Mechanics by W. Chester
Mechanics by W. Chester
MECHANICS
w. CHESTER
University of Bristol
London
GEORGE ALLEN & UNWIN
Boston Sydney
First published in 1979
© W. Chester, 1979
Softcover reprint of the hardcover 1st edition 1979
Chester, W.
Mechanics.
1. Mechanics
1. Title
531 QC125.2 78-41320
Preface page xi
Notation xiv
Units xv
CHAPTER 1. VECTORS
1.1 Introduction 1
1.2 Definitions 2
1.3 The Rules of Vector Algebra 5
1.4 The Resolution of a Vector 10
1.5 The Products of Vectors 14
1.6 The Derivative of a Vector 24
1.7 The Integral of a Vector 30
1.8 Scalar and Vector Fields 32
1.9 Line Integrals 36
1.10 The Gradient Operator 42
1.11 Equipotentials and Field Lines 48
Exercises 48
CHAPTER 2. KINEMATICS
2.1 Introduction 57
2.2 Velocity and Acceleration 58
2.3 Constant Acceleration 63
2.4 Projectiles 64
2.5 Dimensions 71
Exercises 73
CHAPTER 3. DYNAMICS
3.1 Newton's Laws 80
3.2 Units 83
3.3 Applications 83
Exercises 89
CHAPTER 6. OSCILLATIONS
6.1 Simple Harmonic Oscillations 136
6.2 Damped Oscillations 140
6.3 Forced Oscillations '142
6.4 Coupled Oscillations 146
6.5 Non-Linear Oscillations 154
Exercises 161
Index 429
TABLES
4.1.1 Values of I1s for some common substances page 95
7.6.1 Planetary data 201
PREFACE
When I began to write this book, I originally had in mind the needs of university
students in their first year. May aim was to keep the mathematics simple. No
advanced techniques are used and there are no complicated applications. The
emphasis is on an understanding of the basic ideas and problems which require
expertise but do not contribute to this understanding are not discussed. How-
ever, the presentation is more sophisticated than might be considered appropri-
ate for someone with no previous knowledge of the subject so that, although it is
developed from the beginning, some previous acquaintance with the elements of
the subject would be an advantage. In addition, some familiarity with element-
ary calculus is assumed but not with the elementary theory of differential
equations, although knowledge of the latter would again be an advantage.
It is my opinion that mechanics is best introduced through the motion of a
particle, with rigid body problems left until the subject is more fully developed.
However, some experienced mathematicians consider that no introduction is
complete without a discussion of rigid body mechanics. Conventional accounts
of the subject invariably include such a discussion, but with the problems
restricted to two-dimensional ones in the books which claim to be elementary.
The mechanics of rigid bodies is therefore included but there is no separate
discussion of the theory in two dimensions. The argument for a preliminary
treatment of two-dimensional mechanics is that it avoids the introduction of
angular vectors. It is true that the student finds this notion difficult to grasp. But
if he is to understand rigid body mechanics in any real sense, an understanding
of angular vectors is a prerequisite and it is as well to accept this rather than
delay the inevitable. In any case most of the interesting applications are three
dime nsional.
As already stated, the original aim was to produce an elementary text, but the
decision to include three-dimensional rigid body mechanics promotes it to the
intermediate category. Tensor calculus has, however, been omitted. This is not
because I underestimate its importance, but because its inclusion would not be
consistent with the aim to keep the mathematics simple. Furthermore, it is not
usually taught at this level primarily with an eye on its application to mechanics.
It is far more important for the student of mechanics to gain experience in the
application of the theory to specific problems, and to learn from that experience
how to choose by inspection the most appropriate set of axes for a given
problem. On the other hand the book would now be incomplete without some
introduction to Lagrange's equations. Again it seemed inappropriate to give a
full discussion of the theory of analytical mechanics, which requires a knowledge
of variational calculus and which, since it is now taught primarily as a pre-
requisite for quantum mechanics and field theory, requires a different develop-
ment and point of view.
xii Preface
Chapter 1 deals with the elementary theory of vectors. The rather full
treatment is in keeping with the present-day status in most universities, where it
is taught in its own right and for application over a wider field than classical
mechanics. The mechanics of a particle is introduced in Chapters 2-8 inclusive,
which are genuinely elementary in the sense discussed above and which in my
view form a satisfactory introduction to the subject.
Chapter 9 is exclusively devoted to a full discussion of angular vectors,
including the theory of rotating axes, and Chapter 10 deals with the second
prerequisite for an understanding of rigid body mechanics, namely the theory of
moments. Although these preliminaries take some time, the subsequent deri-
vation of the equations of motion is readily achieved and is discussed in Chapter
11. Whereas the derivation of the equations is relatively straightforward, their
application to the many and varied problems can be confusing, and it takes
further time and effort before the student becomes competent at solving such
problems. For this reason the bulk of Chapter 11 is devoted to worked ex-
amples, beginning with simple two-dimensional problems and ending with the
more sophisticated three-dimensional applications.
Virtual work and Lagrange's equations are introduced in Chapter 12 and
finally Chapter 13, which is unconventional for a mechanics text, describes some
important techniques used in non-linear mechanics and elsewhere. My own view
is that these techniques are more useful at the elementary level than the
conventional two-dimensional rigid body mechanics.
lllustrative examples are provided throughout each chapter, and a substantial
number of exercises is given at the end of each chapter. I know of no subject
where the working of examples is more important as an aid to understanding,
and it should occupy the major part of a student's effort.
All practising mathematicians would agree that understanding is helped by a
good notation and hampered by a bad one. I have tried to bear this in mind and
the student is well advised to do likewise. It is a mistake to have a hard and fast
rule, but generally speaking alliterative notation is to be preferred. Thus rn for
mass, rnp for the mass of a particle P, rno and rn 1 for the masses of particles Po
and PI are superior aids to identification than, say, m, Jl, M, M'. This also allows
the symbols Jl and M to be reserved for the coefficient of friction and the
moment of a force respectively. Since it is also desirable to use a notation which
is widely accepted and therefore easily recognised, I have used the recommend-
ations of the British Standards Institution (BS 1991, Part 1, 1967). The notation
is also compatible with that recommended in the report of the symbols commit-
tee of the Royal Society of London (Quantities, Units and Symbols, second
edition, 1975). For similar reasons of wide acceptance, the MKS system of units
is used exclusively.
I am indebted to those authors, too numerous to mention, who have written
on this subject before me and thereby helped to crystallise my own views. It is
also a pleasure to acknowledge the helpful suggestions and advice from W. H. H.
Banks, C. Illingworth, D. W. Moore and M. B. Zaturska. lowe a special debt of
Preface xiii
gratitude to Nancy Thorp for her indispensable help in the typing of the
manuscript and for her tolerant acceptance of my seemingly endless revisions.
Any suggestions for improvement will be welcomed.
W.CHESTER
NOTATION
The following list shows the most common usage of the various symbols.
length metre m
time second s
mass kilogram kg
force newton N
work joule J
power watt W
angle radian rad
1 Vectors
1.1 INTRODUCTION
The first pair of relations are the commutative rules of addition and multi-
plication. The second pair are the associative rules and the third is the distribu-
tive rule. *
A vector requires a direction, as well as a magnitude, for its complete
specification. We also require certain rules of procedure which have a formal
resemb lance to (1.1.1)-( 1.1.3). This is because it is convenient to adapt the
familiar notation of scalar algebra in order to explain and manipulate the rules
appropriate to vector algebra. In doing so, it is important to avoid the assump-
tion that the relations which hold in scalar algebra are automatically valid in
vector algebra simply because of the similarity of notation. While it is true that
the basic rules are a matter of definition when regarded as the basis of a branch
of mathematics, we are particularly concerned in mechanics with the behaviour
and manipulation of physical quantities. Typical examples of vectors in mechan-
ics are velocity, acceleration and force. It is the behaviour of physical quantities
such as these which provides the motivation for vector algebra, and the rules are
formulated with this in mind.
* It is worth noting that the theorems of elementary algebra need not be confined to the
system of real numbers. They can be interpreted to apply to any set of entities which
combine in a way that can be described by the rules of elementary algebra. Conversely
algebras in which the basic rules are different can be, and are, studied.
2 Vectors [1.2
1.2 DEFINITIONS
Free Vector
Let A, B be two points in three-dimensional space. Let C, D be two other points
such that
The class of all such pairs of points (C, D) is defined as the free vector AB. A
particular pair of points such as (A, B) or (C, D) is called a representation of the
vector. Any particular representation is sufficient to determine the vector.
However the vector itself is the class of all its representations and defines a
magnitude and direction, but no particular position in space. In Figure 1.2.1,
(A, B), (C, D) and (E, F) are all representations of the same vector AB or, more
briefly, a.
We say that two vectors a, b are equal if they refer to the same class and we
then write a = b. Thus in Figure 1.2.1 we have*
AB = CD = EF = a.
Note that if two vectors are unequal they will have no representation in
common. Hence if a representation of a vector a has the same magnitude and
direction as a representation of a vector b, one can infer that a = b. This should
be borne in mind in the subsequent development of the theory, for in the actual
manipulation of vectors we are constantly dealing with their representations as
* There are occasions when it is convenient to restrict a vector to a given line; it is then
called a line vector. In Figure 1.2.1, for example, (A, B) and (E, F) are representations of
the same line vector AB or EF, which is to be distinguished from the line vector CD. This
distinction will not, however, be used here.
1.2] Definitions 3
BA=-AB,
so that the negative sign reverses the direction of a vector without changing its
magnitude.
If A is a scalar, then Aa defines a vector of magnitude I A II a I. Its direction is
the direction of a if f... >0 and the direction of -a if f... <0. If eitherf... = 0 or
I a I = 0, it is a null vector.
Null Vector
The class of all pairs of coincident points (A, A) is called a null vector or zero
vector. It has zero magnitude and no specific direction associated with it. It is
comparable to the zero of scalar algebra although they are not identical con-
cepts. If a is a null vector we write a = o.
Position Vector
If the point 0 is fixed, OA is called the position vector of the point A relative to
O. Once the point 0 is chosen, the representation of any position vector relative
to 0 is fixed, and other representations cease to be appropriate. If two points
have the same position vector, then the points coincide.
The position vector will be used in the subsequent analysis as a useful way of
specifying a point relative to a chosen origin, that is as an alternative to its
specification by means of its coordinates relative to a chosen set of axes.
No special notation is used to distinguish a position vector and a correspond-
ing free vector. It should be clear from the context which is intended, and where
no restriction is imposed it is the free vector which is being discussed. In
particular, the actual algebra of vectors, with which we are concerned in this
chapter, refers to free vectors.
Let (0, A), (0, B) be representations of the vectors OA, OB respectively. The
angle between OA and OB is defined as AOB, with 0 < AOB < 11'.
This definition requires a choice of representations having a common point.
This is always possible for free vectors, even if initially they are determined by
4 Vectors [1.2
A'
Figure 1.2.2 The angle between vectors.
representations which do not satisfy this condition. The angle is clearly inde-
pendent of the particular representations chosen to define it.
We note the follOwing properties:
(i) The angle between two vectors depends on their directions and not on their
magnitudes.
(ii) The angle between AO and BO is the same as the angle between OA' and
OB', where OA' = AO, OB' = BO. Thus it is A'OB' = AOB and so is the
same as the angle between OA and OB (see Fig. 1.2.2).
(iii) The angle between OA and BO is the same as the angle between OA and
OB' and so is AOB' = 1T - AOB.
Coplanar Vectors
If the directions of several vectors are all parallel to a plane, the vectors are said
to be coplanar. The plane of the vectors is any plane parallel to their directions.
Vector Addition
The sum, or resultant, of two vectors AB and BC is defined to be the vector AC,
as in Figure 1.2.3. We write
In particular
- ~ -
AB + BA = AA = O. (1.2.2.)
a + (-b) = a - b,
1.3] The Rules of Vector Algebra 5
a+b
A a
Figure 1.2.3 The sum of two vectors 3 and b.
and this serves to define the difference of two vectors. Thus, from Figure 1.2.3,
we have
(1.2.3)
Comparison of (1.2.1) and (1.2.3) shows that, as in scalar algebra, vectors can
be transferred from one side of an equation to the other with a change of sign.
An immediate consequence of the rule of addition is the inequality
1a + b 1< 1a 1+ 1b I, (l.2.4)
We are now in a position to deduce the following rules, which are a consequence
of the above definitions.
This is made clear in Figure 1.3.1. Let (A, B) and (D, C) be two representations
of a. Let (A, D) and (B, C) be two representations of b. Then ABCD is a
6 Vectors [1.3
k?7
A
Figure 1.3.1
a B
C
parallelogram and
a + b = AB + BC = AC = AD + DC = b + a.
In Figure 1.3.2 let (A, B), (B, C), (C, D) be representations of a, band c
respectively. By successive application of the rule for addition we have
This rule follows from the fact that if a, b, (a + b) are represented by the sides of
a triangle, then Aa, Ab, A(a + b) can be represented by the sides of a similar
triangle, which can then be used to define the sum of Aa and Ab.
a
b
(vi) a + 0= a. (1.3.6)
(vii) 1a = a. (1.3.7)
The last three relations are merely statements of the properties of the null
vector, the unit scalar, and the vector -a. They are included in order to display
the relations which are used as the basis of the theory of linear vector spaces. In
this more general theory the quantities a, b, c ... are not further defined beyond
the requirement that they satisfy relations (1.3.1)-(1.3.8) above. The con-
sequences of such a theory are quite versatile in that they can be applied to any
set of quantities which obey the above rules. An elementary example is the set
of real numbers; a less obvious example is the set of all polynomials with real
coefficients. It is therefore worth bearing in mind that although we are con-
cerned here with the development of a theory for geometric vectors as defined in
§ 1.2, many results are capable of a more general interpretation in so far as they
depend simply on a manipulation of relations (1.3.1)-(1.3.8). It should not,
however, be assumed that vector algebra is appropriate in all applications where
a physical concept is described by a magnitude and direction. One exception is
the finite rotation of a rigid body; this has a magnitude, the angular displace-
ment, and a direction defined by the axis of rotation. It does not, however,
satisfy the rules of addition for vector algebra.*
Whenever a branch of mathematics is applied to a physical problem, there is
an assumption that the physical quantities involved obey the rules of the
mathematics used in their manipulation. The only justification for this assump-
tion depends on the extent to which it leads to profitable conclusions, which can
be satisfactorily confirmed by observation.
Example 1.3.1
Let r be the position vector of a point P relative to an origin O. Let a and b be
constant vectors, and let A be a variable scalar. We consider the locus of P when r
* For example let us define two perpendicular axes OX and OY and consider the
displacement of a rod lying originally along Ox. If it is given first a rotation through an
angle 11/2 about OY, so that it is perpendicular to both OX and OY, and then a rotation 11/2
about Ox, its fmal position will be along OY. If, however, the rotations are carried out in
the reverse order, that is first a rotation 11/2 about OX leaving the rod along OX, and then a
rotation 11/2 about OY, the final position of the rod will be perpendicular to OX and OY.
Thus the commutative law of addition is violated.
8 Vectors [1.3
o __------~r------~p
Ab
(i) (ii)
Example 1.3.2
Let the points A, G, B in Figure 1.3.4 lie on a straight line, in that order, and
such that
mAG=nGB,
mAG=nGB,
since mAG and nGB have the same magnitude and the same direction. Hence
m(AO + OG) = n(GO + OB),
or, by rearrangement,
This result is not restricted to points G which lie between A and B. 'The
essential condition is that mAG =nGB should be satisfied. This can include
points G on the extensions of AB provided negative values of min are allowed.
Example 1.3.3
Vector algebra is particularly useful in the derivation of certain results in
Euclidean geometry. The following example proves, among other things, that the
medians of a triangle (that is the lines joining the vertices to the mid-points of
the opposite sides) intersect at a single point, which is called the centroid of the
triangle.
In Figure 1.3.5. let G be a point on the median AD of a triangle ABC such
that
3AG=2AD. (1.3.10)
Show that
(i) (1.3.11)
c
B
Hence
It has already been shown that a single resultant of several given vectors can be
dermed, and we now consider the converse process of resolving a given vector
into several constituent vectors. We first introduce another definition.
If there is a linear combination of three vectors which is zero, the vectors are
said to be linearly dependent. In other words a, band c are linearly dependent
if there are suitable scalars X, p., v, not all zero, such that
Xa + p.b + vc:::: o.
Conversely, if the relation is satisfied only when X:::: p.:::: v:::: 0, the vectors are
said to be linearly independent.
Physically speaking the distinction is one between coplanar and non-coplanar
1.4 ] The Resolution of a Vector 11
vectors, apart from certain degenerate cases which are not exceptional if we
accept the convention that any two vectors and a null vector are coplanar. The
relation Aa + Ilb + vc = 0 implies that Aa, Ilb, vc are coplanar since -Xa is the
resultant of Ilb and vc. If a, b and c are not coplanar there is a contradiction
unless A =Il =v =O. For if, say, A:;6 0 we have -a = A-Illb + A-I vc which
implies that a is coplanar with band c.
Theorem
Any vector in three dimensions is equal to a linear combination of any three
given linearly independent vectors.
Let (0, P) be a representation of r, and let Ox, Oy, Oz define the directions
of the three given vectors (see Fig. 1.4.1).* It is possible to construct a
parallelepiped whose adjacent edges OX, OY, OZ are in the given directions and
whose diagonal is OP. First construct QP parallel to Oz so that Q lies in the plane
of Ox, Oy. Then construct XQ parallel to Oy so that X lies on Ox. The vectors
OX, XQ, QP define the parallelepiped and
r = OX + XQ + QP = OX + OY + OZ. (1.4.1)
This construction is always possible provided that the directions Ox, Oy, Oz are
not coplanar.
We say that OX, OY, OZ are the resolutes of r in the three given directions.
If a, b, c are the three given vectors whose directions are defined by Ox, Oy,
Oz respectively, it is possible to choose scalars A, Il, v such that
.,----------:;;iII p
°
Figure 1.4.1 The resolutes of r parallel to Ox, Oy, Oz.
* In spite of the notation, the results of this section are not confIrmed to the position
vector. They apply to the free vector r, of which (0, P) is a representation based on 0.
12 Vectors [ 1.4
(1.4.4)
we have (1.4.5)
and so (1.4.6)
It follows that if two vectors are equal their resolutes are equal. This result can
be useful in demonstrating vector identities.
Consider now the vector
(1.4.7)
where (1.4.8)
and this dermes, uniquely, the resolutes of the resultant f of several vectors fi. It
shows that the sum of the resolutes in a given direction is the resolute of the
resultant.
A most important application of the above results is to the situation in which
Ox, Oy, Oz form a mutually orthogonal right-handed set of coordinate axes.*
Let OP = f and let P have coordinates (x, y, z) relative to such axes. Let i,j, k be
vectors of unit magnitude in the directions Ox, Oy, Oz respectively. Then
We call x,y, z the components of f in the directions Ox, Oy, Oz. More generally
the component of f in any direction is the projection of its representation on to
a line in that direction, associated with the appropriate sign.
In Figure 1.4.2 let
* A right-handed orthogonal set of axes Oxyz is such that the rotation of a right-handed
screw through an angle nl2 from Oy to Oz, or Oz to Ox, or Ox to Oy propels it along the
positive direction of the third axis. A left-handed set is obtained if the direction of one of
the axes is reversed. It is conventional to choose a right-handed set. Such axes are also
described as cartesian or rectangular.
1.4] The Resolution of a Vector 13
_-------....p (x,y, z)
Z f----+----;/---{
o x
Figure 1.4.2 Cartesian components of r.
(1.4.14)
We call (I, m, n) the direction cosines of the line OP, relative to the given axes.
As a special case of (1.4.9) we note that the sum of the components, in a
given direction, of a set of vectors is equal to the component of their resultant.
That is
( .i:;Xi)i+(.fYi)j+(.fzi)k=.fri=r=Xi+yj+Zk (1.4.15)
1=1 1=1 1=1 1=1
s s
LXi =x, L Yi =y,
i= 1 i=1
Example 1.4.1
Find the vector AB, where A has coordinates (1, 2, 3) and B has coordinates
(4,5,6).
14 Vectors [1.5
Here OA = i + 2j + 3k,
OB = 4i + 5j + 6k,
and AB=OB-OA
= 3i + 3j + 3k.
Note that the radius vector to the point with coordinates (3,3,3) has the
same magnitude as AB and is parallel (but not identical) to the line segment AB.
It is the representation of AB that starts at the origin of coordinates.
Scalar Product
The scalar product of two vectors is defined as the product of the magnitudes of
the vectors and the cosine of the angle 0(0';;; 0 .;;; 1T) between the vectors (see
Fig. 1.5.1). It is written
a. b = ab cos 0 (1.5.1 )
This follows at once when X;;" O. It is also true when X< 0 for then
j.k=k.i=Lj=O. (1.5.5)
(vi) a . (b + c) = a . b + a . c. (1.5.7)
For
Note, however, that in general (a. b)c * a(b . c) and that a. b. c is not
defined.
Vector Product
The second kind of product is a vector whose magnitude is defined as the
product of the magnitudes of the vectors and the sine of the angle e(O ~ e ~ Tr)
between them. The direction is perpendicular to the plane of the two vectors in
the sense of the translation of a right-handed screw rotated from the first to the
second vector. Note that the order ofthe vectors is now relevant. The notation is
a x b =ab sin en
where, in Figure 1.5.1, n is a unit vector perpendicular to the plane of the figure
and into (rather than out of) this plane. We note the follOWing properties:
(i) Ifax b = 0 then one (or more) of a, b and sin e is zero and vice versa.
Thus a x a = 0 and in particular
i xi =j xj =k x k = O. (1.5.9)
i = j x k, j = k x i, k = i x j. (1.5.10)
(iii) Some care is necessary in the handling of this product since the commuta-
tive and associative laws are not satisfied. In fact, one can see from the
defmition that
a x b = - b x a.
Also by counter example it can be seen that, in general,
(a x b) x c * a x (b x c).
For example (i x i) xj = 0 but i x (i xj) = i x k = - j.
ax(b+c)=axb+axc, (1.5.12)
are, however, satisfied. The first relation (1.5.11) follows from the defini-
tion, with a little care necessary if A is negative.
As a preliminary to the proof of (1.5.12) we note that if a =0 the
*
relation is clearly satisfied. When a 0 it is sufficient to prove the result for
a = 1, since for other values each side of the equation can be reduced by the
factor a-I.
1.5] The Products of Vectors 17
or
a x b + a x c = a x (b + c).
a b+c
P lL~--"tb----ll Q 1
1 1
1----------1 I---~
!~ ------t-1~
p'
Q'
a. (b x c) = (b x c). a = - a. (c x b).
a . (b x c) = b. (c x a) = c. (a x b). (1.5.14)
which is the x component of the left-hand side. The other components are dealt
with in similar fashion.
1.5] The Products of Vectors 19
Example 1.5.1
We use the scalar product to establish the cosine rule for a triangle, that is, if
BC =a, CA = b, AB = c, then
(1.5.17)
Since
Note that BAC is the angle between AB and AC; the angle between BA and AC
is 1T - BAC (see Fig. 1.5.3).
,/
,/
/
,/
A,/ n-BAC
c b
B~---------+----------~C
a
Example 1.5.2
We prove that the altitudes of a triangle are concurrent.
Let the altitudes through B and C, of the triangle ABC, meet in H (see Fig.
1.5.4). Then
AH . CA + AH . AB = 0 or AH. CB=O
since BA . CA + CA . AB = O.
20 Vectors [1.5
B C
Figure 1.5.4 Illustration for Example 1.5.2.
Example 1.5.3
We use the vector product to prove the sine rule for a triangle, namely that in
any triangle ABC with sides a, b, c
Then - c=a+b,
This is equivalent to
b x c = c x a,
b x c = c x a = a x b.
Example 1.5.4
Let (J" (J2, (J3 be the angles between the vectors r2 and r3, r3 and r" r, and r2
respectively. We consider the problem of finding the angle between r3 and
r, x r2' Geometrically this is the angle between r3 and the perpendicular to the
plane of r, and r2' If this angle is cp, we have that
1r3 x (rl x r2) 12 = (r3 . r2)2ri + (r3 . rlir~ - 2(r3 . r2)(r3 . rl)(rl . r2)
= rir~r~(cos2 (Jl + cos 2 (J2 - 2 cos (Jl cos (J2 cos (J3)'
It follows that
Example 1.5.5
We have already seen, in § 1.4, that
r=Xa +JLb+vc
r. (b x c) = Aa . (b x c).
Similarly
It follows that
There are certain reciprocal relations satisfied between the two sets of vectors
a, b, c and A, B, C. It follows from their defmition that
a.A=b.B=c.C=I,
(1.5.22)
a . B = a . C =: b . C = b . A = c . A = c . B = 0,
(cxa)x(axb) {(cxa).b}a a
BxC= = =--- (1.5.23)
{a. (b x C)}2 {a. (b x C)}2 a. (b x c)'
1
A . (B x C) = (b ). (1.5.24)
a. xc
Hence
BxC b= CxA AxB
a= c= . (1.5.25)
A. (B x C)' A. (B x C)' A. (B x C)
This last relation shows that any vector r is specified by its scalar products with
three linearly independent vectors. A corollary of this result is that if the scalar
products of r with three linearly independent vectors are zero, then r is a null
vector.
A useful expression for the magnitude of r can be deduced from the use of
1.5] The Products of Vectors 23
r2 = {(r. A)a + (r. B)b + (r. C)c} . {(r. a)A + (r. b)B + (r . c)C}
= (r. A)(r. a) + (r. B)(r. b) + (r. C)(r . c).
Example 1.5.6
Sometimes a vector is defmed only through a vector equation. We take as an
example the equation
ax +x xa=b, (1.5.27)
(1.5.28)
aa. x =a. b.
x x a=b.
x = a -2 (a x b) + A.a.
24 Vectors [1.6
provided such a limit exists. Let us choose a representation of ret) based on some
origin 0, so that ret) is the position vector of some point P relative to O. This is
not essential to the above defmition, but it does allow us to give it a geometrical
interpretation. In Figure 1.6.1 we have
Figure 1.6.1 The locus of a point P whose position vector r(t) varies con-
tinuously with t.
1.6] The Derivative of a Vector 25
where pp' refers to the length of the chord PP'. We note that PP' /PP represents a
I
unit vector and that, as Dt ~ 0, the direction of this unit vector tends to coincide
with that of the tangent at P. Also
lim pp' = ds
(1.6.4)
6t-O ot dt'
where s is the arc length. This is so because
PP' = os(1 + e), (1.6.5)
where os is the arc length from P to pi, and e ~ 0 as os ~ O. In fact one can
introduce the notion of arc length along a curve as the function whose derivative
is defined by (1.6.4), the arc length itself then being obtained by integration of
the resulting function. For future reference we derive the following result from
this definition. Let
ret) = x(t)i + y(t)j + z(t)k
so that x(t), yet), z(t) are the coordinates of P, and
(PP')2 = {x(t + ot) - X(t)}2 + {yet + ot) - y(t)}2 + {z(t + ot) - z(tW.
(1.6.6)
When substituted in (1.6.4) this gives
ds = lim {(X(t + ot) - X(t))2 + (y(t + ot) - y(t))2 + (z(t + Or) - z(t) )2}1/2
dt 6t-0 ot ot ot
(1.6.7)
d dr dA
dt (Ar) = A dt + dt r, (1.6.9)
d dr2 drl
dt (rl . r2) = rl ' Tt + dt' r2, (1.6.10)
d dr2 drl
- (rl x r2) = rl x - + - x r2' (1.6.11 )
dt dt dt
26 Vectors [1.6
The proofs of these results are straightforward generalisations of the proofs used
for differentiation involving ordinary functions. For example,
and so
In the final relation (1.6.11) it should be noted that the terms in the vector
products are not commutative.
One important application of the rules allows us to express the derivative of a
vector in terms of its components relative to three fIXed directions. Thus if
dr dA dtL dv
- =- a + - b + - c. (1.6.13)
dt dt dt dt
In particular, if
dr dx. dy. dz
-=-l+-J+-k. (1.6.15)
dt dt dt dt
Q ""'.,.-------.... ...........
/ / ' ,,
I
/
/
,,
\
P
\
\
\
\
~---'-_ _ _----II_'x
I
I
Figure 1.6.2 Perpendicular unit vectors a and b.
unit circle at P. Since the arc length here is just the angular distance 0, measured
in radians from some initial line, we have the result that
(1.6.16)
db d dO
- =-(0 + rr/2)(-a) =- -a. (1.6.17)
dt dt dt
Alternatively one can choose fIxed unit vectors i and j along and perpendicular
to the initial line, so that
da . dO . dO. dO
dt = - sm 0 dt 1 + cos 0 dt J = dt b, (1.6.20)
db dO.. dO. dO
- = - cos 0 - 1 - sm 0 - J = - - a (1.6.21)
dt dt dt dt'
as before.
The above results can be used to calculate the derivative dr/dt of a vector ret)
which always remains parallel to a fIxed plane. Let a, b be unit vectors along and
28 Vectors [1.6
Figure 1.6.3 Cartesian coordinates (x, y) and polar coordinates (r, 0).
perpendicular to r and let the direction of r make an angle 0 with a fIxed axis, as
shown in Figure 1.6.3. We have
r = ra, (1.6.22)
dr dr da dr dO
-=-a+r-=-a+r-b. (1.6.23)
dt dt dt dt dt
Also d2 r = d2 r a + dr da + dr dO b + r d2 0 b + r dO db
dt 2 dt 2 dt dt dt dt dt 2 dt dt
d2 r dr dO dr dO d20 (dO )2
=-a+--b+--b+r-b-r - a
dr dt dt dt dt dt 2 dt
(1.6.24)
Equations (1.6.23) and (1.6.24) give the components of dr/dt and d2 r/dt 2
along and perpendicular to r. In some applications these components can be
more useful than the expressions obtained using cartesian components, namely,
Example 1.6.1
If rand r' are two parallel vectors, the following relation holds:
,dr ,dr
r .-=r-.
dt dt
For if r and r' are parallel vectors we can write r' = A(t)r for some scalar variable
A(t). It is then sufficient to prove that
dr dr
r.-=r-.
dt dt
dr dr dr dr
to get r . - + - . r = 2r . - = 2r - .
dt dt dt dt
Example 1.6.2
If a, b, C are three mutually perpendicular unit vectors, their derivatives da/dt,
db/dt, dc/dt are linearly dependent. To demonstrate this, it is sufficient to show
that
da . (db x dC) = o.
dt dt dt
db dc
b .-=c.-=O.
dt dt
30 Vectors [1.7
Po
Figure 1.7.1 A typical increment Orj = fi - rj-l of a continuously varying
vector r.
where €; -+ 0 as ot; -+ 0 for all i such that 1 < i < n. Now with the help of a
generalisation of (1.2.4), namely
we have (1.7.6)
where €m is the maximum value of €i for 1 < i < n. Hence, as all the Of i -+ 0, so
does €m(tn - to) and therefore
n
rn - ro = lim ~ v;(tnot;. (1.7.7)
ot;-O;=1
When this limit exists, and is independent of the way of proceeding to the limit,
it is called the Riemann integral of v and we write
= Jt
n n
rn - ro = lim ~ v;(tnot; vet) dt. (1.7.8)
6t;-0 ;= 1 to
Example 1. 7.1
If d is a constant vector we have
r('dt
J i -dr + j ::::..r
x dr + k -dr
dt dt
z)
dt = ~dr
- dt = r + c.
dt
r 1
Ja dt =- (i sin wt - j cos wt) + c =- - b + c,
1
w w
r 1
Jb dt = - (i cos wt + j sin wt) + c' = - a + c'.
1
w w
32 Vectors [1.8
Example 1. 7.3
r x -d22r dt
Jr = Jr-d ( r x -dr) dt =r x -dr + C.
dt dt dt dt
and (ax, ay, az ) respectively, then the relation IjJ = constant can be written
and this is the usual equation for a plane in coordinate geometry. In particular, if
the axes are chosen so that ay = az = 0, then a is parallel to the x-axis and the
surfaces IjJ = constant are equivalent to the surfaces x = constant. These are
clearly planes perpendicular to the x-axis.
With a vector function F(r) we are interested in the direction of F as well as
the magnitude. In many applications the direction is of particular interest and is
displayed by a family of lines for which the tangent at each point is in the
direction of F. Such lines are called field lines. Note that the vector function
ljJ(r)F(r) has the same field lines as F(r) for any scalar function ljJ(r) which is not
zero, since IjJ affects only the magnitude of the vector. For example, the vector
function ljJ(r)r is always in the direction of the radius vector, and any straight
line through the origin is a field line.
If F = FAx,y)i + Fy(x,y)j, then the inclination of F at any point is Fy/Fx
and so the field lines in two dimensions are represented by the integral of the
equation
(1.8.1)
dy =x 2
dx y
which integrates to
The family of field lines is obtained by plotting these curves for different values
of the constant (see Fig. 1.8.1).
The extension to three dimensions is straightforward and is most con-
veniently described parametrically. We suppose that the position vector r is itself
defined by a parameter t so that, as t varies, ret) defines a point P which moves
along a prescribed path. At all points on this path dr/dt is parallel to the tangent
(see § 1.6). Now the tangent to the field lines of F(r) is everywhere parallel to
F(r), and this is satisfied by the locus of P if
dr
-= F(r). (1.8.2)
dt
34 Vectors [1.8
Figure 1.8.1 The family of curves if> = h 2 - ix 3 = constant for the scalar field
if>. The curves also represent the field lines of the vector. field F = yi + x 2 j. The
arrows indicate the direction of F.
When F(r) is known as a function of r, equation (1.8.2) defines the slope of the
field lines at each point. In principle, this is sufficient information to enable the
field lines to be constructed. In practice it entails the integration of the set of
equations
dx dz
dt = Fx(x,y, z), - = Fz(x,y, z) (1.8.3)
dt
dr
dt = if>(r)F(r),
with an appropriate choice for if>(r), for the purpose of calculating the field lines.
1.8] Scalar and Vector Fields 35
Example 1.8.1
For the vector function a Xl, where a is a constant vector, the field lines are
everywhere perpendicular to a and l. Hence they are circles in planes perpen-
dicular to a and with a as the common axis.
To derive the result formally, the analysis is simplified if the x axis is chosen
to coincide with the direction of a, so that a = ai. Then a x r has the components
a(O, -z,y) and the field lines are given by
dx dy
-=0 -=-az
dt ' dt '
dy
Y -+z=O
dz '
or y2 + Z2 = constant,
the field lines are the circles
x = constant, y2 + Z2 = constant.
Example 1.8.2
For the vector function
dx dy dz
dt = yz, -=zx dt = -xy.
dt '
dy dz dz dx
Y -+z ·--=0, z-+x-=O.
dt dt dt dt
the field lines are given by the lines of intersection of two families of circular
cylinders, whose axes coincide with the x-axis and the y-axis respectively.
When the scalar field 4>(r) is defined as a function of the vector r, and r is itself
defmed as a function of the scalar variable t, cp is given implicitly as a function of
t and we can operate on it as such. If ret) is interpreted as the position vector of
a point P which describes a curve as t varies, we are in effect defining cp(r) at
points ofthis curve. For this reason the integral
Jcp(r(t)) dt (1.9.1)
is called a line integral. Indeed the parameter t may, in some applications, be just
the arc length along this curve from a given origin. In practice it is evaluated in
the usual way by expressing cp explicitly as a function of t. When r is defined in
terms of its components, we can also express (1.9.1) in the equivalent form
rI
then J cp dt =
o
J xyz dt =)
0
I
0
I
t 6 dt =~ . (1.9.4)
(1.9.5)
where the parameter is one of the variables defining the coordinates. This
integral can be evaluated either by expressing y and z as functions of x to get
Y =X 2/3 , z =X l/3
1 1 1
and so ~
o
t/> dx = ~
0
xyz dx = J0 X2 dx =l. (1.9.8)
1 1 dx 1
i t/> dx = i xyz - dt = i 3t 8 dt = ! . (1.9.9)
)0 )0 dt )0
L t/>(x,y, z) dx (1.9.11)
is often used, where C denotes the curve involved along which the integration is
taken. Only when t/> is given as a function of x alone, independent of y and z,
can (1.9.11) be integrated without further information.
The generalisation to the line integral of a vector
IF(r(t)) dt (1.9.12)
F = Xi + Yj + Zk, (1.9.13)
r dr
l(r). dt dt, (1.9.15)
~ F. dr. (1.9.16)
dr dx. dy. dz
-=-I+-j+-k (1.9.17)
dt dt dt dt
r(1)
Figure 1.9.1 The vector F(r) associated with the point P whose position vector
is r.
1.9] Line Integrals 39
which is suggested by the shorthand form of (1.9.16) used here on the left-hand
side of (1.9.19). Whether or not the different contributions are actually evaluated
with the above choices for t, (1.9.19) is the usual expression used for the work
in terms of the components of F.
We repeat the warning here that, in general, expressions of the type (1.9.19)
cannot be integrated until the curve of integration is specified. One obvious
exception is the integral
Example 1. 9.1
Evaluate the line integral
} xy2 dy,
e
The <integral can now be evaluated with the substitutiony = sin 8. But it is more
direct to make the transformation
x=cos8, y=sin8
in the original integral, and integrate directly with respect to the parameter 8 .
Then
Jc 2
xy2 dy = cos 8 sin 8 d(sin 8) = J
e O
r 21T 2 2
cos 8 sin 8 d8
21T
= i) o (1 - cos 48) d8 = 1T/4.
Example 1.9.2
Evaluate the line integral
(n ~O)
= t 2 + 1.
J xn ds
- dt =
) t, n t7.+ 3 t n+1
t (t2 + 1) dt = _ _ + _1- .
c dt 0 n +3 n +1
Example 1.9.3
We consider the work of F = y2 i + x 2j along the following (two-dimensional)
curves between the points (0, 0) and (1,2) in Figure 1.9.2, namely:
(i) The two straight lines between (0,0) and (1,0) and between (1,0) and
(1,2).
We evaluate the integral in two parts. Along Ox we can take the
parameter to be x itself and write
o x
since dx/dy =0 and x = 1 along this path. Thus for the integral from (0, 0)
to (1,2) we have
JF. dr = 0 + 2 = 2.
(ii) The straight line joining the points (0, 0) and (1, 2). On this line y = 2x and
so, with x as the parameter, we get
Example 1.9.4
Evaluate
~ F. dr dt
,c dt
42 Vectors [1.10
r = a cos t i + b sin t j + et k.
Since -dr = -a sm
. t °
1 + b cos t J0+ e k
dt
we have
Hence
Hence
r dr (d
JF . - dt = J - {(x 2 + z)y + !z3} dt =(x 2 + z)y + !Z3 + constant.
dt dt
This agrees with the previous calculation, but this alternative way of proceeding
does not use any information concerning the particular path of integration, and
the result is true whatever path of integration is chosen. We shall see eventually
under what conditions this alternative procedure is possible.
We have seen that the work of a vector mayor may not depend on the path of
integration. When it does not, the vector field is called conservative, and such
fields play an important role in applied mathematics. It is useful to consider the
conditions which a vector field must satisfy in order that it should be conserva-
1.10] The Gradient Operator 43
and Similarly for the other variables, the above relation for dcp/dt can be
expressed in the form
" cp(x,y,z+Sz)-cp(x,y,z)5z
+ I~ -.
6t ..... O Sz St
where acp/ax, acp/ay and acp/az are the partial derivatives of cp with respect to
X,Y and z respectively. Integration of equation (1.10.2) gives the indefinite
integral expression for cp,
integration C, the value of the integral between two points of C depends only on
the end points, and not on the curve joining them. This must clearly be so since
the integral represents the change in cp between the two end points, and cp is a
function only of position. *
A comparison of (1.10.4) with the expression (1.9.19) for the work of a
vector F shows that if F is of the form
(LlO.5)
for some scalar function cp, then the work of F depends only on the end points
and not on the path. Hence F is a conservative vector field.
The converse statement is also true, that if
f (X dx + Y dy +Z dz) (1.10.6)
depends only on the end points and not on the path of integration, then a
function cp exists for which
x=acp/ax, Y=acp/ay, z=acp/az.
For if we choose an origin and consider all line integrals starting at the origin,
the integral (1.1 0.6) depends only on the end point (x, y, z) and so defines a
function
cp(x,y, z) = f (X dx + Y dy + Z dz) = fo
t (dx dy dz)
X dt + Y dt + Z dt dt
(1.10.7)
say, when the integral is taken along some path r = ret). From (1.10.2) and
(Ll 0.7) we have
But this must hold for all paths of integration and so the relation is true
whatever the dependence of x, y, anc! z on t. This is only possible if
For example, the first relation follows from (1.10.8) with dy/dt = dz/dt = 0 and
dx/dt =t= O.
The scalar function cp is called the potential of the vector field F. The vector
* More correctly", must be a single valued function of position. We shall consider only
such functions here.
1.10] The Gradient Operator 45
represented by (1.10.5) is called the gradient of cp, or briefly grad cp, and is
written
(1.10.10)
The operator a .a
V'=i-+J-+k-
a (1.10.11)
ax ay az
has many of the properties of vectors. It can also occur in expressions of the type
ax ay az
V'.F=-+-+-, (1.10.12)
ax ay az
called the divergence of F, or div F. The vector product
Y= acp/oy, z = ocp/oz,
it follows by direct substitution that
oY ax a2cp a2¢
---=-----=0
ax oy oxay oyax
and similarly for the other components of curl F. Thus for a conservative vector
field we must have
v x F = O. (1.10.14)
r
cp(x,y,z) = 1o X(~,O,O)d~+}
x Y
0
Y(x, 11,0) dl1 + J
0
Z
Z(x,y,nd~,
(LlO.15)
where X, Y, Z are the components of F. Note that I/J is defined as a line integral
of F with the path of integration chosen as the three straight lines joining the
points (0,0,0), (x,O,O), (x,y,O) and (x,y,z). Hence the path of the line
integral is specially chosen, but nevertheless clearly defmes a function of x, y, z.
46 Vectors [1.10
We now wish to show that the function q, so defined is the potential of the
vector field F =(X, Y,Z) whenever V x F =O. This will also show incidentally
that, although a special path of integration has been chosen initially to define q"
the integral is in fact independent of the path of integration since F is a
conservative field.
By differentiation of (1.1 0.15), we see that
J: a~
ax ay az
V'q, = (X(X, 0, 0) + I Y
aa X(x, 11, 0) d11 + f Z ~ X(x,y, t) d t ) i
I:
'0 11 0 at
Example 1.10.1
Show that the vector field f(r)r/r is conservative and find the potential.
We first check that the curl of this vector field is zero. Since
fer) (.
fer) r = -r-
-r- . k) = X'I + YJ. + Z k ,
XI + YJ + z
1.10] The Gradient Operator 47
say, we have
Now
or or
and so r oy =y, r-=z.
oz
or or
It follows that z-=y-
oy oz
and similarly for the two remaining components of the curl of this vector field.
There is therefore a potential function I/> such that
q,= r f(~) d~ +c
r
and, since the constant will have no effect on the gradient of q" we have
For a conservative vector field F =ViP, the field lines are everywhere orthogonal
to the equipotential surfaces q, =constant.
To prove this result, consider the locus of a point P whose position vector
r(t) is defined as a function of the parameter t. If this curve lies on the equi-
potential surface cp = constant, we have
EXERCISES
(i) for any scalar A, Af(a) = f(Aa), and, in particular, -f(a) =f( -a);
(ti) f(a) + f(b) = f(a + b).
AB =band AC = c,
1.4 Let a, b, c, d be fixed vectors, and let A, fJ., v be variable scalars. Let r be
the position vector of a point P. Show that the loci of P represented by
r = a + fJ.b, r = c + vd
AI b +c
-=--
ID a
Show that
BD c
(i) -=-
DC b'
1.8 The three vectors OA, OB, OC are linearly independent, and the three
vectors AD, BD, CD are linearly dependent. Show that
where a: + /3 + 'Y = 1.
1.9 Find the angle between the two non-zero vectors a and b in each of the
following cases:
(i) 2a + 3b = 0,
(ii) lal=lbl=la+bl,
(iii) a. (a + b) = 0 and I b I = 2 I a I.
(iv) (a x b) x a = b x (b x a).
(i) a. k = 1,
(ii)b.k=l,
(iii) a is perpendicular to b,
(iv) a is perpendicular to i + j + k,
(v) b is perpendicular to i + j + k,
(vi) b is perpendicular to i,
(i) a x (b x c) + b x (c x a) + c x (a x b) = 0,
(ii) (a x b) . (c x d) + (b x c) . (a x d) + (c x a) . (b x d) = O.
a2 r = (a. r)a + (a x r) x a,
1.16 Find a unit vector which is at right angles to both of the vectors i - 3j + k
and 2i+j -k.
(Ans.: 62- 1 / 2 (2i +3j +7k.)
1.17 The three vectors a, b, c are non-zero and satisfy the relation
a x (b x c) =b x (a x c).
Show that one possible relation between them is that a and bare
parallel, and fmd a second possibility. Show that the magnitudes of the
vector a x (b x c) in the two cases are lallb x cl and la . bllcl respectively.
1.18 Prove that the four points with position vectors
are coplanar.
1.19 The points A, B, e lie on the surface of a sphere of unit radius and centre
O. The angles BOe, e~A, AOB are denoted by a, {3, 1 respectively. Show
that
OB . oe = cos a, I OB x oe I = sin a.
Show also that the angle between the planes OAB and OAe is
cos
-1 (COS a - cos (3 cos 1) .
sin {3 sin 1
1.20 The two triplets a, b, c and i,j, k are mutually orthogonal right-handed
sets of unit vectors such that
a = 111i + 112j + 113k,
b = 121i + 122j + 123k,
c = 131i + 132j + 133k.
if a = {3,
Show that
if a:f={3.
ih = {3,
if a:f={3.
1.21 The two triplets a, b, c and i, j, k are mutually orthogonal sets of unit
vectors such that
Ii x a + mj x b + nk x c = O.
52 Vectors
x={(a. b)a+b}/(l +a 2 ).
(i) 1 + a . b =f= 0,
(ii) 1 + a . b = O.
ax + a x x + (x . b)a = a
ax + (x . b)a = a,
and use this result to find x given that a =f= 0 and a + a . b =f= O.
(Ans.: (a + a . b)x = a.)
1.25 The position vector of a moving point P with respect to a fixed origin is r,
and the unit vector in the direction of r is a. Show that
(r x i) x r = r3 a.
1.26 The position vectors of the two points A and B relative to an origin 0 are
a(t) and b(t) respectively. Given that
db da
ax-=bx-,
dt dt
d
show that -(axb)=O.
dt
Hence show that the two points A and B move in a plane and in such a
way that the area of the triangle OAB remains constant.
1.27 A moving point P has position vector r and polar coordinates (r,O).
Calculate the radial and transverse components of the vector d 3 r/dt 3 .
Given that:
(i) d 3 r/dt3 is parallel to r,
(ii) dOl dt = w, where W is constant,
(iii) r = ro, 0 = 0, drfdt = u when t = 0,
Exercises 53
rxi = Hr/r + a,
where a is a constant vector which has component -H in the direction of r.
Hence show that r lies on the surface of a circular cone with vertex at the
origin and axis a.
1.29 Show that the solution of
a xr=b,
Use this result to deduce that the field lines of the conservative vector field
V cp are orthogonal to the curves cp = constant.
54 Vectors
1.32 Sketch the curves ifJ = constant for the following two-dimensional scalar
fields ifJ(x, y):
1.33 Evaluate the following integrals along the straight-line paths joining the
end points:
where C is the semi-circle x = (1 - y2 )112 between (0, -1) and (0, l).
(ii) Ie (y dx + xdy),
where C is the parabolay = x 2 between (0, 0) and (2, 4).
where C is the square with vertices (1, 1), ( -1, 1), (-1, -1), (1, -1),
(ii) Ie (y dx -x dy),
where C is the triangle with vertices (0, 0), (1, 0), (1, 1),
(iv)
Exercises 55
0) fc (x 2 - y2) ds,
(iii) Jc ds,
(iv) Jc ds,
where C is the curve defined by x = sin t, y = cos t, z = t for 0";;;; t";;;; 21T.
(Ans.: 0, 2- 1/2 , !{1og(2 + 5112 ) + 2.S 1l2 }, 2 3/2 1T.)
1.37 Which of the following vector fields are conservative? Give the correspond-
ing potential functions for those that are.
(i) x- 1 i+y-lj,
(in y-l i + x -lj,
(iii) yzi + zxj + xyk,
(iv) (2xy + z3)i + (x 2 + 2y)j + (3xz 2 - 2)k,
(v) e z- 2 {yz(1 +x)i+(zx+z 2e-X )j-xyk}.
X
(Ans.: (i) </> = logxy, (iii) </>= xyz, (iv) </>= x 2y + xz 3 + y2 - 2z,
(v)</>= y(1 +Z-l xeX).)
1.38 Prove that
1 df
(i) gradf(r) = ~ dr r,
(ii)
Fx = al + bllX + bl 2Y + b 13Z,
Fy = a2 + b 2lX + bny + b23Z,
F z = a3 + b3lX + b32Y + b 33Z,
where the a;'s and the bi/S are constants. Show that such a field is locally
conservative if bij = bji for all i, j, and that the associated potential is then
2.1 INTRODUCTION
Then its velocity v and its acceleration a are defined by the relations
dr .
v=-=r (2.2.1)
dt '
dv d 2 r ..
a=-=-=r (2.2.2)
dt dt2 '
d - , dr' dr
-PP = - - - . (2.2.3)
dt dt dt
Note that the velocity of P' is the vector sum of the velocity of P and the
velocity of P' relative to P. This can be a useful rule for the calculation of the
velocity of a particle. In the same way the acceleration of P' can be calculated
from the vector sum of the acceleration of P and the acceleration of P' relative
to P.
Example 2.2.1
An aeroplane, flying in a north-east wind,* travels a distance d due north and
returns to its original position. The time for the outward journey is to and for
the return journey t r. Find the speeds v and w of the aeroplane in still air and of
the wind, respectively.
d
1;
(a) (b)
Figure 2.2.1 Illustration for Example 2.2.1: (a) outward journey; (b) return
journey.
By subtraction, we have
or
Example 2.2.2
A plane P takes off with constant velocity u and climbs at a fIxed angle a to the
horizontal. A helicopter H is in steady horizontal flight with velocity v at a
height h, and the two flight paths are in the same vertical plane. If the horizontal
distance between the helicopter and the plane is d when the latter is about to
take off, find the distance of nearest approach.
If a velocity -v is superimposed on both H and P, then H is reduced to rest
and the direction of the combined velocities -v and u will give the flight path of
P relative to H. If this path is PN, and if HN is the perpendicular from H on to
this path, then HN is the distance of nearest approach (see Fig. 2.2.2).
Let ~ be the inclination of PN to the horizontal and let 1 be the inclination of
PH to the horizontal. We then have that the distance HN, without regard to sign,
is
and a knowledge of HN depends on the calculation of~. To fmd ~ note that the
velocities -v and u must combine so that there is no component perpendicular
to PN. This implies that
v -H
+------------d-------------+
Example 2.2.3
A rabbit runs in a straight line with constant speed u. A fox chases the rabbit
with velocity v whose magnitude v(>u) is constant and is always directed
towards the rabbit. At time t =0 the fox is at a distance d from the rabbit and
62 Kinematics [2.2
the velocities of the two animals are perpendicular. Show that the fox catches
the rabbit after a time vd/(v2 - u2 ).
The curve followed by the fox is called the curve of pursuit and can be
calculated by considering the velocity of the fox relative to the rabbit. Denote
the rabbit by R and the fox by F, and let their initial positions be Ro and F 0
(see Fig. 2.2.3). Let RF = r, and let the angle between rand u be 0 + n/2. Then
re = u cos O. (2.2.6)
since r =d, 0 = 0 when t = O. This is all that is required to find the time taken to
catch the rabbit, for then r = 0 and hence the time is vd/(v 2 - u 2 ).
In order to find the path, it is necessary to integrate the relation
1 dr -v + u sin 0
--=---- (2.2.8)
r dO u cos 0
2.3] Constant Acceleration 63
which is implied by the ratio of (2.2.5) and (2.2.6). The integral of (2.2.8) is
r v
log d= - ;; log (sec e + tan e) - log cos e,
r 1
or (2.2.9)
d= cos e(sec e + tan et 1u .
Equations (2.2.7) and (2.2.9) contain all the information required to fmd the
path taken by the fox. They determine, at least implicitly, rand e (and hence r)
as functions of t. Thus the position vector Ro F of F, relative to Ro, can be
determined as a function of t since
There are several problems of practical interest where the acceleration is con-
stant in both magnitude and direction. The relations governing such motion are
easily derived by integration of the relation
(2.3.1)
dr
v=-=u+at (2.3.2)
dt '
r =ro + ut +! at 2 , (2.3.3)
v2 = u2 + a2 t 2 + 2u . at
= u2 + 2a. (ut + ~ at 2 )
=u 2 + 2a. (r- ro). (2.3.4)
The most general motion of this kind is two dimensional. For we can without
loss of generality choose the origin so that ro = o. Then the plane of the motion
is that plane of the constant vectors u and a which passes through the origin.
When u and a are parallel the motion is one dimensional and if
x = ut +tat 2 , (2.3.7)
(2.3.8)
2.4 PROJECTILES
g=9'81 ms- 2 •
6400)2
g (- - =0'997 g
6410 .
There would hence be a 0'3% error at this height if the variation in g were
ignored.
The fact that the acceleration is towards the centre of the earth means that
the assumption that it is constant in direction is also approximate. Two points
10 km apart on the earth's surface subtend an angle of 10/6400 radians or 0'1 0
of arc at the centre of the earth. This represents the departure from parallelism
between two such points.
The flattening of the earth towards the poles means that the earth is not a
true sphere and that consequently there is a slight variation in g along a
2.4] Projectiles 65
where cp is the latitude, so that there is a total variation ing of 0'5% between the
poles and the equator.
Finally the accuracy of the result will depend on the extent to which
additional mechanisms, which influence the motion of the projectiles, can be
ignored. In practice the most important of these is the resistance of the air. The
modifications to which this gives rise will be discussed in <hapter 4. For the
moment we investigate the consequences of the assumption that the gravita-
tional acceleration is constant in magnitude and direction and is the only
operative mechanism. Then the motion is described by equations (2.3.2), (2.3.3)
and (2.3.4) with a equal to the gravitational acceleration. In particular
r=ut+!gt2 (2.4.1)
if the particle starts from the origin with velocity u at time t =O.
We take axes Ox horizontally and Oy vertically in the plane of the motion.
The gravitational acceleration then has components (0, -g) and the initial
velocity u makes an angle a with the horizontal as in Figure 2.4.1. The governing
equations can be written
(2.4.4)
O ••__________~x___________ R----------------------+
Figure 2.4.1 The path of a projectile moving under the influence of gravity.
66 Kinematics [2.4
u sin ex
(=-- (=! T,say).
g
Then, from (2.4.3),
u
2 sin2 ex
h-~--~· (2.4.5)
g 2 g 2g
o= u( sin ex _ !gt2 ,
2u sin ex
and so (=0 or (=T).
g
Thus T is the time of flight and the range, R, on the horizontal plane through
the origin is therefore, from (2.4.3),
R =u cos ex ~ i
2U sin ex
g
1= u 2 sin 2ex
g
. (2.4.6)
It follows that the maximum range, for a fixed value of u and varying values of
ex, is Rm = u2 /g achieved when ex = 1T/4. Any intermediate range can be achieved
with two possible values for ex since, if ex = fJgives the required range, so also
does ex = t'1T - e. These two directions lie on either side of ex = 1T /4 and are
equally inclined to it.
Equations (2.4.3) are the parametric equations of the path of the projectile,
with ( as the parameter. The cartesian equation of the path can be obtained by
the .elimination of ( . It is
gx2
Y = x tan ex - 2 2 • (2.4.7)
2u cos ex
X R] R 2' R2 Rm
Figure 2.4.2 Illustration for Example 2.4.1: the parabola of safety.
The initial inclination a which takes the projectile through the point (xo ,Yo)
must, from (2.4.7), satisfy
gxo2 gxo2 2
Yo =xo tana- 2 =xo tana--2 (1 + tan a).
2 (2.4.9)
2u cos a 2u
2g ( gx2)
tan a = [ I ± { 1 - u2 Yo + 2U~:
}112] gxo
u .
2
(2.4.10)
For real values of a, and hence possible trajectories, the part of the expression
within the square root must not be negative, and so we have
gx5) ,
2g ( Yo +-2
1 #2
u 2u
or (2.4.11 )
The point (xo,Yo) must therefore not lie above the curve
(2.4.12)
are two possible values of a, say al > 1f/4 and a2, with al > a2. For these two
values of a, the right-hand side of (2.4.9) takes the value Yo. Hence, for
al ~ a ~ a2, the· right-hand side of (2.4.9) must have a stationary value as a
function of tan a. Two differentiations with respect to tan a shows that the
second derivative is _gx~/U2 < 0, so that the stationary value is a maximum.
Thus for al ~a~a2 the trajectory passes through the point (xo, y) where
y ~ Yo. This is the range of a for which it is possible for the projectile to pass
over the point (xo,Yo).
Corresponding to the two values al ,a2 of a are the horizontal ranges
• 2
R 1 -_u
2
SIn al
, (2.4.13)
g
with R 1 ~ R 2 • This inequality follows because the two trajectories can only
meet in one other point in addition to the origin. Since this point is (xo, Yo)
withxo ~O,Yo ~ 0, a contradiction is implied unlessR l ~R2.
If a2 > 1f/4, so that (xo ,Yo) lies between the trajectory for maximum range
and the parabola of safety, then Rl ~ R~R2 represents the interval of ranges
on the horizontal plane accessible to a projectile which has to pass over the point
(xo,Yo). Only one trajectory is possible for a given range R, for a second
trajectory would imply that a2 < 1f/4 which contradicts the assumption that
a2 >1f/4.
If a2 < 1f/4, the maximum range Rm = u2/g is attainable. Here Rl ~R ~Rm
is the interval of accessible ranges. For Rl ~R ~R2 only one trajectory is
possible, for which a> 1f/4. For R2 <R <Rm there are two trajectories
possible, one with a > 1f/4, the other with a < 1f/4.
Figure 2.4.2 shows two trajectories through the point (xo ,Yo) for which
at > a2 > 1f/4. The accessible points on the horizontal plane for a projectile
passing over (xo ,Yo) lie between R 1 and R 2 • The two trajectories through the
point (x~ ,y~) are such that al > 1f/4 > a2. In this case the accessible points lie
betweenR l andR m •
x =ut cos a - ~gt2 sin (3, y =ut sin a - ~gt2 cos (3. (2.4.1 5)
The maximum range, for varying values of ex, will occur when sin (2ex + (3) = I;
the inclination for maximum range is therefore
(2.4.18)
Note that since this gives the maximum distance that can be attained in a
direction inclined at an angle (3 to the horizontal, only points within the curve
whose polar equation is
u2
r=---- (2.4.20)
g(l + sin 8)
can be reached by a projectile starting from the origin. Equation (2.4.20) must
therefore be an alternative expression for the equation of the parabola of safety
70 Kinematics [2.4
(equation (2.4.12». One can readily show the equivalence of the two expres-
sions. For, by (2.4.20), we have
r + r sin 8 =r + y =U 2 /g.
Hence
Where the projectile strikes the plane we have, from (2.4.14) and (2.4.16),
2U sin lX)
x =U cos lX - g sin {3 (
g cos {3
=u cos lX - 2u sin lX tan {3,
2U sin lX)
y = u sin lX - g cos {3 (
g cos {3
= -u sin lX.
Example 2.4.3
A particle is projected from a horizontal plane to pass over two objects at
heights hand k, and a slant distance d apart. Show that the least possible speed
of projection is {g(h + k + dW12.
Since, from equation (2.4.4),
u~ =u~ - 2gh,
where uo,u p are respectively the speeds at 0 and P (see Fig. 2.4.4), it follows
that when the speed of projection U o is a minimum so is up. Now up is a
minimum when the maximum range along PQ is equal to PQ, and so
2
PQ = up- or up2 = g(d + k - h) .
g(l + sin (3)
o
Figure 2.4.4 Illustration for Example 2.4.3.
In this problem the point of projection 0 is not fIxed, but has to be chosen to
be consistent with the solution.
2.5 DIMENSIONS
x km = 1000x m,
On the other hand the measure of area, which has dimensions L2 , will change in
proportion to the square of the measure of length, so that
(2.5.2)
For acceleration we have
k IOOOa
h -2 -- 1000a m h- 2 -- (3600)2 -2 (2.5.3)
a m ms .
Note the special use of the equality sign in relations such as (2.5.1), (2.5.2),
(2.5.3). The actual numbers on the two sides of the relation
72 Kinematics [2.5
are different and have to be read in conjunction with the units in order to make
sense of the statement. Only when the same units are used on both sides of the
equation, or when the quantities are dimensionless, can the units be ignored and
the equation interpreted in a strict mathematical sense.
Although a meaning is attached to the product or ratio of quantities of
different dimensions, only quantities having the same dimensions are added. This
is a consequence of another assumption of fundamental importance, which is
that the validity of all physical laws is independent of the units used to measure
the quantities involved. Consider, for example, the statement
x = ut + !at 2 ,
x=u
can have no such general validity since it will cease to hold if the unit of time is
altered.
All the relations subsequently derived will be consistent in the above sense,
and this dimensional consistency is a very useful check of the algebra, or even of
the basic assumptions, in any physical problem.
The law of dimensional consistency can even be used to find the form of a
solution, and this can be put to significant practical use where problems are too
complex to be treated in full mathematical detail. As a simple example, suppose
we consider the distance travelled from rest by a particle given a constant
acceleration a. Since the distance travelled can depend only on a and the time t,
and since it must have the dimensions of a length, it must be proportional to at 2
and we have the result, apart from a numerical factor. If we introduce further
parameters, for example if the particle is given an initial velocity u, the correct
dimensional combination at 2 is still required, but the factor of proportionality
can now be a function of any non-dimensional combination of the parameters.
Since u/at is a dimensionless combination (and there is no other independent
combination) the distance travelled must be expressible in the form at 2 F(u/at),
where F is some unknown function. Of course in this simple example the
formula can be calculated explicitly (see equation (2.3.7)). It is
and is indeed of the right form. In more complex problems arguments of this
kind can bring to light the non-dimensional parameters on which a problem
Exercises 73
depends. From the practical point of view it is clearly useful to know that a
certain function F(u, a, t) of three parameters appearing in the problem is in fact
of the form F(u/at).
EXERCISES
2.1 Show that the hour and minute hands of a 12 hour clock coincide every
12/11 hours.
The hour, minute and second hands of the clock have a common axis
and they coincide at 12 o'clock (mid-day and mid-night). Do they coincide
at any other time?
2.2 An aeroplane flies at a constant speed v and has a range of action (out and
home) R in calm weather. Prove that in a north wind of constant speed w
its range of action is
tan ex - 1
tan e= I - cot f3 •
Find the ratio of the speeds of the ship and the wind in terms of ex
and f3.
(Ans.: u/w = (tan ex cot f3 - l){(tan ex - 1)2 + (1 - cot f3)2r 1l2 .)
2.4 A torpedo, which drives itself through the water at a constant speed u, is
fired from 0 at an enemy ship P whose constant speed through the water
is v. At the moment of firing the direction of motion of the enemy ship
makes an angle ¢ with OP (when the ship is moving directly away ¢ = 0).
Prove that, in order to allow for the motion of the enemy ship, the axis of
the torpedo must be aimed ahead of the ship by an angle D, where
74 Kinematics
sin 5 = v sin rp/u. Prove that this result is unaffected by a uniform tidal
current.
Prove that relative to the sea bed the locus of positions of torpedos
fired simultaneously in all horizontal directions from a fixed point is a
circle whose radius is independent of the tidal current. Find the centre of
the circle.
2.5 A boat, which can travel at speed u in still water, makes a trip from point
A to point B and back again. Because of the tide the shortest time for the
journey is achieved with a true speed Vo for the outward journey and vr
for the return journey. Show that if w is the speed of the tide,
2.6 A river, of uniform width d, flows along a straight channel and its speed at
a distance y from the bank is u(y). A boat travels from a point 0
(x = y = 0) on one bank of the river to the opposite bank. The speed of the
boat relative to the current is v, and the boat is set at an angle 8 to the
upstream direction.
Show that the velocity w of the boat is given by
If, further,
v = constant,
ad 3
- d cot 8.
6v sin 8
On the return journey the boat is set at a constant angle rp with the
upstream direction. Prove that it returns to the original point of
Exercises 75
1 6v - ad 2 1
6v >ad2 , tan 2cf> = 6 d2 cot zO.
v+a
(iii) Show that if both u and v are constant, and if the boat is always
directed towards a point 0 1 on the opposite bank directly across from
the starting point 0, the differential equation for the path is
the distance travelled and a and b are positive constants. Show that if it
starts from rest at the origin x =0, it next comes to rest when x =2alb
after a time Tr/b 1l2
2.13 A particle moves with constant speed u along the spiral curve r =ae - k8 ,
where a and k are positive constants. Show that if the particle starts with
8 > 0 at time t =0,
(v - U)2
6(v+u)'
or less than
where c is the speed of sound in air. Why is the larger root not appro-
priate?
Show that the depth of the well is approximately !gr (l - gT/c), if
gT~c.
Given that T = I' 5 sand c = 330 m s -1 show that the small correction
term alters the result by 4'4%.
2.21 A projectile is fired from ground level at A with the least speed needed to
clear an obstacle of height h at a horizontal distance a from A. Show that
if it reaches the ground at B, where AB = a + b •. then
2.22 A fort and a ship are both armed with guns which give their projectiles a
muzzle velocity (2ga}1I2, and the guns in the fort are at a height h above
the guns in the ship. The greatest (horizontal) ranges at which the fort and
ship can engage are d 1 and d 2 respectively. Show that
d1 =(a+h Y12
d2 a - h)
h cos 2 (3)1/2 }
2a sec2 (3 {(
, 1+ a - 1 .
2u 2 cos a sin(a - 8)
r=--------~~---
g cos 2 8
2.27 A projectile is fired from 0 with initial speed u at a target in the same
horizontal plane as O. Show that if a small error E radians is made in the
angle of elevation, and an error 2E in azimuth, the shot will strike the
plane at a distance from the mark 2U 2 E/g.
Show also that if the angle of elevation a is such that tan 2a <
2(a < 0'55 rad) an error in elevation will cause the shot to miss the target
by an amount greater than would be caused by an equal error in azimuth.
2.28 Mud is thrown off from all points of a tyre of radius a, which is rolling
along the ground with constant velocity of magnitude u. Show that, if
Exercises 79
We now introduce the laws of motion first postulated by Newton, which form
the basis of the whole of classical mechanics. We shall first state the laws in the
form in which they are usually quoted, and then proceed to discuss them.
It should be noted that the above three statements are formulated in terms of
certain concepts, namely force, momentum, action, which it is necessary to
define before the significance of the statements can be appreciated. It will
appear that their definitions depend, in turn, on the additional concept of mass,
this being the third and last fundamental concept of mechanics, after length and
time.
The first law associates the notion of force with that of acceleration, and
attaches significance to the state of zero acceleration. But before any statement
can be made about the acceleration of a particle a reference system is required,
since a particle may have zero acceleration relative to one set of axes, but not
relative to a different set. There is therefore an implication in this first law that
certain axes are preferential. This is connected with the belief that the appropri-
ate way to describe the acceleration of a particle is in terms of the influence of
other bodies, and that if a particle were isolated and free from any external
influence it would move with zero acceleration. A preferential, or inertial, set of
axes should be chosen to be consistent with such a motion. These axes, though
not arbitrary, are also not unique, for axes moving with constant velocity
relative to an inertial set also constitute an inertial set. The laws of mechanics are
in fact identical in the two frames, as will subsequently appear. This is called the
Galilean relativity principle.
As far as the idealised mathematical model is concerned, it is quite legitimate
to postulate the existence of inertial frames of reference. The theory, however,
does not explain how such a frame is to be chosen in practical applications, since
we are not able to isolate a particle from the influence of other bodies. Here we
must appeal to our experience to decide what is reasonable. Thus a system of
axes fixed on the earth's surface is adequate for many problems, in the sense that
a theory of Newtonian mechanics using these axes gives results which compare
3.11 Newton's Laws 81
satisfactorily with observation. Strictly speaking, however, a point on the earth's
surface does have an acceleration because the earth rotates about its polar axis.
At a latitude cp the acceleration is w~ r e cos cp towards the polar axis, where
re , We are respectively the radius of the earth and its angular speed of rotation.
Since, approximately, re = 6·4 x 10 6 m and We = 21T radians per day, we have
so that (3.1.2)
In the same way, by the substitution of any other particle P2 for PI, we can
determine a number rnz associated with Pz , which is invariant for all reactions
with Po.
We further assert that the result i:; transitive, so that ifP, and P2 are chosen
as the two interacting particles, then
F=mr. (3.1.3)
Gmim2(r2 - rd
1 r2 - ri 13
These can now be integrated to give ri and r2 as functions of the time, and
hence to predict the motion of the particles if the appropriate initial conditions
are given. There is a similar rule for the electrical force between charged
particles. For particles connected to a spring or elastic string there is a different
rule which says that the force is proportional to the extension of the spring (see
Chapter 6). Whenever a new mechanism is introduced a corresponding rule will
always be required before the calculations can proceed.
If the theory is to be generally applicable it is necessary to consider situations
more complex than the interaction between two particles. A particle may, for
example, be subjected to several forces (measured by the accelerations they
would produce if acting separately). We then assert that the resultant force,
obtained from the vector sum of the individual forces, is equivalent to the
product of the mass and the acceleration of the particle. This is the postulate of
the independence of forces. It enables us to use the empirical laws to calculate
the forces on a given particle, irrespective of the fact that there may be other
factors influencing its motion. We may also speak of an unaccelerated particle
being subject to several forces whose vector sum is zero.
3.3] Applications 83
Finally, for those who find it difficult to reconcile the idealised assertions
which have been made with their intuitive idea of what actually happens in
practice, the following quotation is the answer given by H. lamb (Dynamics
(1923) Cambridge University Press).
'The laws which are to be imposed on these ideal representations are in the
first instance largely at our choice, since we are dealing now with mental
objects. Any scheme of abstract dynamics constructed in this way, provided
it be self-consistent, is mathematically legitimate; but from the physical point
of view we require that it should help us to picture the sequence of
phenomena as they actually occur. The success or failure in this respect can
only be judged a posteriori, by comparison of the results to which it leads
with the facts. It is to be noticed, moreover, that available tests apply only to
the scheme as a whole, since, owing to the complexity of real phenomena, we
cannot subject anyone of its postulates to verification apart from the rest.'
3.2 UNITS
The unit of mass, as for all fundamental units, is a matter of choice. In the SI
system it is the kilogram (kg), which is 103 gram (g), defined in terms of a
material standard kept in Paris. In dynamics there is no particular reason for the
preference of the kilogram over the original unit of the gram, apart perhaps from
the fact that it tends to be more useful in everyday life. For the theoretical
reason for choosing the kilogram, reference must be made to the complete
system of SI units, which includes those used in electromagnetism. In this
discussion it is used simply because it is internationally acceptable.
Force is here a derived concept with the dimensions of mass times acceler-
ation, or MLT- 2 • The unit in the SI system is the newton (N), which is the force
represented by 1 kg having an acceleration of 1 m s -2.
Since it is observed that the attraction of the earth gives all bodies an
acceleration g, it follows that the associated force of attraction is mg. This will
vary from place to place because of the variation of g, but at a given place g is
constant and the mass of a body is proportional to the gravitational force, called
the weight (wt) of the body. Since the weights of two bodies can be compared
on a balance, this also gives a practical means of comparing masses. In fact the
force of magnitude mg newtons is sometimes referred to as m kilogram weight
(kg-wt), but it should always be kept in mind that it has the dimensions MLr2
and refers to the weight of a mass m kg. In problems possible confusion is best
avoided by working entirely in terms of the force mg newtons rather than
m kg-wt.
3.3 APPLICATIONS
In the following examples certain assumptions and idealisations are made, the
chief of which are discussed below.
84 Dynamics [3.3
Some bodies, which are manifestly of fmite size, can be treated as if their
equations of motion are identical with those of a particle subject to the same
forces. Strictly speaking this requires a study of rigid body dynamics, but at this
stage such treatment may be accepted, at least for one-dimensional motion or
where the size is clearly not an important parameter.
When two bodies are in contact they exert, by Newton's third law, equal and
opposite forces on each other, called contact forces. A contact force may vary
according to the dynamical situation, but when bodies are smooth this force is
always perpendicular to the plane of contact of two such bodies. This is an
idealisation which neglects the frictional force (see Chapter 4).
When two particles are connected by an inelastic string, the string transmits a
force, called the tension of the string. Consider an element os of a straight string
moving parallel to itself with acceleration a (see Fig. 3.3.1). The contact forces
between os and the rest of the string are T(s + os) and T(s), say, so that the
resultant force on os is
where p is the mass per unit length of the string. Division by os gives, in the limit
as os -+ 0,
If the string passes round a smooth peg or pulley this result is not affected
since the contact force between the string and the pulley is perpendicular to the
string and hence does not affect the equation of motion along the string.
There is an extra term if the string is subject also to a gravitational force. For
example, for a string pulled vertically upwards with acceleration a, equation
(3.3.1) is modified to
dT
ds - pg= pa. (3.3.2)
os
Example 3.3.1
A particle, of mass m!, is free to slide on the inclined face of a smooth wedge, of
mass m2 and angle G!. The wedge is itself free to slide on a smooth horizontal
plane. Find the acceleration of the particle and the wedge.
Let the acceleration of the particle relative to the wedge be al down the face
of the wedge, and let the acceleration of the wedge be a2, as shown in
Figure 3.3.2.
The forces consist of the normal reaction Nl between the particle and the
wedge, which is a mutual force perpendicular to the wedge face, the reaction N2
between the wedge and the plane, and the gravitational forces on both the wedge
and the particle.
The equations of motion, when resolved horizontally and vertically, are, for
the wedge,
NI sina=m2a2,
m2g-N2 +N I cosG!=O,
The four equations are to be solved for the four unknowns ai, a2, NI and N 2 . In
particular
(3.3.3)
(3.3.4)
Nj \ 1mjg
~ ~~ -a2
\ lm2 g aj~
ex
Nj
N2 i
~ :0- ~
(a) (b)
Figure 3.3.2 Illustration for Example 3.3.1: (a) the forces; (b) the accel-
erations.
86 Dynamics [3.3
llJCample 3.3.2
A light inextensible string passes over a smooth pulley and particles, of mass m 1
and mz respectively, are suspended from the ends of the string. The pulley has a
constant acceleration ao vertically upwards. Find the accelerations of the masses.
The accelerations of the particles relative to the pulley are equal and opposite
since this is true of their displacements. Let these be ± a so that ml has an
acceleration ao + a and mz has an acceleration ao - a as shown in Figure 3.3.3.
The forces involved are the tension in the string, which is constant if the mass
of the string is neglected, and the gravitational forces .
The equations of motion are
Elimination of T gives
(3 .3.5)
ao + a
T rao - a
m2g
Tcos a= mg.
3.3] Applications 87
.
T
..-
(
'---_C2.._--/~ mg
T sin ex = mw 2 [ sin ex
2 g sec ex g
or w =--=- (3.3.6)
[ OC'
where OC is the depth of the particle below the fixed point. The time for one
complete revolution is
(3.3.7)
llxamp/e 3.3.4
A particle is slightly displaced from rest in its position of eqUilibrium at the
highest point of a smooth fixed sphere of radius a. Determine the point where
the particle leaves the sphere.
When the radius vector from the centre of the sphere to the particle makes an
angle e with the vertical as in Figure 3.3.5, the components of acceleration along
and perpendicular to the radius vector are (-a(j2, a8). If the normal reaction
88 Dynamics [3.3
I
I
I ;'
ie>;,;' ;'
L-/
between the sphere and the particle is N, the equations of motion are
Since N cannot become negative, the particle will lose contact with the sphere
when N = 0, that is when cos 8 = 2/3. It will then begin to move freely under
gravity.
Example 3.3.5
A bead is free to slide on a smooth, straight, horizontal wire which rotates with
constant angular speed w about a vertical axis through a point 0 of the wire (see
Fig. 3.3.6).
e
o
Figure 3.3.6 Illustration for Example 3.3.5: bead sliding on a rotating wire.
Exercises 89
f - rw 2 = 0. (3.3.10)
This is a second order linear differential equation with constant coefficients. The
solutions of such equations are, in general, of exponential type, possibly with
complex exponents, and so are of the form r = A eAt , where A and A are
constants. Substitution in (3.3.10) shows that
u u
r=-(e wt _e- wt )=- sinhwt.
2w w
EXERCISES
F = Ai + Btj + Ct 2 k,
where t is the time and A, Band C are constants. Initially the particle is at
rest at the origin. Find its position vector r at any subsequent time and
show that
where the integral is taken along the path of the particle from the origin to
its position after unit time.
3.2 A particle of mass m is moving with constant velocity u relative to an
inertial frame. A second frame of reference has its axes always parallel to
those of the inertial frame but its origin has the position vector
ai + btj + ct 2 k, where a, b and c are constants and t is the time. Find the
displacement of the particle relative to the second frame and show that the
particle appears to be subject to a force -2mck.
90 Dynamics
3.3 A particle rests in equilibrium under the action of three forces. Show that
each force is proportional to the sine of the angle between the other two.
3.4 For time t < 0 a particle rests at the top of a smooth hemisphere of radius
a. The hemisphere rests with its plane face on a horizontal plane. For t;;;;' 0
the hemisphere is constrained to move with constant velocity, of magni-
tude U, along the plane. Show that for u 2 ""ga the particle leaves the
hemisphere when
3 cos () - 2 - u 2 /ga = 0,
where () is the angle between the radius vector to the particle and the
upward vertical. Deduce that if u 2 ;;;;. ga the particle will hit the horizontal
plane at a point vertically below its initial position.
3.5 A man of mass m is holding on to a light rope which passes over a smooth
fixed pulley and has a counterpoise of mass m attached to the other end.
Initially the system is at rest. Show that if the man begins to climb his
portion of the rope the man and the counterpoise move equal distances.
3.6 Over a smooth pulley A, of mass 2 kg, passes a light inelastic string having
masses 6 and 9 kg at its ends. Over a smooth fixed pulley B passes a light
inelastic string having a mass 15 kg at one end and being attached to the
axis of the pulley A at the other end. The system moves so that the parts
of the string not in contact with the pulleys move vertically. Show that the
acceleration of the 15 kg mass is 0 '44 m s -1.
3.7 Two particles, of masses m 1 , m2, are joined by a light inelastic string and
are initially close together on a smooth table. A smooth ring, of mass m 3,
is threaded by the string and hangs over the edge of the table and the
portions of the string on the table are at right angles to the edge. Find the
acceleration of the ring.
(Ans.: m3(ml +m2)g/{4mlm2 +m3(ml +m2)}')
3.8 A bead slides under gravity, on a smooth straight wire, from rest at a point
o to a point X. Show that if X lies on a prescribed circle having 0 as its
highest point, the time of descent is independent of the position of X on
the circle. Hence find the straight line of quickest descent from:
(i) a point to a straight line,
(ii) a straight line to a point,
(iii) a point to a circle.
3.9 A smooth cone of angle 2a has its axis vertical. A particle moves in a
horizontal circle on the inner surface of the cone at a height h above the
vertex. What is the speed of the particle and what is the period of the
motion?
(Ans.: (gh)1/2, 2rr(h/g)1/2 tan a.)
3.10 A light string ABC passes through a small, smooth ring B, and has its ends
fixed at two points A, C such that C is vertically below A. If the ring is
whirled round so as to describe a horizontal circle with constant angular
speed W, show that
2(a + c)bg = {(c + a) 2 _b 2 }(c - a)w2 ,
3.11 A bead can slide freely on a smooth straight wire OA, of length I, which is
rotated in a horizontal plane with constant angular speed w about its end
O. Initially the bead is projected along the wire with speed u from O.
When the bead leaves the wire, show that the angle between the line of the
wire and the direction of motion of the bead is tan -1 {lw/(u Z + IZ W Z )112}.
3.12 A smooth straight wire OA, of which 0 is fixed, is made to rotate in a
fixed vertical plane with constant angular speed w. A bead of mass m,
which is free to move along the wire, is initially at rest at 0 when OA
points vertically downwards. Show that if r is the distance of the bead
from 0 at time t
;: - W Zr = g cos wt
3.13 A ring of mass ml is free to slide along a fixed horizontal smooth wire. An
inelastic string of length 1 has one end fastened to a fixed point of the
wire, is threaded through the ring and has a particle of mass mz fastened
to its other end. The particle is allowed to fall from an initial position in
which the string is taut and horizontal with the ring at its tnid point.
Write down the accelerations of the ring and the particle when the
portion of the string between the ring and the particle is of length r( < I)
and has turned through an angle 0«
1T/2) from the horizontal.
Hence show that:
r=-1 ( 1 +3-mz
- 04 ) •
2 56 mt
3.14 A ring, of mass ml, is free to slide along a fixed smooth wire inclined at an
angle a to the horizontal. An inelastic string has one end fastened to a
92 Dynamics
fixed point of the wire, is threaded through the ring and has a particle of
mass m2 fastened to its other end.
(i) Let the system be released from a position of rest in which the string
is taut and that portion of it between the ring and the particle inclined
at an angle a + 8 to the horizontal. Show that the initial accelerations
of the particle along and perpendicular to the string are ;: ( 1 - cos 8)
and g cos(a + 8), where
(ii) Show that a motion is possible in which the ring moves with constant
acceleration up the wire provided 8 is constant and such that
3.15 A smooth circular cylinder of radius a is fixed, with its axis vertical, to a
smooth, horizontal table so that its plane base is in contact with the table.
A light inextensible string of length 21 is fastened to a point of the
circumference of the base of the cylinder and a length 1 of the string is
wrapped round the cylinder. A particle of mass m is attached to the other
end of the string and moves on the table in such a way that the string
always remains taut. Initially the particle is given a horizontal velocity u
perpendicular to the unwound part of the string in such a direction that
the string starts to unwind.
Show that the position vector of the particle relative to the axis of the
cylinder at any instant can be expressed as
r=ap+(l+a8)q,
where p and q are unit vectors along a radius of the cylinder and along the
string respectively. Hence prove that the string is completely unwound
after a time 3z2 j2au, and calculate the tension in the string as a function of
the time t «312 j2au).
(Ans.: mu 2 (12 + 2aut) -1/2.)
4 Resisting Forces
We have seen in the previous chapter that there are mutual reactions between
two bodies in contact and that these reactions are equal in magnitude and
opposite in sense according to Newton's third law. When the bodies are ideally
smooth the reactions are perpendicular to the plane of contact, which is the
common tangential plane for bodies of continuous curvature. It is, however, a
matter of everyday experience that when one body slides on another there is a
force which opposes the relative motion, and that a body at rest in contact with
a fixed body requires a certain minimum impressed force before it will move.
This means that, in addition to a component N normal to the plane of contact,
the reaction R has a component F in the plane of contact, called the frictional
force. Account is taken of the frictional force according to the following
idealised laws:
surfaces, a process which is assisted by the heat generated when there is a relative
motion. It is this welding and subsequent shearing that is the source of the
frictional force. Since the true area of contact, where the shearing and welding
take place, tends to be proportional to the load, so does the frictional force.
There are exceptions to the rule. For very hard substances, such as diamond, the
true area of contact is not proportional to the load and f.l.s decreases as the load
increases. A substance such as mica can be molecularly smooth in some circum-
stances and then the true and apparent areas of contact are comparable. The
frictional behaviour is then quite different from that for substances like metals.
The coefficient J.l.s varies from one material to another, and even for the same
material under different conditions. The degree of roughness of the contact
surfaces is clearly a factor, and f.l.s tends to decrease to a constant value as the
surfaces become smoother. The trend can, however, be dramatically reversed for
highly polished surfaces. For example f.l.s = 1 for glass under normal conditions,
and does not vary appreciably between ground glass and plate glass surfaces. But
if the surfaces are optically flat, highly polished and free from all surface
impurities, the adhesion can be so great that the glass will fracture before sliding.
Similarly two pieces of lead with flat and highly polished surfaces weld together
over an appreciable area and seizure occurs on contact, whereas under normal
conditions J.l.s is about 1·5. Other metals also react in this way, though they may
require very careful treatment to aohieve substantial adhesion.
The law of proportionality can be violated for some metal surfaces because
under normal conditions most metals have a coating of oxide, and J.l.s is smaller
for the contact of such surfaces than it is for the uncoated metal itself. For
example J.l.s = 0·4 for copper surfaces if N < 10-3 kg and 1·6 if N> 10-2 kg.
Between these values of N there is a continuous transition. The reason is that for
light loads friction arises from the oxide coatings, but as the load becomes
heavier this coating is gradually broken through by the local high spots on the
surfaces so that actual metallic contact increases. If the thickness of the oxide
film is deliberately increased the transition is delayed but the range 0·4 ~
f.l.s ~ 1·6 is essentially unchanged. For aluminium, which is a softer metal,
f.l.s = 1·2 and does not vary appreciably over the range 10 -5 to 1 kg for N since
the surface film is easily broken. By contrast f.l.s = 0·4 for chromium over the
same range for N because both the oxide and underlying metal are hard
enough to resist deformation and remain intact.
The coefficient of kinetic friction is generally somewhat lower than f.l.s, but
the difference is not substantial. It arises because the welding of local high spots
has ample time to take place when the surfaces are not in relative motion. When
they move the welds are continually being broken before they have time to
consolidate. However, the source of the frictional force, namely the local welding
of the two surfaces, is essentially the same.
A change in the mechanism of friction can, of course, substantially affect the
sliding properties. Thus the introduction of a thin mm of lubricant can reduce
the coefficient of friction between metals by a factor of 5, and is an important
consideration in the design of the moving parts of most machines. If the relative
[4.l Solid Friction 95
speed between the surfaces is sufficiently high, frictional heating can result in
lubrication by a melting of the contact surfaces themselves. This can occur at
quite low speeds for ice on ice, for which f.ls = 0·1 at 0 °c and 0·5 at -50°C,
whereas 11k = 0·1 over this range of temperature at a speed of 4 m s -1. Similar
decreases occur for metals, but the speeds required for melting are much higher.
The conclusion is that the law of proportionality between F and N is a
reasonable approximation if attention is paid to the conditions under which it is
being applied. It is not to be used indiscriminately.
Table 4.1.1 gives the values of J1s for some common substances. It should be
interpreted as a rough guide since experiment indicates that the coefficient can
vary according to the conditions. The coefficient 11k is rather less than I1s and its
variation with conditions is more marked.
Example 4.1.1
A particle is at rest on a rough horizontal plane. The plane is slowly tilted until
the particle starts to move, and is then kept fixed. If the static and kinetic
coefficients of friction are I1s and 11k «J.1.s) respectively, find the velocity of the
particle after it has travelled a distance d.
Let Q be the inclination of the plane to the horizontal (see Fig. 4.1.1). In
equilibrium we have
F
Hence tan Q= -=J.1.s
N
mgsin a- ma
Hence =J.1k,
mg cos a
Example 4.1.2
A rough wedge, of mass m2 and angle a, is free to move on a horizontal plane. A
rough particle, of mass m 1, is free to move on the inclined face of the wedge.
The accelerations and forces are displayed in Figure 4.1.2. The horizontal and
vertical components of the equation of motion for the wedge are
\;:::'
~ \ lm2g
.!!1-~
al~
-a2
Fl Nl lN2
rt F2 -
/::: '/;
(a) (b)
Figure 4.1.2 Illustration for Example 4.1.2: (a) the forces; (b) the accel-
erations.
There are only four independent relations to be obtained from the equations
of motion, although different combinations may simplify the calculations. For
example FI and NI can be eliminated by the substitution of (4.l.2) in (4.1.1) to
get
In addition (4.1.2) can be solved to give FI ,NI in terms of aI, a2' The resulting
equations are equivalent to resolving the equation of motion of the particle
along and perpendicular to the wedge instead of horizontally and vertically.
Thus we have
In the four independent relations (4.1.1), (4.1.2) or (4.1.3), (4.1.4) there are
six unknowns, namely F I , N l , F 2 , N 2 , aI, a2' Hence two supplementary
conditions are required. There are four possibilities to consider:
(i) Both the wedge and the particle are in equilibrium. The two further con-
ditions are al =a2 = O. Equations (4.l.3) and (4.1.4) then give FI ,NI , F 2 ,
N2 immediately.
(li) The particle slides on the wedge and the wedge remains at rest. Here
a2 = 0 and FI = ±J.lINI are the two further conditions, where J.lI is given
and the sign depends on the direction in which the particle slides. If, for
example, it starts from rest, then FJ = J.lJ NJ and equations (4.l.4) show
that
(iii) The wedge slides and the particle does not slide relative to the wedge. Here
iiI = 0 and F2 = ±1l2N2' Equations (4.1.3) are then sufficient to determine
F 2 ,N2 anda2. If, for example, F2 = 1l2N2 we have
The coefficients III and 112 refer to the coefficients of kinetic friction at the
appropriate points of contact.
Note that in cases (i)-(iii) there will be conditions to be satisfied by the
coefficient of static friction arising from the fact that there is no slip. For
example in case (i) /J.s > tan a for the particle.
Example 4.1. 3
A light string is wrapped round a rough circular cylinder and is in a plane at right
angles to the generators of the cylinder. Show that when the string is on the
point of slipping the tensions at its two ends are in the ratio eMsO , where 8 is the
angle between the normals at the two ends of that part of the string in contact
with the cylinder.
We consider an element of the string of length os subtending an angle 08 at
the axis of the cylinder. Let the tensions in this element at its two ends be T and
T+oT. Let the notmal component of the reaction be Nos, where N is the
,NOS
Fos_~ _ _~
T~ ~
T+oT
reaction per unit length, and let the frictional force be Fos (see Fig. 4.1.3). Since
the string is in equilibrium, the components of the equation of motion along and
perpendicular to the normal reaction give
These equations are now divided by os and considered in the limit a'S os -+ 0,
and hence 0 T -+ 0,00 -+ O.
oT dT sin !08
----+--
1 d8
Since --+-
os ds' os 2 ds
dT dO
we have that -=F T-=N
ds ' ds '
1 dT F
or --=-
T dO N
For limiting equilibrium, FIN = Ils' and the resulting equation integrates to
When a body moves through a fluid, such as air or water, the fluid is set in
motion and the body thereby experiences a force. For a particular body and a
particular fluid the force depends on the velocity, but a detailed theoretical
solution is one of the outstanding unsolved problems of fluid dynamics,
although a great deal of experimental information is available. To simplify the
analysis we resort to approximate expressions, but it should be kept in mind that
there is no adequate universal formula.
We know from the law of dimensions that the force should be represented
by a dimensionally correct combination of the parameters on which it depends.
Let us consider what these parameters might be. The speed v and the size of the
body, which we account for by a representative length t, are clearly involved. We
might also expect the density p of the fluid to be a factor. It is easily verified
that the correct dimensions of force are given by the combination pt2 v2 •
Moreover there is no non-dimensional combination, since p is the only parameter
whose measure varies with the mass of the fluid, and v is the only parameter
whose measure varies with time. Thus a practically useful representation of the
100 Resisting Forces [4.2
through air or water would achieve a Reynolds number of several hundred. If,
however, the fluid were olive oil the linear law of resistance would be quite
reasonable (see Exercise 4.10).
We mention one other non-dimensional factor which is velocity dependent
and which may affect the value of C. This is the Mach number, which is the ratio
of the speed of the body to the speed of sound. In the transonic regime, where
the Mach number is of order unity, this can be a significant effect. But for
speeds which are small compared with the speed of sound, which is about
330 m s -1 in air at ground level, this effect can be ignored. It is invariably
ignored in liquids. The speed of sound in water, for example, is 1· 4 x 103 m s -1 .
Apart then from certain exceptional regimes, it appears that the simplest rule
which might be generally acceptable is one in which the force varies as the
square of the speed provided, as will usually be the case, that the Reynolds
number is not small. If the Reynolds number is small then a linear law of
variation is appropriate. Unfortunately, apart from one-dimensional problems, it
is only the linear law which is amenable to simple analysis. This is because it
leads to linear differential equations which are easier to handle than non-linear
equations.
In some of the following examples the linear law of resistance is used where
the conditions do not seem to be appropriate. This is partly because a theory is
best illustrated by examples which are not too complicated. But they should not
be dismissed as completely artificial, because the effects of resisting forces of
different kinds have certain features in common. It is often useful to have a
simple solution where the gross features are qualitatively acceptable, even
though quantitatively they are not (see also Exercise 4.11).
Example 4.2.1
We consider the one-dimensional motion of a particle subject to a resistance
whose magnitude per unit mass is equal to k times the square of the speed, and a
constant force Iper unit mass.
(i) Suppose first that the constant force also opposes the motion, so that the
total force is -m(1 + ku 2 ) as shown in Figure 4.2.1. The equation of motion
is then
du _ 2
- - -I - ku . (4.2.1)
dt
mf ......- - -
mk,,2 .....t--_ _ •
Note that this case only arises when the particle is given an initial
velocity u in the positive x direction. Eventually the particle will come to
rest, and then begin to move in the opposite direction since the acceleration
is negative. When this happens the resisting force will change direction, since
it always opposes the motion. Equation (4.2.1) is valid only up to that
point.
It is convenient to write f= k y2 , so that
dv
-= _k(V2 + v2 ). (4.2.2)
dt
dv dx dv dv
- = - -=v-=-k(V2 +V2)
dt dtdx dx '
V2 +u2 )
log ( 2
.V +v
2 =2kx . (4.2.4)
Equations (4.2.3) and (4.2.4) show that the body will come to rest in
a time {tan- 1 (u/V)}/kV after travelling a distance {log(I +U 2/V2)}/2k.
For example, suppose a ship reverses its engines when going full speed
ahead. Here u =V (see below) and it will travel a distance !k- 1 log 2 in time
1T/4kVbefore corning to rest.
(ii) When the particle moves in the direction of the propelling force, it is less
confusing if x is measured in the same direction, as in Figure 4.2.2, so that
the equation Cif motion is now
dv dv
- = v - =f - kv 2 = k(V 2 - v2 ). (4.2.5)
dt dx
mkt,2 .. • mf
• • x
• v
I
'(V + v)(V - u) \
log (V _ v)(V + u) = 2kVt, (4.2.6)
log I VV22 - U2 2
1
= 2kx (4.2,7)
v
1
x =-log (1 u2 )
+- (4.2.8)
2k V2
1 _ U
in time tl = -tan 1 -, (4.2.9)
kV V
1
x == -k log 2
(V2) == -log
2
1
cosh kVt. (4.2.11)
2 V - v k
t2 = -
1 sm
. h- 1 -.
U
(4.2.12)
kV V
Example 4.2.2
A particle, of mass m and position vector r relative to 0, is projected from
with initial velocity u at an angle a to the horizontal. It is subject to a
°
gravitational force mg and a resisting force R == -mkv, where v is its velocity (see
Figure 4.2.3).
104 Resisting Forces [4.2
~mg
"'--------'------------------.
o x
Figure 4.2.3 Projectile subject to gravitational force mg and resistance R.
f= g - lei. (4.2.l3)
(4.2.14)
(4.2.l5)
u cos ex ( 1-e-kt)
x=--- (4.2.l6)
k '
(4.2.l7)
y = ~ log (1 - u~
k cos ex
) + x (tan ex + g
ku cos ex
). (4.2.18)
4.2] Motion Through a Fluid 105
Figure 4.2.4 shows the effect of resistance on the trajectory of the particle.
The various trajectories shown in the figure were calculated using a non-
dimensional form of equations (4.2.16) and (4.2.17). This preliminary reorgan-
isation of the equations is an invariable procedure in all numerical work, for the
reasons discussed in §2.5. It is always an advantage in practice to use non-
dimensional variables whose magnitudes are of order unity, and this should be
one of the guides in the choice of the variables. Now in the special simple case of
no resistance we know from §2.4 that the values of t, x and y at the highest
point of the trajectory are, respectively,
(4.2.20)
and hence to one basic shape, independent of u,g, CI!, in the plane of (~, 11). This
is another hopeful sign that (4.2.19) is a useful transformation, because another
important consideration is to seek transformations which minimise the number
of significant parameters on which the problem depends.
Substitution of relations (4.2.19) into (4.2.16), (4.2.17) and (4.2.18) gives
the follOwing equations:
(4.2.21)
where
kusinCl! usinCl!
Kl = =-- (4.2.23)
g V
and V= g/k is the terminal speed of the particle in free fall under gravity. In
these variables, then, the reduced trajectory depends on the four parameters k,
u, g, CI! only in the non-dimensional combination K 1 = kug -1 sin CI!, which we call
106 Resisting Forces [4.2
the specific resistance (the suffIx denotes that the resistance is linear). The
advantage is that calculation can now be confmed to the family of curves
depending on the single parameter K 1. In any specific problem the trajectory
will be given by that member of the family having the appropriate value of K 1 ,
and the values of k, u, g, a merely determine the appropriate scales of length and
time through the transformations (4.2.19).
When Kl ~ 1, the expression for 1/ in (4.2.22) can be expanded to give
or, equivalently,
(4.2.25)
(4.2.26)
(4.2.27)
4.2] Motion Through a Fluid 107
(4.2.28)
(4.2.30)
(4.2.31)
(4.2.33)
where € = O· 0025 for K 1 = 2· 3 and is smaller than this for all other values of
K l ' There is no a priori reason why such a simple formula should exist and no
general method of deducing such formulae, although to be successful they must
have some gross features in common with the exact result. In this case, with
€ = 0, expression (4.2.33) agrees with (4.2.30) for Kl ~ 1 and with (4.2.32) for
K 1 ~ 1. Although the exponentially small terms are not modelled correctly,
they are of no significance compared with Ki 1 when K 1 ~ 1.
108 Resisting Forces [4.2
0'4
0-2
0-4
0'2
--------- a = 0
----a=rr/4
-a=rrl2
Example 4.2.3
When the resistance R varies with the square of the speed the projectile problem
is more complicated. Since the resistance is given, in magnitude and direction, by
an expression of the form -mkuv, the equation of motion is
r=g-kuv (4.2.34)
where (4.2.36)
where (4.2.39)
(4.2.41)
(4.2.42)
Next we use (4.2.42) to eliminate dVdr from the second of equations (4.2.41),
which then becomes
d 2 7] = -1 _ K2 d7]
dr2 1 + K2 r dr'
or
(4.2.44)
d7]=
-
dr
( 1+-1) 1 1
--(1+K 2 r),
2K2 1+K2 r 2K2
(4.2.45)
110 Resisting Forces
1 ( 1) 1 1
1/ = K2 1 + 2K2 10g(1 +K2T) - 4K~ (1 + K2T)2 + 4K~
= (1 + _1_)~
2K2
_ IrA (e
4A2
2K 2 ~ - 1). (4.2.46)
If the exponential term is first written as a power series, it is easily seen that
(4.2.46) agrees with (4.2.20) in the limit K2 ~ O. This means that, in addition to
being correct for finite values of K2 in the limit a ~ 0, (4.2.46) also yields the
correct trajectory in the limit K2 ~ 0 without restriction on a. The range of
validity is consequently greater than might otherwise have been expected. On
the other hand it should be borne in mind that the validity of the approximation
really depends on the inclination of the trajectory to the horizontal being small
throughout the flight of the projectile. This is clear from equations (4.2.35),
which are the dimensional form of (4.2.41) with the approximation v =x.
x
Equation (4.2.36) shows that this requires y ~ at all times. If the trajectory is
bounded by the horizontal plane this is ensured by a ~ 1. But if the proj ectile is
allowed to continue through to negative values of y, the approximation will
eventually fail.
For general values of K2 and a it is not possible to integrate equations
(4.2.37) and (4.2.38) in terms of elementary functions. However, it is a straight-
forward matter to integrate them numerically using a computer. The trajectories
shown in Figure 4.2.5 were produced from such a numerical integration and
should be compared with the equivalent Figure 4.2.4 for the linear law of
resistance. Note that the trajectories labelled a = 0 and a = n/2 are limiting
trajectories for the non-dimensional variables (~, 1/). Of course in the physical
(x, y) plane these trajectories degenerate to a point and a vertical straight line
respectively because the non-dimensionalisation by means of (4.2.19) also in-
volves a. The trajectories in the (t 1/) plane are preferred because they exhibit
information in a more revealing fashion and, as in the previous example, the
trajectory in the physical plane is readily obtained by the appropriate change of
scale, which will involve other parameters as well as a. We see in particular that
in the (t 1)) plane the variation with a is less significant than the variation with
K 2 , and that the nearly vertical descent for K 1 = 5 is not duplicated in the
trajectory for K2 = 5.
EXERCISES
4.1 A rough plane is inclined at an angle a to the horizontal. A particle of mass
m is attached to a fixed point of the plane by a light string, and rests in
eqUilibrium with the string taut and making an acute angle (J with the line
of greatest slope down the plane. Show that sin (J ,;;;; J1 s cot a, where J1s is
the coefficient of static friction.
When sin (J = J1s cot a, show that the tension in the string is
Exercises 111
4.2 Two particles, each of mass m, lie at points A and B on a rough horizontal
plane, and are connected by a taut light inelastic string AB. The particle at
B is pulled by a horizontal force T applied to a second light inelastic string
BC, where ABC = 1T - a Ca < 1T/4).
Show that the least force T which will cause both particles to be about
to move is T = 2f..Lsmg ICOS a.
Explain the significance of the restriction a < 1T/4.
4.3 Two particles of equal mass are connected by a light string. The first
particle rests on the surface of a rough circular cylinder, whose axis is
horizontal. Part of the string is in contact with the cylinder over an angular
distance a, the remainder hangs vertically with the second particle sus-
pended from its end. The two particles and the string are in equilibrium in
a vertical plane. The coefficient of friction between the particle and the
cylinder, and between the string and the cylinder, is I1s.
Write down the equations for equilibrium. Show that when the friction
is limiting,
Ca - 1) ea = 1, or a = 1·2785.
4.4 The mass of a train is 5 x lOs kg, the driving wheels support a weight of
4 x 10 4 kg-wt and the coefficient of friction between the driving wheels
and the rails is O· 6. Show that at the end of lOs after starting from rest
the speed will be less than 17 km h -1 .
4.5 Two equally rough planes slope down at angles a and (3 to the horizontal
from the horizontal ridge in which they meet. A particle is projected from
the bottom of the first plane up the line of greatest slope with a speed v
just sufficient to carry it over the ridge. After descending the other face it
reaches the level from which it started with speed kv. Prove that, if
11k < tan (3,
e
where a is the value of for which v = o.
Show that the frictional force is of magnitude I1kmg cosh {2I1k(a - e)}.
e e
Show that the line integral of this force, between = 0 and = a, is! mV6,
e
where Vo is the value of v when = o.
4.8 Two particles, of masses ml, m2 respectively, are connected by a long
light string which passes over a rough circular cylinder whose axis is
horizontal, and lies in a plane at right angles to the axis of the cylinder.
Show that the two particles can hang in equilibrium if m2 < ml <
m2 ells1T • Show that if m I;;;' m2 ells1T the acceleration of the particles is
when 7= o.
Integrate these equations, and discuss the subsequent motion of the
particle, on the basis of the simplifying assumption that tan a ~ 1.
In particular show that, if 11k cot {3 > !, the ultimate value of ~ is
(211k cot (3 - 1) -I approximately.
Exercises 113
Show also that if Ilk cot (j> 1, the particle will finally come to rest but
that if ! < Ilk cot (j < 1, the particle will ultimately slide in the y direction
with constant acceleration.
4.10 (i) A spherical grain of sand, of radius 1= 10 -3 m, falls freely under
gravity in air. Show that its terminal speed V is 9 m S-1 and that the
corresponding Reynolds number VI/v is 600.
If the medium is water instead of air, show that V is 0'23 m S-1
and that the Reynolds number is 230.
Assume that the drag coefficient CD = 0'5 and g = 9'8 m s-2
(ii) When the medium is olive oil, it is found that V= 2 x 10-2 m S-I.
Show that the Reynolds number cannot be greater than 0'2 and
deduce an appropriate law of variation of resistance with velocity.
Verify from the equation of motion that the terminal speed will
never be achieved in finite time but that a speed of 99 V /100 will be
achieved in about 0'02 s after starting from rest.
Use the following physical constants:
Kinematic viscosity,
Specific gravity v(m 2 S-I)
sand 2
air l'3xlO- 3 1'5 X 10-5
water 1 10-6
olive oil 1 10-4
v/V= W + w
a =[(W) - W!'CW),
b = !'(W)/V.
dw
W"dO+ (11+ ka)(w 2 -wI)=O,
where 8 is the angular displacement of the particle about the axis and WI
is a constant which should be determined. Hence determine W as a
function of 8 .
Show that if aW5 > g tan a, the particle does not lose contact with the
pipe immediately, but that contact will cease when 8 satisfies the equation
2
exp{2(11 + ka)8} 11 w -
=-1-% 1) .
ka\wl
g tan a
11
( Ans.: wi = 11 + ka -a--' W2 =wI +(W~ -wi)eX P{-2(I1+ka)8}.)
V10g 3
2(b - a)
and that the friction and air resistance alone can then bring it to rest in a
further time
V
----:-;-;:-tan- I (b - a)1/2
--
{a(b - a)}112 4a
4.14 A particle falls from rest under gravity in a medium offering a resistance
av + bv 2 per unit mass, where a and b are positive constants and v is the
speed of the particle. Find the velocity acquired and distance fallen in
time t.
b
T
( Ans.: v =!: tanh _!:.. y = !.IOg (COSh
2b ' b cosh a
T)_!!.!..,
2b
where c =(bg + !a 2 )112 , tanh a =a/2e, T =(ct + a). )
the same velocity as the air. The force exerted on the particle by the air is
gv2 /W 2 per unit mass, where g is the gravitational acceleration, v is the
speed of the particle relative to the air, and w is a constant.
Show that the maximum height, above its initial level, attained by the
particle is
provided u < w.
Show that the time taken to reach this height is wg- 1 tanh -1 (u/w).
What happens if u > w?
4.16 A particle is projected vertically upwards with initial speed u in a medium
offering a resistance whose magnitude is kv 2 , per unit mass, when
its speed is v. Prove th~t it returns to the starting point after a time
tan -1 (ku 2 /g) 11 2 + sinh -1 (ku 2 /g)1I2. Show that if ku 2 ~ g the time is
approximately
2U(1_ kU 2 ).
g 4g
4.17 A particle is projected vertically upwards with initial speed u and moves in
a medium offering a resistance kv4 , per unit mass, when its speed is v.
Prove that the greatest height reached above the point of projection is
ex/(4kg) 1 I 2 , where g tan2 ex = ku 4 • Show that if the particle has speed won
regaining the point of projection, g tanh2 ex = kW4.
4.18 A particle is projected from a fixed pOint 0 with speed u in a resisting
medium in which the force per unit mass opposing the motion at any
distance x from 0 is 2ex2 x times the speed of the particle, ex being
constant. No other forces act on the particle. Given that the maximum
distance of the particle from 0 is X, show by a dimensional argument that
X = ku 1l2 lex, where k is a numerical constant.
Calculate x explicitly as a function of time, prove that X exists, and
find k. Show that when x approaches X and is written in the form
x = (1 - €)X, where € ~ 1, the corresponding time t for the particle to
reach x is given approximately by
log(2/€)
t=
2exu 1l2
4.20 A particle of mass m is projected horizontally with initial speed u from the
top of a tower which stands on a horizontal plane. It is subject to a
constant gravitational force mg and a resisting force equal to -mkv when
its velocity is v. What is the height of the tower if the particle hits the
plane at a distance x from the base of the tower?
Let Xg be the distance from the base of the tower at which the particle
would hit the plane if there were no resistance. Prove that, approximately,
x = Xo (1 - kxo /3u) provided that kxo <Ii; u.
( Ans.:!
k
10g(_U
u-kx
)_ u gx .)
(i) Show that if the angle of inclination is such that tan ex = g/kv, the
particle eventually returns to its point of projection.
(ii) Show that the particle is directly above its point of projection at the
highest point in its trajectory if the angle of inclination is such that
g) log (1 + ku sin ex )
(1 + kusinex u cos ex
= 1 +--- .
g v
where
·f d17
1
~0
::r,
dr
·f d17.;;::: 0
1 """'.
dr
Exercises 117
4.23 The non-dimensional range t for a quadratic law of resistance, satisfies the
relations
when a = 0, and
when a= 0 and
f u ~d~
x = u R(~)' Y =
f
u
u ( .
Sill a-
f
1)
U gd~
~R(~)
) 11d11
R(11)'
R(v) per unit mass, where v is the speed of the projectile. The velocity
vector makes an angle If; with the horizontal. Initially v = u and If; = a.
(i) Show that
Hence show that the speed at the highest point of the trajectory is
Au cos a, where
(iii) Verify that the result obtained in (ii) agrees with that obtained from
the solution derived in Example 4.2.3 when a -+ 0, K2 =1= O.
(iv) Verify that the fractional variation in A, for K2 = 0·2 and
0";; a";; nl2 is about 7%.
(v) Show that the corresponding value of A, when R(v)=kv, is
(I + K I) -I , where K 1 = kg -I u sin a.
5 Impulse, Momentum,
Work and Energy
In this chapter we discuss certain integrals of the equation of motion which can
sometimes be used with advantage as alternatives to the basic equation.
Let a particle of mass m be subject to a force F and thereby experience an
acceleration dv/dt, where v is its velocity. The integral with respect to the time
of the equation of motion
F =m dv/dt (5.1.1)
gives immediately
f t2
t,
Fdt=m(v2 -Yd· (5.1.2)
The time integral of the force is called the impulse of the force. The product mv
is called the momentum. Thus equation (5.1.2) states that the impulse of the
force is equal to the change in the momentum of the particle.
Suppose that F is large in magnitude but acts for a short time interval only.
We can idealise the situation by letting F ....'>-<>0 and (t2 - t d -+ 0 in such a way
that the expression
-
F= lim f t2
F dt (5.1.3)
t2~t, t,
Example 5.1.1
Consider the impact of two particles. By Newton's third law the impulses of the
forces at contact are equal and opposite and so there is no net change in
120 Impulse, Momentum, Work and Energy [5.1
momentum. If the particles have masses ml, m2 and velocities Ul, U2 before
impact, their velocities VI , V2 after impact must be such that
One further condition is required to give Vi and V2 separately. If, for example,
the two particles coalesce on impact, we have
Example 5.1.2
A light inelastic string OAB has one end attached to a fixed point 0 and
particles of masses ml , m2 are attached to the string at A, B respectively. The
system rests on a horizontal table with each of the portions OA, AB of the string
straight and horizontal and inclined to each other at an angle 1T - a (a < 1T12), as
A _Uj
- m2
F B
W (~
Figure 5.1.1 Illustration for Example 5.1.2: (a) the impUlses; (b) the
velocities.
in Figure 5.1.1. The particle B is given an impulsive blow P in the plane of OAB
and in the direction perpendicular to and away from ~A. Calculate the instan-
taneous velocities acquired by m 1 and m2 .
In addiiion to the applied impulse P, there will be impulses 1\ in OA and 7'2
in AB, say, since the string is inelastic. * The velocity of m 1 will be perpendicular
* In an elastic string no such impulse is produced because the tension, being pro-
portional to the extension, cannot become large.
5.2] Work and Energy 121
These four equations are readily solved for the four unknowns u 1, U2, TI , T2 •
The solutions for uland U2 are
The next relation is obtained from the line integral of the basic equation of
motion (5.1.1); this integral is evaluated along the path actually traversed by the
particle. Let the position vector of the particle be r and let the initial and final
values of r be rl and r2' We have
(5.2.1)
The line integral of the force is called the work done by the force. The
product !mv2 is called the kinetic energy of the particle. Equation (5.2.1) states
that the work done by the force is equal to the change in kinetic energy of the
particle.
A special simple case arises when F is a constant force, say ma, so that the
particle is moving with constant acceleration a. Then
I r2
I,
F . dr = ma . (r2 - rl) (5.2.2)
Example 5.2.1
A parti~le slides up a fIxed rough slope inclined at an angle a to the horizontal.
We consider the subsequent motion of the particle given that its initial speed is
VI and that the coefficient of friction is p..
The sum of the components of the forces perpendicular to the plane is zero,
and the net force down the slope is mg sin a + F (see Fig. 5.2.1). Since the
particle is sliding, we have F= p.N = p.mg cos a, so that the force on the particle is
mg(sin a + p. cos a) down the slope and this is constant. Hence the work done by
this force when the particle travels a distance x up the slope is -mg(sin a
+ /1 cos a)x. If the particle then has speed v, it follows from (5.2.1) that
Alternatively one can use the fact that the constant force mg(sin a + J.1 cos a)
produces a constant acceleration -g(sin a + /1 cos a) in the direction of x increas-
ing. Equation (5.2.4) then follows from (2.3.8). The distance d travelled by the
particle up the slope is obtained from (5.2.4) by putting v = 0, and so
d=
v21
(5.2.5)
2g(sin a + p. cos a) .
Let US suppose that the frictional force is sufficiently small to allow the
particle to slide back down the plane. The magnitude of the frictional force will
remain the same, but the sign will change. The net force on the particle is now
mg(sin a - JJ. cos a) down the plane. When the particle reaches its initial position,
the work done by this force is mg(sin a - JJ. cos ex)d. Accordingly the net work
done during the whole motion is
mg(sin a - JJ. cos a)d - mg(sin a + JJ. cos a)d = -2mgp.d cos a. (5.2.6)
The net loss in kinetic energy is 2mgp.d cos a, and if V2 is the final speed, we have
Note that the net work done on the particle by the gravitational force mg is
zero. This is because the sign of this force does not change, therefore the
negative work done by mg in the upward motion exactly balances the positive
work done in the downward motion. All the contribution comes from the
frictional force, which does change its sign so that it always opposes the motion,
and hence does a negative amount of work leading to a loss in kinetic energy.
The two c1asses of dissipative forces and conservative forces do not exhaust
all the possibilities and there are others, in electromagnetism for example, which
do not fit into either c1ass. It is interesting to note the following implication for
a force field in which the force is a function only of position but which does not
have the property that the net work done along a closed path is zero. If the
direction of traverse is chosen appropriately, the net work done will be positive.
A particle under the influence of such a force could, by repeated traverse of such
a path, acquire unlimited kinetic energy without permanently changing its
position. If the gravitational field had this property we would have a readily
available source of unlimlted energy.
For the most part we shall be concerned with the dissipative and the
conservative types of force. The latter plays a particularly important part in
mechanics and its special propetties can be exploited with advantage.
For a conservative force field the work done in moving from Al to Az by any
path is equal and opposite to the work done in moving from Az to Al by any
path. EqUivalently the kinetic energy lost by a particle moving from AI to Az
under the influence of such a force is equal to the kinetic energy gained in
moving from Az to AI. It is convenient to introduce the idea of a potential
energy which is stored during the motion from AI to Az at the expense of the
kinetic energy. This potential energy is released if the particle returns to A I .
We define the potential energy at a point in a conservative field as the work
done by a conservative force in bringing the particle from the given point with
position vector r, to some standard point with, say, position vector ro. Thus the
potential energy V is given by
V= f ro
F. dr= -
f F. dr.
r
(5.3.1 )
r ro
F = -\7V.
Note the negative sign which arises from the way in which the potential energy is
defined.
If the particle moves from rl to r2, it follows that
irrespective of the path between rl and r2. This is the loss in potential energy.
By (5.2.1) it is equal to the gain in kinetic energy and so we have the
conservation equation
or !mv 2 + f
r
ro
F. dr = constant. (5.3.3)
The simplest type of conservative field is a uniform force field such as gravity.
Since mg is a constant vector, the potential energy is easily evaluated to be
f r
ro
mg. dr = mg. (ro - r) =mg(y - Yo), (5.3.4)
where y is the vertical coordinate of the particle (remember that g has the
component -g vertically upwards). If Yo is taken at the origin of coordinates,
(5.3.4) reduces to mgy and so
or
as in equation (2.4.4).
Example 1.1 0.1 shows that the force field f(r)r/r, called a central force field,
is conservative. The associated potential is
f ro fer)
- r . dr =
f Yo
fer) dr,
r r y
which clearly depends only on the distance from the origin. This type of force
field will be studied in more detail in Chapter 7.
It is important to note that the concept of potential energy is appropriate
only for conservative force fields, whereas equation (5.2.1), which equates the
work done by the force along the prescribed path of the particle to the increase
in kinetic energy, is quite general. Energy conservation is likewise restricted to
problems involving conservative force fields. In Example 5.2.1, which involves
dissipative frictional forces, the particle loses kinetic energy for which there is no
compensation within the framework of Newtonian mechanics. In a wider con-
text, however, energy conservation, which has been presented here as a de-
duction from the laws of motion, is regarded as a very basic law. Suitably
modified, It is retained throughout the whole of modern physics. In Example
5.2.1 the loss of kinetic energy of the particle is equated to a dissipation into
heat through the mechanism of the frictional forces. This is a form of energy not
included in the mechanics of a single particle although Example 8.3.3 shows how
it might be accounted for at a molecular level.
Example 5.3.1
A particle of mass m2 hangs vertically from a light inextensible string, which
passes over a smooth peg and is attached to a ring of mass ml (<m 2 ), free to
slide on a smooth vertical wire. The horizontal distance between the peg and the
126 Impulse, Momentum, Work and Energy [5.3
. d __ ·
1
YI 1
T
Y2
N_ J T
j
1
mIg
m2g
1
Figure 5.3.1 Illustration for Example 5.3.1.
wire is d. Initially that part of the string between m 1 and the peg is horizontal. If
ml is released from rest in this position find the maximum depth to which it
descends.
The configuration and the forces acting on m 1 and m2 are shown in Figure
5.3.1. The vertical distances of ml and m2 below the peg are denoted by Yl and
Yz respectively. Because the wire is smooth, the reaction Nbetween it and ml is
horizontal. Hence there is no work done by N during the motion of ml' There
is, however, a tension T in the string which does work when the calculation is
made separately for each mass. Moreover T is one of the unknown quantities in
this problem
Let us calculate the work equation for each mass separately. We have
(5.3.5)
for m2' Here e is the inclination to the vertical of that part of the string between
the ring and the peg, and I is the total length of the string, so that Yz = I - d
initially. We also have the relations
_ Yl
cos e - (2 2 liZ' (5.3.7)
\d +yr)
(5.3.8)
and so (5.3.9)
5.3] Potential Energy 127
fo T cos 8 dY1 = -
Yt jY2
I-d
T dY2.
We can therefore eliminate Tby the addition of (5.3.5) and (5.3.6) to get
or (5.3.11)
(5.3.12)
One point to note here is that (5.3.12) is obtained from (5.3.11) by squaring
each side of the relation. Hence the solution of (5.3.11) satisfies (5.3.12), but
the converse is not necessarily true. In fact (5.3.12) satisfies (5.3.11) only if
m2 > mi' If m2 < m 1 there is no solution of (5.3.11) and in this case the ring,
once released, never comes to rest while the string remains suspended over the
peg.
The result that the tension in the string does no net work during the displace·
ment of the system is capable of wider application in that it depends on the
following conditions, which are satisfied for many other configurations involving
light, elastic strings:
(i) the tensions at the two ends of the string are equal in magnitude, because
the peg is smooth and the mass of the string is negligible;
(li) when the two ends of the string suffer small displacements, the components
of those displacements along the length of the string (and hence parallel
to T) are equal in magnitude. This follows immediately from the assumption
that the string is inextensible, and ensures the validity of (5.3.9).
128 Impulse, Momentum, Work and Energy [5.4
5.4 POWER
fF . dr = f F . v dt,
the time derivative of the work done, or the rate of working, is F . v. This is
referred to as power. The unit of power in the SI system is the watt (W), or
1 J S-l, and its dimensions are ML2 T- 3 • The kilowatt hour (1 kWh =
3·6 x 106 J) is often used as a practical measure of work or energy, particularly
with reference to heat. The latter is also a form of energy although it plays no
part in Newtonian mechanics.
Power is a useful practical measure of a machine's capacity for doing work.
For example the powers of a small electric motor and a typical automobile
engine are likely to be in the rangesO·l-l kWand 10-100 kW respectively.
Example 5.4.1
Water of density p is pumped from a depth d and delivered through a pipe
of cross-section A. What is the power required to deliver a volume V per unit
time?
There are two parts to this calculation. The first part represents the work
done in raising a mass of water p V from a depth d in unit time, which is p Vgd.
In addition there is the work required to generate the requisite kinetic energy. In
unit time a volume Vof water is delivered through a pipe of cross-section A. The
speed is therefore VIA and since the mass is p V the kinetic energy is 1P V( V/A)2 .
v
The total work per unit time, which is the power required, is p Vgd + 1P 3 fA 2 •
If we substitute the particular values g = 9.81 m s -2, P = 10 3 kg m -3 ,
d = 10 m, A = 10- 2 m 2 , V= 10- 1 m 3 S-l we get (9·81 + 5)10 3 W or about
15 kW.
Example 5.4.2
A train of mass m runs on a level track at constant speed u. We assume that the
power exerted by the engine is constant and that the resistance to the motion
arises from two sources. The air resistance is modelled by a term K v2 , where v is
the speed and K is a constant, and frictional resistance is assumed to be equal
to a constant fraction {3 of the weight of the train. We investigate the change in
the speed of the train when it slips a coach of mass (X}'J'l.
Initially the total resistance is Ku 2 + {3mg and, since the speed is constant, this
must be balanced by the tractive force. The power is therefore (Ku 2 + {3mg)u
and is constant. When the coach is slipped, let the subsequent speed be v. The
tractive force is then (Ku 3 + {3mgu )fv and the resistance is Kv 2 + {3(1 - cx)mg. It
follows that the equation of motion is
dv Ku 3 + {3mgu
(1 - cx)m - = Kv 2 - {3(1 - cx)mg,
dt v
Exercises 129
dv K ~g
or v- = - (V 3 - U3 ) + - - (u - v + 00).
dt m(1 - a) 1- a
and approximate this equation by retaining only the first order terms in o. The
result is a linear equation of the form
do=
- -T
-1
(0 - e)
dt '
where T -1 ~g ( 1 +
=- 3KU 2 )
e = - - -a- - - -
u m~g(1- a)' 1 - a + 3Ku 2 /m~g .
The solution is
and we can now see more clearly the conditions for which the approximation is
valid. As t -+ 00,0 -+ e and the terminal speed is u(1 + e). In order that 0 should
be uniformly small, it is necessary that e ~ 1, or a ~ 1 + 3Ku 2 /(m~g). The
parameter T, which has the dimensions of time, is a measure of the time interval
over which the speed changes by a significant proportion of itself.
EXERCISES
5.1 A golf ball of mass m is struck by a club, the contact lasting for a time T,
and is thereby given a speed v. Assume that, during the contact, the force
increases at a constant rate from zero to a maximum and then decreases at
a constant rate from this maximum to zero. Find: (i) the impulse of the
blow, (ii) the maximum force on the ball.
(Ans.: mv, 2mv/T.)
5.2 A light rod rests on a smooth horizontal plane and has two particles, of
masses m 1 and m2 respectively, attached to it at A and B. A horizontal
impulse F is given to A in a direction making an angle a with AB. Find the
velocities of A and B, and show that the kinetic energy imparted to the
system is equal to 1Fw, where w is the component of the velocity of A in
the direction of the impulse.
130 Impulse, Momentum, Work and Energy
What would be the effect on the initial motion if the plane were rough?
(Ans.: A component F cos al(m1 + m2) along AB for both particles and a
component F sin aIm 1 perpendicular to AB for A.)
5.3 A man of mass m 1 is standing in a stationary lift of mass m2, the
counterpoise being of mass m 1 + m2. Suddenly the man jumps with an
impulse which would raise him to a height h if he were jumping from the
ground.
Given that the inertial and frictional effects in the cable and pulley
which connect the lift to the counterpoise may be neglected, calculate the
velocity of the man, relative to the lift, immediately after the impulse.
Find also the subsequent acceleration of the man relative to the lift.
Deduce that the height above the floor of the lift to which the man
jumps is 2h(m 1 + m2 )/(m 1 + 2m2).
5.4 If the earth is regarded as a sphere, of radius re, the gravitational force on a
particle of mass m at a distance r (>r e ) from the centre of the earth is
mg(re/r )2.
Show that the work done in raising the particle from the earth's surface
to a height re above the earth's surface is !mgre.
5.5 A mass of 200 kg is suspended at the lower end of a light vertical rope and
is being hauled up vertically. Initially the mass is at rest and the pull on the
rope is 300 kg-wt. The pull diminishes uniformly at the rate of 6 kg-wt for
each metre through which the mass is raised.
Find the work done by the pull on the rope, and the speed of the mass
after it has been raised 20 m.
(Ans.: 4800g J, (8g)1I2 m S-l.)
5.6 The tension in a stretched elastic string is given to be "AxIl in magnitude,
where I is the natural length of the string, x is the extension beyond its
natural length, and A is a constant for a particular string.
x
Show that the work done in increasing from Xl to X2 is! A(X~ - xi)/1.
A uniform elastic string is stretched in a straight line between two
points A, B which are fixed. A point C of the string is then displaced so
that the components of the displacement are x along AB and y perpen-
dicular to AB.
Before the displacement let AC = a, CB = b and let the tension in the
string be T. Let the unstretched length be k(a + b), where 0 < k < 1.
Show that if the string remains taut during the displacement, the work
done is
T(a
2ab
+ (~+
b)
1- k Y
2) .
Exercises 131
When the tensions in AC and CB are the same after displacement, show
that the work done can be written in the form
Tk(a+b){ r2 ( }
r2 )112 +1.
- - - 1+-
1- k 2kab ab
5.7 A particle moves along the x-axis under the action of a force -11/x2
(11 > 0) per unit mass in the x direction. Initially x> O. Establish the
energy equation in the form
where E is a constant.
(i) Show that if E = -11/2a (a> 0), the particle comes to rest and then
falls into the origin. Use the substitution x = a(1 + cos 8) to obtain a
parametric relation between x and t.
(ii) Show that if E = 0, the particle moves off to infinity and
(iii) For E =11/2a (a> 0), use the substitution x = a(cosh cp - 1) to obtain
a parametric relation between x and t.
5.8 A parcel, of mass m, rests on a horizontal conveyor belt. For time t;;" 0 the
belt is constrained to move with a constant horizontal velocity of magni-
tude u. The coefficient of friction between the parcel and the belt is 11.
Find:
(i) the time that elapses before the parcel is again at rest relative to the
belt;
(ii) the distance the particle slides relative to the belt;
(iii) the energy dissipated during this sliding;
(iv) the impulse of the frictional force.
F = -Fo(xi + yj + ck)/l
at time t. Show that its motion is compatible with that arising when it is
subject to the force F, with appropriate values for F 0 and c. Show that the
magnitude of the force remains constant during the motion of the particle.
132 Impulse, Momentum, Work and Energy
F = Fo(yi + xj + zk)II,
where Fa and I are constants. Show that this is a conservative field. Find
the potential energy and deduce the energy equation
v; + 4 VI = 0, V~' + 16 V2 = 0, ....
and, hence, that the speed v of the particle is given by the equation
d2 q 3v' 2
-+-q=O
dz 2 4v2 .
Find the reaction between m I and the cylinder at this instant. Hence
show that m I will certainly remain in contact with the cylinder initially if
m2 < 3m l.
(Ans.: mlg{(3ml + m2)sin 0 - 2m20}/(ml + m2).)
5.15 A light string rests over two smooth pegs, which are at a distance 2a apart
and at the same horizontal level. A smooth ring, of mass m 3, slides freely
on the string and two particles, of masses ml, m2 respectively, are
attached to the ends of the string.
Initially the system is held with that part of the string between the pegs
horizontal and the ring midway between the pegs. The portions of the
string beyond the pegs are vertical with the particles hanging at rest.
The system is then released from rest. Let Yl, Y2, Y3 denote, respect-
ively, the distances whi~h m 1, m2, m3 descend below their initial levels, so
that Yl =Y2 =Y3 = 0 initially.
134 Impulse, Momentum, Work and Energy
2ak
1 - k2'
mu 2 mu 2
(! log 3 - if; log 5 - ~ log 2) - = 0'25 -
Po Po '
(6.1.1)
Here n > 0 is a constant and the negative sign ensures that the force is a restoring
force acting always towards the equilibrium value x = 0, which is a particular
solution of (6.1.1). Such a force is conservative and can be derived from a
potential energy !n 2 x2 per unit mass. Hence there is an energy equation of the
form
(6.1.2)
say, where the constant total energy has been written as !n 2 a2 • It is easily
verified that (6.1.1) follows from (6.1.2) by differentiation. Equation (6.1.2) is
equivalent to
nt + a = ± f (2
dx
\a - x
2 )112 = ±cos -I
x
-
a
which, for a linear differential equation, can be combined linearly to give the
general solution
The motion described by the basic equation (6.1.1) with the corresponding
solution (6.1.4) or (6.1.5) is called simple harmonic motion. It is a pure
sinusoidal displacement in time, with a single frequency n. Note that a displace-
ment such as
T=ax (6.1.6)
T= "Ne/I (6.1.7)
Example 6.1.1
A light spring has one end attached to a fixed point and a mass m suspended
from the free end. The equilibrium extension of the spring is c and the
equilibrium tension is T() as in Figure 6.1.1. Therefore
To = ac=mg.
m5:=mg-T,
and T=a(c+x)=mg+ax.
;
I
G:,.
c.o
c.o
c.o
G
<0
~ ~
fg G
G
TGr
T~tTo G
G
i~ G
xG
~ +~ .
mg
~
mg
time is 2rr(c/g)1/2. The amplitude and phase can be determined once the initial
conditions have been prescribed.
~xample 6.1.2
A light elastic string has one end attached to a fixed point and a mass m
suspended from the free end. The equilibrium extension of the string is c. The
mass is pulled down a further distance 2c and then released. Show that the
period of the subsequent motion is 2(C/g)1I2(3 1/ 2 + 2rr/3).
Initially the motion is simple harmonic, as in the previous example, with
frequency n =(g/c) , / 2, amplitude 2c and zero phase. Hence x =2c cos nt, where
x is measured from the eqUilibrium position (see Fig. 6.1.1). This motion persists
until x = -c. This is when cos nt = -! and first arises when nt =2rr/3, so that
-2rr + (3C)
- 112 = (C)
- 1/2 (2rr
- + 3 112 ) •
3n g g 3
It will take an equal time to return to its original position after which the
motion will repeat itself.
Note that although the motion is periodic, and although the first part of the
motion is governed by an equation of the form (6.1.1), it is not simple harmonic
motion because for part of the time the particle is moving freely under gravity.
If, however, the string is replaced by a light elastic spring of the same length and
stiffness, the relation x = 2c cos nt holds throughout the motion.
~xample 6.1. 3
A particle of mass m is connected by a light spring of natural length I to a point
o on a rough horizontal table. Let us consider the equation of motion when the
particle is at a distance 1+ x from O. Here the magnitude of the frictional force
is pmg, but the direction will depend on the direction of motion ofthe particle.
Suppose first that x is increasing. The equation of motion is
"lu:
rnX = -T - J1.mg= - T- p.mg,
or (6.1.8)
where T = "lu:/l is the tension and F = p.m.g is the frictional force (see Fig. 6.1.2).
Equation (6.1.8) is similar to (6.1.1) with n2 = A/ml and the variable changed to
(x + p.mgl/A). The motion is thus similar to a simple harmonic oscillation but
140 Oscillations [6.2
or ~
dt 2
(x _ pmgl) = _
X
~(x
ml
_pmgl) .
X
(6.1.9)
The centre of the motion thus switches to x =pmgljA. This switch of the centre
of the motion each time the direction of motion changes decreases the effective
amplitude of oscillation until eventually the particle comes to rest within a
distance pmgljA on either side of the origin x = O. When this happens, the
restoring force -Axil is smaller in magnitude than the limiting frictional
force pmg and the particle will remain permanently at rest (see Fig. 6.1.2).
Note that the ultimate position of rest can be anywhere in the interval
-pmgllX ..;; x ..;; pmgllX depending on the initial conditions.
6.2 DAMPEDOSOLLATIONS
(6.2.1 )
This is still a linear differential equation with constant coefticients, and to solve
it we look for solutions of the form
(6.2.2)
6.2] Damped Oscillations 141
have been chosen to give a real form of the solution (6.2.4). In this case the
damping is conveniently described as light. The motion is oscillatory, but the
amplitude decays according to the factor e- kt / 2 and so tends to zero as t tends
to infinity. A typical example is a pendulum oscillating in air.
For k > 2n, the solution can be written directly as
(6.2.5)
where Xl and X2 « AI) are now both real and positive. This solution represents
exponential decay, without oscillation, to x =O. If Al/A2 < -1 it will pass
through x = 0 just once, otherwise it will decay monotonically to zero. Such
heavy damping might arise from the motion of a pendulum in a very viscous
medium.
This approach fails to give two independent solutions in the critical case
k =2n, when Al and A2 are both equal to !k. We can see more clearly what
happens in this case if equation (6.2.1) is rewritten in the eqUivalent form
(6.2.6)
With xe kt / 2 as the dependent variable, (6.2.6) is of the form (6.1.1) if k < 2n, or
(3.3.10) if k > 2n. If k =2n it is easily integrated to give
or (6.2.7)
142 Oscillations [6.3
(6.3.1)
The solution of equations such as (6.3.1) can be divided into two parts. The
complementary function, which is the solution of the equation with f(t) omit-
ted, is exactly the solution discussed above in §6.2. The second part is the
particular integral, which can be any particular solution of the full equation.
The most general solution is the sum of the complementary function and
the particular integral.
There are standard techniques for calculating particular integrals. Here we
simply verify that a solution of the following form is possible. Let
X = Jo X(i - 7)f(7) d7
t
(6.3.2)
x = X(O)f(t) + r
o
X(t - 7)f(7) d7, (6.3.3)
(6.3.5)
(6.3.6)
where z = x + iy is a complex function of t. Since an equation involving complex
quantities can only be satisfied if the real and imaginary parts of the two sides of
the equation are separately equal, it follows that if z is a solution of (6.3.6) then
the real part x is a solution of (6.3.1) withf= fo cos pt (and y is also a solution
of (6.3.1) withf= fo sin pt).
A particular integral of (6.3.6) is z = zoe iPt , where Zo is a constant satisfying
the equation
144 Oscillations [6.3
j; eipt
It follows that z =x + iy = -::-~o--::-_ _
n 2 _ p2 + ikp
to cos (pt - ~)
(6.3.7)
and so
kp
where tan ~ = 2 2· (6.3.8)
n -p
This is called the forced oscillation to distinguish it from the free oscillation,
represented by the complementary function, to which the particle can respond
without the help of the forcing term. As in all resisted oscillations, the free, or
transient, oscillations eventually decay and ultimately the motion is just that
represented by (6.3.7). The frequency of this forced oscillation is that of the
forcing term, but there is a change in amplitude and phase. When k is small there
is a significant increase in amplitude for a narrow band of frequencies near the
frequency n of free, umesisted oscillations, that is when (n - p) is small. Indeed
there is a singularity in the amplitude for n =p and k =o. This phenomenon is
called resonance and can be a source of potential danger in mechanical struc-
tures. In fact in many applications involving spring supports and periodic forcing
terms, it is desirable to include sufficient damping unless the frequency of natural
vibrations differs substantially from the frequencies of the forced oscillations
that might be encountered. One practical example is the vibrations of a motor
with a resilient mounting. Equation (6.3.7) shows that the amplitude of the
forced oscillation can be written
to
dx -prJ .)
fo cos pt dt ={(n2 _ p2)2 + k2p2 }1/2 cos pt sm(pt - (j ,
Hence the work done in one complete period °<; t <; 21r/p is
Jo21r/P fo cos pt dxdt dt ={(n2 _ 1rfJ sin (j 1rkpfJ
p2)2 + k2p2 }1/2 =(n2 _ p2)2 + k2p2'
(6.3.10)
where (6.3.8) has been used. For small values of k this is also small, and
approximately proportional to k, except near resonance where there is a sharp
increase in the supply of energy. When n =p this supply is actually proportional
to k- 1 , so giving rise to large amplitudes.
The case for which n = p and k = 0, namely
x =b.
n
J sin net - r) cos
0
t nT dr
The solution is no longer periodic, but has an amplitude which grows linearly
with time. This behaviour is typical of undamped resonance. In practice, how-
ever, there will invariably be some damping present and often there will be
146 Oscillations [6.4
non-linear effects brought into playas the amplitude grows, thereby modifying
the ultimate behaviour.
There are many physical applications of the mathematics relating to the inter-
action of oscillating systems. We introduce the ideas with a discussion of a pair
of equations which typify such interactions, namely
(6.4.1)
The terms involving the constants k12 and k21 are those arising from the
interaction between the two systems. When they are absent, the equations
become uncoupled and are then of a type already discussed.
There are standard techniques for the integration of equations such as (6.4.1),
but it is instructive to see how the solution can be built up in the following way.
Equations (6.4.1) can be combined linearly to give
(6.4.2)
(6.4.3)
(6.4.4)
This equation for the single variable Xl + KX2 is of a known form. Now equation
(6.4.3) can be written as a quadratic equation for K, namelY
(6.4.5)
(6.4.6)
6.4] Coupled Oscillations 147
For these two values of", (6.4.4) yields a pair of equations for x I +" IX2 and
XI + "2X2, which can be used as alternatives to the original pair (6.4.1). A
discussion of these equations in the general case presents no particular difficulty.
However, we shall concentrate on the case for which the frequencies nl , n2
given by
(6.4.7)
say, and this pair of equations can be solved to give XI and X2 as a linear
combination of two harmonic oscillations.
Alternatively we can solve (6.4.1) by looking for solutions such that both XI
and X2 are proportional to e~t; such a procedure has more general applicability
for linear equations with constant coefficients. If the solutions are known to be
purely oscillatory, this fact can be recognised from the beginning by choosing
A= in, say, where n is real. The procedure is then equivalent to looking for
possible solutions such that
(6.4.9)
XI kl2 n 2 - k22
or, equivalently, -::: = (6.4.11)
X2 n 2 - kl1 k21
This relation yields a quadratic equation for n 2 , whose roots are given by (6.4.7)
without the intermediate calculation of ". When nl and n2 are known, the
general solution can be written as a linear combination of the corresponding
solutions of (6.4.9), say
148 Oscillations [6.4
where the relation for Xl follows with the help of (6.4.11). That equations
(6.4.12) are equivalent to (6.4.8) may be verified directly.
We note that the oscillation frequencies are the same for both the displace-
ments Xl and Xl, but the amplitUdes are different. In general the actual
oscillations represented by (6.4.12) will not be harmonic, indeed they will not
even be periodic if nl/nl is irrational. But if one of the arbitrary constants al or
al happens to be zero, the resulting oscillation is harmonic with the same
frequency for both coordinates. Such an oscillation is called a normal mode of
vibration. Here there are two possible normal modes of vibration, represented by
(6.4.12) with either al =0 or al =0, and the system can be made to vibrate in
one or other normal mode by suitable choice of the initial conditions. In general,
however, the motion will be given by a superposition of the two normal modes
of vibration.
There is one special case for which the above arguments fail, namely when
(6.4.13)
Equations (6.4.6) and (6.4.7) then yield only one value for" = (kll - kl d/
2kl I and one associated value for n ={(k2 2 + kl d/2}1/2. This will yield only one
mode of oscillation and is not sufficient information to obtain the general
solution. Although we shall not use this case, a method of solution is indicated
for mathematical completeness.
The above values of" and n give rise to one relation of the form (6.4.8), say
(6.4.15)
k 21 a . , )
X2 = - - - t sm(nt + al) + a cos(nt + a2 . (6.4.16)
2n
The associated expression for XI then follows immediately from (6.4.14).
Example 6.4.1
Two particles m 1 ,ml are connected by a light elastic string, of natural length I,
which passes over a smooth peg, as in Figure 6.4.1. The particles are released
6.4] Coupled Oscillations 149
from a rest position in which a length 1/2 of the string is hanging vertically on
either side of the peg.
To discuss the subsequent motion, let ml, m2 be at distances Xl, X2
respectively from the peg. Then
(6.4.18)
(6.4.20)
Equation (6.4.19) is just (6.1.1) with an appropriate change ofvariable and with
2 h(ml + m2)
n = .
lmlm2
150 Oscillations [6.4
2g1mlm2
Xl +X2 = 1+ ) (1- cosnt). (6.4.21)
"A.(ml + m2
Example 6.4.2
Three light springs AB, BC, CD are joined end to end and stretched between two
fixed points A, D on a smooth horizontal plane. Masses ml, m2 are fixed to the
springs at B and C and the system is allowed to oscillate in the line AD (see Fig.
6.4.2).
Let the displacements of the particles B and C from their equilibrium
positions be XI and X2 respectively. Let the spring stiffnesses for AB, BC and CD
be aI, a and a2 respectively. In order to discuss the motions of Band C we
require only the changes in the tensions of the springs, since the equilibrium
values must cancel. For particle B, the net force is -alXl - a(xi - X2) and for
Cis -a2x2 + a(xi - X2); The equations of motion are therefore
These equations are solved by looking for solutions such that both B and C have
oscillations of the same frequency, that is solutions for which, simultaneously,
(6.4.23)
(6.4.25)
This is a quadratic equation for n2 which in general has two roots, say nt and
n~. For each value there is a possible solution, with the ratio Xl /X2 fixed and
given by (6.4.24). These solutions are the normal modes of oscillation. The most
general solution of (6.4.22) is a linear combination of these two solutions, say
(6.4.26)
(6.4.27)
where ai, a2, al , a2 are arbitrary constants to be determined from the initial
conditions. If the latter are adjusted so that al or a2 is zero then the system will
oscillate in a normal mode with a definite frequency of oscillation. In general,
however, this will not be the case, and the solution is a combination of the two
modes of oscillation.
Consider the special case ml = m2, al = a2 so that, from (6.4.25),
(6.4.28)
and (6.4.29)
(6.4.30)
Here the first normal mode is such that the two particles have the same
amplitude al and the same frequency nl but are 11" out of phase so that they
always move in opposite directions. In the second slower normal mode the
amplitude, frequency and phase are all identical. In this mode the spring BC
exerts no influence. For arbitrary initial conditions the motion is a superposition
of these two oscillations. Suppose for example that the initial conditions are
whent=O.
(6.4.31)
X2 = 2a sin ( nl -2 n2)
t sin (nl + n2 t ) .
--2- (6.4.32)
and ifa= 0, we see that the solution iSXl = 2a cos (at/ml )1/2 t, X2 = 0 as expected.
152 Oscillations [6.4
carrier wave
,r beat oscillation
--
Figure 6.4.3 Superimposed oscillations of nearly equal frequency.
If 0 ~ 01, the frequency !enl + n2) approximates to (odmd l/2 • But note that
there is still the slowly varying factor cos! (n 1 - n2 )t, and provided that
(n 1 - n2) is not zero this will eventually alter the solution for x 1 substantially.
Furthermore X2 will eventually acquire a substantial oscillation. This is an
example of a situation in which a small effect can accumulate over a sufficient
length of time to produce a Significant modification of the solution. A vibration
of the kind described by (6.4.31) or (6.4.32) is called a modulated oscillation.
The slow oscillation superimposed on the modulated oscillation is called a beat
oscillation and can be used in tuning musical instruments. The combined effect
of two notes of approximately equal frequency has an amplitude which slowly
rises and falls as shown in Figure 6.4.3. These surges in volume are easily
detected by the ear and tuning is effected by arranging that the frequency of the
beat oscillation tends to zero. This type of oscillation also has applications in the
theory of electromagnetic waves. Radio waves are transmitted at frequencies of
the order of 106 oscillations per second. These are called carrier waves. But the
actual information is transmitted by modulating the carrier wave at relatively
lower acoustic frequencies, not more than 104 oscillations per second. The
receiving apparatus then separates the acoustic oscillation from the carrier wave
for further use.
Example 6.4.3
The calculation of normal modes of oscillation can be extended to more
complex systems. We consider here the motion of r equal particles fastened at
Yk-l
+.
Figure 6.4.4 Illustration for Example 6.4.3: transverse oscillations of particles
on a stretched string.
6.4 ] Coupled Oscillations 153
equal distances to a string of length (r + 1)1. Both ends of the string are fIxed and
the only force influencing the motion of the particles is assumed to be the
tension T in the string.
We consider small lateral displacements Yl ,Yz, .. . ,Yr of the particles. Then
the change in the tension T from its equilibrium value To will be small and we
can write, correct to the fIrst order,
as the equation of motion of the kth particle. To fInd the normal modes we look
for solutions for which
Yo = Yr+l = O.
so that
cos~= l-!a.
asine=O, asin{(r+l)~+e}=O,
and so E=0 and siner + 1)~ = O. Hence there are r normal oscillations
154 Os cilla tip ns [6.5
krr
~ =r + l' (k =1,2, ... ,r).
The corresponding frequencies n can be calculated from
tx + vex) = E,
2 (6.5 .1)
where E is the total energy which can be determined from the initial conditions.
Note that the acceleration is given by - V'(x) and that, if the acceleration is zero
at X=Xo, then V'(xo)=O. If the initial conditions are such that, in addition,
E = V(xo), both the velocity and acceleration are zero at x = Xo and the particle
can rest in equilibrium. Otherwise the details of the motion are deducible from
the relation
2
112
t
=+
-
f(E _ V(X»112
dx
' (6.5.2)
where b < a and ljJ(x) is strictly positive and single-valued in the range b ~ x ~ a.
Then the motion is confined to this range, since only values of E - Vex) which
are non-negative are appropriate. Equation (6.5.1) now becomes
occurs in one complete oscillation. The periodic time T is the time for one
complete oscillation from b to a and back again. From (6.5.2), with due
attention paid to sign, T is given by
21!2T= fa dx
b (E- V(X»1/2 r (E - ~X»1I2 = s: (E _ 2~»112 .
(6.5.5)
(6.5 .6)
(6.5.7)
or (6.5.8)
(6.5.11)
156 Oscillations [6.5
Suppose first that y"(xo) > 0 so that V has a minimum at x =xo. It is clear by
comparison with (6.5.6), which represents simple harmonic motion, that ~ is
then confined to the range I ~ I <; E and that the perturbation is a simple
harmonic oscillation given by
T= 21/'(V"(XO))-1I2. (6.5.13)
Iii
/
mgSina/
Figure 6.5.1 Illustration for Example 6.5.1: the simple pendulum.
6.5) Non-Linear Oscillations 157
to·2 = g1cos 0 + E,
which is the energy equation. We write this in the alternative form
(6.5.16)
1 fa dO
T=~ -a (sin 2 ta- sin 2 !O)1/2 .
(6.5.17)
The integral is called an elliptic integral. The standard form is obtained with the
transformation
sin !O = sin ta sin IjJ = a sin 1jJ, (6.5.18)
and if we write To = 2rr/n as the periodic time for small oscillations, equation
(6.5 .17) can be written
T _2 J. 1r / 2 dljJ 2K(a)
(6.5.19)
To - -; 0 (1 - a2 sin 2 1/»112 - rr
where K(a) is called the complete elliptic integral of the first kind. It can be
found tabulated in all the more comprehensive tables of functions. Note that
K(a)~rr/2 asa~O
158 Oscillations [6.5
r 1f /2
= J (1 + !a 2 sin2 ef> + ... ) def> = !1I"(l + ia2 + ... ), (6.5.20)
o
so that, approximately,
(6.5.21)
When a > 1, e
never becomes zero and the pendulum swings in complete
circles rather than to and fro. Here the periodic time is the time for e to increase
by 211", and so
T
To
(6.5.22)
where the substitution e/2 = ef> has been used. Note that
T 1
asa~oo.
To 2a
1·0 I-----~======F======~--___._j 1!
a
1!12
O·SI-----------I----F----------I
1!14
()os 1 1·5
T1To
Figure 6.5.2 Variation of the periodic time with a for a simple pendulum.
6.5] Non-Linear Oscillations 159
Figure 6.5.2 shows the variation of TITo with a, produced from tables of the
function K. When Q = 7r/4(a = 0·383), for example, the periodic time is 1·04To
so that there is a 4% error in using the small amplitude approximation To. The
periodic time has increased to 1·17To at Q = 7r/2(a = 0·707) and tends to infmity
as Q -+ 7r(a -+ I). When a = 1 ()16 the periodic time is again To, and for a > 5 the
periodic time is (20) -1 with an error of less that 1%.
Example 6.5.2
A bead B, of mass m, is free to slide on a fIXed horizontal smooth wire. It is
constrained by two springs AB, CB, each of natural length I and modulus A,
which are attached to the bead and to fIXed points A and C (see Fig. 6.5 .3). The
bead can rest in equilibrium with AB =BC and ABC perpendicular to the wire.
Let us consider the possible oscillations of the bead on the wire.
We first note that there is an energy equation since the tension in a spring is
conservative. If the tension is AyII when the spring is stretched a distance y, the
potential energy is
(6.5.23)
Let AC =2b, and let x denote the displacement of the bead from AC. Then
each spring is stretched a distance (b 2 +X2)1/2 -I and the total potential
energy is, from (6.5.23),
(6.5.24)
(6.5.25)
Let us consider the equilibrium positions and their stability. From (6.5.24)
we have that
(6.5.26)
(6.5.27)
V"(O) = 2A
ml
(1 ~ ~ )
b
(6.5.28)
(6.5.29)
mbl )1/2
(
21T 2A(b -I)
about x = 0 and
(6.5.31)
(6.5.32)
where the substitution x = a cos 1> has been used. The function K(2 -1/2) is the
complete elliptic integral with argument r 1f2 (see equation (6.5.19» and has
the value 1·854. Hence the periodic time is
Unlike the simple harmonic case, the periodic time is now inversely proportional
to the amplitude.
EXERCISES
6.1 Verify that x =a sinmnt describes simple harmonic motion, about a suit-
able origin, for m = 1, 2 but not for m ;;;., 3.
6.2 A mass is suspended at the lower end of a light, perfectly elastic vertical
string whose upper end is fixed. The extension of the spring is 0'06 m
when the mass is in its equilibrium position C. The mass is pulled up
0'02 m above C and set in motion with an initial upward speed of
OA m s -}. Find the potential energy in any position and hence use
energy considerations to determine:
Explain why the last two results can be deduced without detailed
calculation.
(Ans.: 0'037 m above C, 0'4 m s -} , C.)
6.3 A particle P is constrained to move along the x-axis under the several
forces krPP r , r = 1, 2, ... , n, where kr is a constant scalar and Pr is a point
on the x-axis.
162 Oscillations
X Xl+ X2 X
2- , ---=:::-.
n X Xl +X2 X
(ii) At a given instant the displacement of the particle from the equi-
librium position 0 is x, and it is moving away from 0 with speed v.
Show that it next reaches 0 after a time
l(
;;-1T-tan- 1 nx)
-; .
(iii) At the instant that the particle is moving away from 0 with speed v, it
is given an impulse which instantaneously changes the speed from v to
v + ov, where ov ~ v. Show that the time to reach 0 alters by
ov(n2 a2 - v 2 )1/2 /a 2 n 3 approximately, where a is the amplitude.
6.6 Two light springs, of the same natural length I and of stiffnesses 01,02,
can be combined in parallel to form a combination also having the natural
length I. Show that a particle of mass m, when suspended from such a
combination, oscillates with a periodic time 21T{ m/( 0 1 + 02)} 1/2.
Show that if the springs are combined in series to form a combination
having the natural length 2/, the particle will then oscillate with a periodic
time 21T{m(oi 1 + 02"1 )}1/2.
Which is the longer period?
6.7 A particle of mass m is attached to one end of an elastic string, of natural
length I, whose other end is fixed at a point O. The particle is allowed to
fall from rest at 0 and the maximum depth of the particle below 0 is
written in the form I cot 2 !8, where 8 is a positive acute angle. Show that
the particle attains this depth in a time (21/g)1!2 { 1+ (1T - 8)cot 8}. Show
that the modulus of elasticity of the string is !mg tan2 8.
6.8 A particle is attached to the mid point of a light elastic string of natural
length I. The ends of the string are attached to fixed points A and B, A
being at a height 21 vertically above B, and in eqUilibrium the particle rests
at a depth 51/4 below A. The particle is projected vertically downwards
from this position with speed (gl)I/2.
Prove that the lower string slackens after a time 1TU/g)I/2 /12, and that
the particle comes to rest after a further time (j(l/2g)I/2 , where (j is the
acute angle defined by tan (j = (3/2)1/2.
Exercises 163
6.9 A light spring, of stiffness 01, is fixed at one end and a smooth pulley of
negligible mass is suspended from the other. Over the pulley passes a light
elastic string of stiffness 02' One end of the string is fixed to the ground,
from the other end is suspended a particle of mass m. In equilibrium the
spring and the two portions of the string are vertical. Show that, provided
the string remains taut and vertical, the particle can perform simple
harmonic oscillations with period
6.1 °
A particle, of mass m, moves along the x-axis and is subject to a restoring
force equal to -mn 2 x. It is also subject to a constant frictional resistance,
of magnitude mn 2 a, but which acts only when x is positive. The particle is
released from rest when x = -b(b > 0). Show that it next comes to rest
when x = (a 2 + b 2 )1 /2~ a and that it will then stay permanently at rest if
b < 3 1 / 2 a.
6.11 The displacement of a particle is given by
XIX3 - x~
Xo = ,
Xl +X3 - 2X2
112
164 Oscillations
approximately.
6.13 A spring balance consists of a pan of mass m supported by a spring of
stiffness u. Its motion is subject to a resistance equal to K times its speed.
A particle of mass mp is gently placed on the pan, which is initially at rest.
Show that when the pan comes to rest again for the first time, its
displacement from its initial position is
mpg
- - { 1 +exp ( -
u (4u(m
K1T
+ mp) - K 2)1/2
')I} '
e+ikf} + n 2 sin () = 0,
(ii) Show that if the motion is oscillatory and k ~ n, to the first order the
resistance affects only the amplitude and the buoyancy affects only
the frequency,
(When a body is immersed in a liquid it experiences a buoyancy force
which is equal and opposite to the weight of the liquid displaced by the
body,)
6.15 A convenient measure of the spread of the values of x(t) (0 < t < 00) about
°
the value x = is
in the sense that if the integral is small, I x(t) I must itself be small most of
the time.
Exercises 165
6.16 For time t";; 0 a particle hangs in equilibrium at the lower end of a light
elastic string, which is extended a distance a beyond its natural length. For
t > 0 the upper end of the string moves vertically, its downward displace-
ment being b sin pt at time t. Show that, provided the string remains taut,
the downward displacement x of the particle satisfies the equation
where n 2 =gla.
Show that the string remains taut throughout the motion if
pb
-2--2 (n sin nt - p sin pt) <a
n - p
pb < In - pia.
bp2
e = nl(n 2 - 2
p )
(n sin pt - P sin nt)
Show that the next time the angular speed vanishes the value of t is
2rr/(n + p).
6.18 Reciprocating or rotary engines are frequently mounted on their foun-
dations with rubber supports. Let these supports be modelled as a com-
bination of a restoring force proportional to the displacement and a
damping force proportional to the velocity. Because of its vibration, the
engine is subject to a periodic force with frequency equal to that of its
periodic motion. Thus the equation of motion, for vertical oscillations
about equilibrium, may be written
x == fo cos pt - (ki + n 2 x),
where fo cos pt is the periodic force per unit mass and (ki + n 2 x) is the
force per unit mass which arises from the supports and hence, by Newton's
third law, is the force transmitted to the foundation. Show that when the
transient oscillations are negligible, the ratio of the amplitudes of these
two forces is
Hence show that the amplitude of the force transmitted to the foun-
dations is always greater than that of the exciting force if p < 21/2 n.
Show that the ratio is less than ~ if either
x + n2 x == fo cos nt
fo t
- sin nt.
2n
Exercises 167
x + ki + n 2 x = fo cos nt
which satisfies the same boundary conditions. Show that the two solutions
agree in the limit k ~ O.
Sketch the solution when a = u = k = O.
(Ans.: kx =n -1 fo sin nt + e- k tl2 {ka coswt +w- 1 (ku +!k 2 a - fo)sin wt},
where w = (n 2 _ k 2/4)1/2.)
6.21 Consider forced oscillations. governed by the equation
x+kx+n2x=fo cospt,
when free oscillations are negligible. Verify from the solution that the
energy dissipated by the resisting force in one complete period is equal to
the work done by the forcing term as given by equation (6.3.10). Why
does the instantaneous rate of working of the forcing term not equal the
rate of dissipation of energy?
6.22 From the equation
This result also shows that the work done by the forcing term is, on
average, equal to the dissipated energy once the motion has become
periodic, since then the right-hand side is zero if t 2 - t 1 is the periodic
time.
6.23 An oscillator satisfies the equation
X =-
n 0
I
f. t
f(r)sin net - r)dr.
Let
t> to.
Show that the solution is then
x=
1 n
f020-cosnt),
2n
10
2
1
sin :Into sin n(t - ItO),
1
0";; t";; to,
t ~ to.
fo cos nt
x==-
X == fo t
x(t - r)V(r)dr,
Exercises 169
X+2wX+5w 2 X=0
Show that the work done by the force field in the time interval
o ~ t ~ Tis 2u 2 sin4 (wT/2) per unit mass.
6.27 Two particles A, B, of equal mass m, rest on a smooth horizontal table and
are connected by a light spring of natural length 1 and modulus A.. Initially
AB = I, A is at rest and B is given a velocity u in the direction AB.
Let x and y be the distances moved by A and B respectively, after time
t. Write down the equations of motion and show that
x + Y = 0, x - y = -n 2 (x - y),
where n 2 = 2A./ml.
Calculate x and y as functions of t.
(Ans.: 2x = ut + un -\ sin nt, 2y = ut - un -\ sin nt.)
6.28 A light spring, of modulus A. and natural length 21, has one end fixed to a
point 0 and has particles of masses m 1, m2 suspended respectively from
its mid-point and from its other end. Show that in equilibrium the
particles are at depths
respectively, below O.
Find the equations of motion for vertical oscillations of the particles
about the equilibrium positions and show that the frequencies of the
normal modes of vibration are given by the roots of the equation
170 Oscillations
6.29 Two simple pendulums AIBt and A 2 B 2 , each oflength I and mass m, are
suspended from two points At and A2 at the same level and at a distance a
apart. The bobs B t and B2 are connected by a light spring of modulus 'A.,
whose unstretched length is a. Show that there are particular oscillations
for which:
(i) the two pendulums are exactly in phase and oscillate with a periodic
time which is identical to that for either pendulum oscillating in-
dependently with the spring removed;
(ii) the two pendulums are exactly out of phase, so that their angular
displacements are 8 and -8 respectively, and such that the energy
equation can be written in the form
. 2g 2'A.
8 2 + - (1 - cos 8) + - sin 2 8 = w 2 ,
I ma
e
where w is the value of when 8 = O. Show that the frequency of
oscillation when 8 is small is (g/l + 2'A./ma)1/2.
(i) Show that if /J > 0 any motion is oscillatory, about the equilibrium
position x = a, for all values of the total energy E. Show that if the
mass oscillates with a small amplitude, such that ~ - a I ~ a, the
oscillations are approximately simple harmonic with frequency
1 2 { 1+
~x (ddxf )2} + gf(x) = constant.
Let f(x) = II x Icx/a cx ,
where I is positive, 0: > 1 and a is the value of I x I for which x = O. Show
that the bead can perform oscillations with periodic time
81 )1/2
T = ( -; F(o:, ajl)
(iii) F(a, (3) = (3G(a)+ 0«(3-1) for (3 ~ 1, where G(a) decreases monotoni-
cally from 2 to 1 as a increases from 1 to 00,
(iv) F(rt;(3) ~ 2 + (3 as a ~ 00,
(v) F(2, (J) = 2yE(y-l),
where y = (1 + (32 /4)1/2 and E is the complete elliptic integral of the
second kind, namely
to F(a, (3) has the properties (i)-(iv) with G(a) approximated by (a + l)/a.
Show that the error in G(2) is less than 5% and the error in F(2, 1) is less
than 4%. (E(2/5 112 ) = 1.18.)
6.34 The one-dimensional equation of motion of a particle is
h2
X = 3+ [(x),
x
a/(a)
---=n>3.
[(a)
Show that, if n < 3, the particle can perform small oscillations about
x = a with frequency h(3 - n)1/2 la 2 .
Sketch the potential energy curve when [(x) = -1)./x 2 (I). > 0). Hence
discuss the motion of the particle qualitatively. State the condition to be
satisfied by the total energy E for oscillatory solutions to be possible, and
calculate the periodic time of these oscillations.
(Ans.: -JP 12h 2 < E < 0, 21(1)./( _2E)3/2 .)
6.35 A jet of water issues vertically upwards from a nozzle. The mass of water
issued per unit time is km and the initial speed is u. The jet strikes a ball of
mass m and the vertical velocity of the water relative to the ball is
destroyed on impact.
Show that the momentum destroyed per unit time is
Hence show that the equation for vertical motion of the ball is
A force field with the property that the force at every point acts along the
position vector of that point relative to an origin 0 is called a central force field.
The motion of a particle subject to a central force is of particular interest
because of its wide application to problems as diverse as the motion of the
planets and the motion of electrons. The central force most commonly found in
nature is one which varies inversely as the square of the distance from the centre
of force O. We shall begin, however, by stating the equations for an arbitrary
central force.
Let the position vector of the particle relative to 0 be r, and let the force
acting on the particle be f(r)r/r per unit mass. The equation of motion is then
.. f(r)r
r=--. (7.1.1)
r
d
rxi =- (r x i) =0 (7.1.2)
dt
or r xi = hk (7.1.3 )
.. . 1 d -2'
r8 +2;0 =- -(,8)=0. (7.1.5)
r dt
"'{J=h, (7.1.6)
Example 7.1.1
We show that the radius vector to the particle from the centre of force sweeps
out equal areas in equal times.
This result follows from the fact that if in time 5t 8 increases to 8 + M , the
area of the elementary triangle swept out by the radius vector is!'" 58 correct
to the first order (see Fig. 7.1.1). Accordingly, if we denote this area by 5A, we
have that
and so the rate of sweep of area is constant and equal to !h. This result is a
direct consequence of the conservation of angular momentum.
Figure 7.1.1 Illustration for Example 7.1.1: elementary area swept out by
radius vector.
176 Central Forces [7.2
VCr) =- I r f(r)r . dr
r =-
I f(r) dr,
r
(7.2.1 )
where the lower limit of the integral can be chosen to be any convenient level of
zero potential energy and
r. dr = d(h . r) = d(!r2) = r dr
!v 2 + V(r)=E, (7.2.2)
where E is the total energy per unit mass. Since the speed of the particle in polar
coordinates is such that v2 =;-2 +r2(j2 we have, with the help of (7.1.6), (7.2.1)
and (7.2.2),
where
h2
2,
h2
U(r) = --::2 + VCr) = - 2
2r
-I r
fer) dr. (7.2 .4)
h2
;= ? + fer), (7.2.5)
dr _ o· -
-- dr_ h dr
--- (7.2.6)
dt de r2 de
7.21 Conservative Forces 177
-h
2r4
Z
(dr)2
-
dO
+ U(r)=E. (7.2.7)
h2
U(r) = 2r2 + V(r).
Note that the contribution h 2 /2r 2 to U(r) arises from an effective force h 2/r3 .
This is simply the term r8 2 , with 8 replaced by h/r2, transferred to the
right-hand side of equation (7.1.4) and treated as a contribution to the force
instead of to the acceleration, as in (7.2.5). When it is interpreted in this fashion
the term ,02 is called a centrifugal force. The motivation is that f can then be
regarded as the effective acceleration, in direct analogy with a one-dimensional
problem, and the arguments used to investigate the associated energy equation
(6.5.1) in the one-dimensional problem are also applicable to (7.2.3).
,2
The motion is restricted to values of r for which U(r) ~ E, since cannot be
negative. When U(r) = E, , (and dr/dO) is zero so that r has a stationary value and
the particle is moving perpendicular to the radius vector. A point on the
trajectory of the particle for which r has a stationary value is called an apse. The
direction of r there is called the apse line and r is the apsidal distance.
When U(r) is such that
where b < a and I/I(r) is strictly positive and single valued in the range b ~ r ~ a,
there is a solution in which r is confmed to the range b ~ r ~ a (see equation
(6.5.3)). There are just two apsidal distances, namely r = b, r = a, and r oscillates
between these two values. The trajectory of the particle is in fact deducible from
its behaviour in one of the intervals b ~ r ~ a. This is illustrated in Figure 7.2.1,
where B1 , A1 , B2 , Az , represent adjacent apses.
The situation can also be discussed using a more physical approach. Since the
force is always directed towards the origin, and is always the same at the same
distance, it follows that if at an apse the velocity of a particle were reversed, it
would retrace its previous path. It also follows that the path is symmetrical
about an apse line. Thus if adjacent apses on either side of OA 1 are B1 and B2 ,
we must have OB 1 = OB 2 . Similarly if A2 is the next apse beyond B2 , then
178 Central Forces [7.2
OAt =OA2 , and the trajectory can be built up by continual reflection about an
apse line.
The angle between two adjacent apse lines is called the apsidal angle and is
given by the expression
h2 JJ.
U(r)=- - -.
2r'2 r
This is the effective potential for the inverse square law, with f(r) =-JJ./r2 . For
the energy l~vel Eo and JJ. > 0 in Figure 7.2.2, the motion is circular with radius
ro. This i~ thl3 minimum energy level for which motion is possible. For E =E t
and JJ. >- 0, tho trajectory is bounded and r lies between r =rl and r =r;. In fact
we shall !lee later that the apsidal angle is 1T and that there are just two apses. For
=
E II" the motion is restricted to r ~ r2 (JJ. ~ 0) or r ~ r; (JJ. < 0). There is only
one apse and the trajectory is unbounded.
The inverse square law is appropriate for the determination of the motion of
the planets round the sun. The apses are then called the perihelion, where r is a
minimum, and the aphelion, where r is a maximum. The corresponding points
for orbits relative to the earth as the centre of force are called the perigee and
the apogee.
7.3] Circular Orbits and Stability 179
h 2 f.1
U(r)=- - -
2r2 r
U(r)
In the more general case, with U(r) given by (7.2.4), a circular orbit with
radius ro is possible if
(7.3.1 )
The result also follows directly from the original equations of motion. In the
first place equation (7.1.6) shows that
ti =h/r5 =w
say, where w is the constant angular speed with which the orbit is traversed.
Then (7.1.4), with f = 0, gives
(7.3.2)
in agreement with (7.3.1). It follows thatf(ro) must be negative so that the force
is one of attraction.
According to the general theory of §6.5, a perturbation of such a solution
will be oscillatory if U(ro) is a minimum and unstable otherwise. Thus to avoid
instability we must have
2
U /I (ro) = 3h ')
4""' -f(ro =w 2(3-n»0, (7.3.3)
ro
rof'(ro)
where n=- . (7.3 .4)
f(ro)
(7.3.5)
r
Hence, if at some particular time in the perturbed orbit = 0, it will also be zero
at a time n/{w(3 _ny/2} later. During this time the radius vector will travel
through an angle n/(3 - ny/2, since the angular speed is approximately w. This
is the apsidal angle for the perturbed orbit.
The critical case n = 3 requires a more accurate analysis, but will be oscil-
latory if U(r) has a minimum at ro.
For an inverse square law of attraction, with f(r) = -IJ./r2 ,we have n = 2 and
an apsidal angle of n. It is fortunate that this is stable, since it is the law of
attraction between the sun and the earth, whose orbit round the sun can be
regarded as a small perturbation of a circular orbit.
7.3) Circular Orbits and Stability 181
Example 7.3.1
A particle of mass ml moves on a smooth horizontal table. It is attached to a
light, inextensible string which passes through a small hole in the table and
carries a mass m2 hanging vertically.
Let r be the distance of m 1 from the hole, and let 8 be the angular
displacement of the horizontal part of the string. During the motion of the
system, there is no work done by the tension in the string (see Example 5.3.1).
The only force which does work on the system is the gravitational force acting
on m2, and this is conservative. The two conservation laws for the system as a
whole are
V 2 + U(r)=E, (7.3.6)
where (7.3.7)
/
/
/
/
/
/
/
/
/
/
U(r) /
/
/
/
/
/ "
/
/
/
/
/
/ ""
ro
r
Figure 7.3.1 Illustration for Example 7.3.1: variation of the function
[]I( r)
_1
- 2
ml h2
2
+ m2
gr.
m1 + m2 r ml + m2
182 Central Forces [7.3
Figure 7.3.1 shows the variation of U(r) with r. There is one stationary value,
when r = ro, given by
It follows that for the particle ml to move in a circular orbit of a given radius ro,
it must be given an angular momentum h which satisfies the relation
(7.3.8)
(7.3 .9)
Since
In fact it is clear from Figure 7.3.1 that all possible orbits are bounded and will
lie between the two circles whose radii are the two appropriate apsidal distances.
Suppose that ml is in a circular orbit and its velocity is increased in
magnitude by a factor k without change of direction. The effect of this on the
energy equation for the subsequent motion will be to change the angular
momentum from h to kh. When this change is made in (7.3.7) and the result
substituted in (7.3.6), the energy equation becomes
or (7.3.10)
Apart from the root r = ro, equation (7.3.1D) has one other positive solution
for r given by
(7.4.1)
where a and b are constant vectors. In terms of a and b, the energy and angular
momentum are, from (7.2.3) and (7.1.3),
(7.4.3)
r
h = 1r x 1= n 1a x b I. (7.4.4)
We proceed to show that the particle describes an elliptic orbit. Let rm be the
maximum value of r, and let the origin of time be chosen so that r = rm when
t = D. Define the cartesian axes Ox in the direction rm and Oy perpendicular to
rm in the initial direction of r. Because of the choice of axes these initial
conditions are that
(7.4.8)
184 Central Forces [7.4
dr =mg- z '
d2 1 T1
m (7.4.9)
(7.4.10)
(7.4.11)
I
I
I
I
z+
I
I
I
I
IL ___ • __ _
r
mg
Figure 7.4.1 Illustration for Example 7.4.1: notation for the spherical
pendulum.
7.51 The Inverse Square Law 185
d2 r gr
dt 2 = - I
correct to first order. This represents elliptic harmonic motion with period
21T(I/g)1 /2 .
The theory of motion under a central force was originally developed by Newton
as a consequence of his law of gravitation, which states that the force on a
particle P1, of mass m1 and position vector r1, arising from the attraction
between P1 and a second particle P2 , of mass m2 and position vector r2, is
Gmlm2(r2 -r1)
1 r2 - r1 13
where
and is called the gravitational constant. There is an equal and opposite force
acting on P2 •
This is the source of the earth's gravitational attraction, but has a much wider
application including, for example, the motion of the planetary system. Since G
is small in magnitude, the force is appreciable only when large masses, such as
the stars and planets, or small distances are involved. Although the law is stated
here for the attraction between particles, it can be shown that spherically
symmetrical distributions of mass attract in the same way as particles of similar
mass at their centres. This is a good approximation for the planets.
One of Newton's earliest calculations in this connection was concerned with
the attractive force of the earth for the moon. Let re be the radius of the earth
and let d be the distance of the centre of the moon from the centre of the earth.
According to the inverse square law, the force of attraction at the moon is
g(re/d)2 per unit mass, where g is the force per unit mass at the surface of the
earth. From equation (7.3.2), with f(d) = -g(re/d)2, it can be seen that the
moon will describe a circle of radius d around the earth if
where w is the angular speed of the moon in its orbit and is equal to 21T /T, where
T is the perio die time. Hence
186 Central Forces [7.5
The quantity sin -1 (rjd) is called the moon's horizontal parallax and is known
from observation, as indeed are all the other quantities in the above relation.
This calculation provided one of the first tests of the law of gravitation.
Subsequent work in astronomy and in the dynamics of satellites has abundantly
confirmed the accuracy of the law in these fields.
The inverse square law also governs the attraction between electrically and
magnetically charged particles, and can be either an attraction or a repulsion
depending on the sign of the charges. One of the important applications here is
to the motion of an electron in an electromagnetic field, for example in its orbit
around the nucleus of an atom. Strictly speaking, the latter problem requires the
theory of quantum mechanics for its complete formulation, although Bohr's
original calculation of the energy levels of the hydrogen atom was based
substantially on the classical theory.
The polar equation of the orbit for a general law of force may be obtained
from the integral of equation (7.2.7). It so happens that there is an elementary
integral for the inverse square law. We put f(r) = -Jljr2 , so that Jl is positive for
an attractive force, and for the moment we restrict the analysis to Jl> O. It is
convenient to take the zero level for the potential energy at infmity and define
Equation (7.2.7) then becomes, with the help of (7.2.4) and (7.5.1),
(7.5.2)
du 1 dr
-=-
dO r2 dO
or, equivalently,
(7.5.3)
Equation (7.5.3) can now be compared with the energy equation (6.1.2) for
simple harmonic motion. In fact the transformations
7.5] The Inverse Square Law 187
(7.5 .4)
change (7.5.3) into (6.1.2) with n=1 and amplitude (1 +2h2E/J1?)1/2. The
solution of (7.5.3) is therefore deducible from (6.1.4) as
h2 u h 2
-=-=1+ ( 1
J.l f.1.T
+7
2h2E)1I2
cos (0 + a). (7.5.5)
2h2E)1I2
e= ( 1 + - - <1 (7.5.8)
112
and the polar axis is chosen so that dr/dO is a minimum there, hence a = O.
An alternative form of the equation can be obtained using cartesian co-
ordinates x = r cos e ,y = r sin 0 , so that (7.5.6) gives
When the terms in this equation are appropriately rearranged, the result is
(x + ae)2 y2
2 +2=1, (7.5.9)
a b
188 Central Forces [7.5
c x
where (7.5.10)
_ I _ 2 1/2 _ h
b - (1 __ e2 )1/2 - a(1 - e) - (_2EY 12 • (7.5.11)
Equation (7.5.9) is the cartesian equation for an ellipse with centre at x = -ae,
y = 0, semi-major axis a, semi-minor axis b, eccentricity e and semi-latus rectum I
(see Fig. 7.5.1).
For an ellipse, the energy equation in its original form
(7.5 .12)
simplifies to (7.5.13)
with the substitution for VCr) and E from relations (7.5.1) and (7.5.10) respec-
tively. This is a useful relation giving the speed in terms of the distance from the
centre. It shows that the speed is always less than (2IJ./r)1/2 in an elliptic orbit.
For the circular orbit r = a the result
(7.5.14)
(7.5.15)
For a given central force, the periodic time in an elliptic orbit varies as the 3/2
power of the major axis. This is a result which has been checked with consider-
able accuracy for the planetary system.
At the surface of the earth P./T~ = g, where Te is the radius of the earth. It
follows that for a particle to travel in a circular orbit just above the earth's
surface requires a speed slightly less than
this gives
which represents a parabola with latus rectum I and vertex at the point x = 1/2,
Y =0 (see Fig. 7.5.2).
Equation (7.5.12) simplifies to
v2 = 2/1/r. (7.5.18)
In this orbit a particle will eventually recede to infinity, but with its kinetic
energy and potential energy both tending to zero. Thus for a given initial value
of r, (2/1/r)I/2 represents the minimum initial value of the speed required if the
particle is to escape to infinity. This is just 21/2 times the speed required for a
circular orbit at the same radial distance.
From the surface of the earth the escape speed is
llr= 1 +e cosO
represents a hyperbola. We also note that when /.1. < 0, so that the force is one of
repulsion, we must have E> 0 because the kinetic energy !mv2 cannot be
negative and the potential energy V = -/.1./r is now positive. It is convenient to
treat the cases E > 0, /.1. ~ 0 together.
For p. < 0 we have an inverse square law of repulsion. We write the energy
equation (7.5.2) in the form
(7.5.20)
l/r = -1 + e cos 0
with a suitable choice of polar axis. The orbits are therefore given by
with the positive sign for p. > 0 and the negative sign for /.1. < O.
Relations (7.5 .21) represent the two branches of the same hyperbola. In
terms of cartesian coordinates x = r cos 0, y = r sin 0, (7.5.21) becomes
(7.5.22)
I I p. I
where a=--=- (7.5.23)
2 e - 1 2E'
_ I _ 2 112 _ h
b - (e 2 _ 1)112 - aCe - 1) - (2E)1/2 . (7.5.24)
192 Central Forces [7.5
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/
/'
Figure 7.5.3 Hyperbolic orbits for Jl> 0 and Jl < 0: 0 is the centre of force.
(7.5.25)
It is clear from Figure 7.5.3 that a particle corning in from infinity along one
asymptote and receding to infinity along the other is ultimately deflected
through an angle 2 tan -1 alb. In terms of the speed at infinity, which is
v~ = I Jl lIa (7.5.26)
(7.5 .27)
tered through different angles X depending on the distance b of their initial line
of approach from the scattering centre (see Fig. 7.5.4); b is called the impact
parameter. The number dN of particles scattered between angles X and X + dX
will be proportional to the flux n of particles in the incident uniform beam, that
is the number of particles crossing unit area normal to the beam in unit time.
Hence we write dN = ndo, where do has the dimensions of area and is called the
effective scattering cross-section. It is determined by the nature of the scattering
process and is an important characteristic of that process. We assume that the
scattering process is symmetrical about an axis through the scattering centre
parallel to the incoming beam, as is the case for a fixed central force field. We
also assume that the particles which are scattered through angles between X and
X + dX are those which originated with impact parameters between b and b + db
(see Fig. 7.5.4). Since the flux of particles across a ring of thickness 1 db 1 normal
to the incident beam is n2rrb 1db I, it follows that
do = 2rrbl db I,
where the modulus sign is used to ensure that do is positive even when db is
negative, as it is in Figure 7.5.4. In order to obtain the dependence of 0 on X, we
regard dX as a positive increment in X and write
This is the desired relation in general terms, and is valid for any law of scatter.
The precise form of (7.5.28) depends on the details of the scattering process,
which determine the variation of b with X. For the inverse square law (both
194 Central Forces [7.5
III I
b =--coth
v~
The result shows that da has a rriarked variation with the scattering angle X,
and this can be checked experimentally from an observation of the flux
dN = nda. In his investigation of atomic structure, Rutherford bombarded atoms
with a-particles. On the assumption that the force between the positive charge of
an a-particle and the positive charge inside the atom varies according to an
inverse square law of repulsion, the experimental observations should agree with
(7.5.29) for the effective scattering cross-section. We expect the assumption to
be valid provided that the charges can be treated as point charges. But if positive
charge were distributed throughout the volume of the atom, a deviation' from
(7.5.29) would be expected for those a-particles which penetrated the atom.
Rutherford obtained agreement between theory and experiment provided that
the distance of closest approach was larger than 10 -14 m and so concluded that
the positive charge was concentrated within such a radius. This was the origin of
the nuclear theory of the atom.
The above calculation leading to (7.5 .29) assumes a fIxed centre of force and
so is applicable only when the nucleus of the atom is much heavier than the
~-particle. The modification required when the masses are comparable will be
discussed in Example 8.3.4. The a-particles will also be deflected by the negative
electrons, but the mass of the electron is relatively very small and the deflection
is consequently insignifIcant (see Example 8.3.4).
Finally it should be noted that the interaction of an a-particle with an atomic
nucleus is strictly a problem of quantum mechanics, where the concept of a
definite particle with a defmite trajectory is not appropriate. On the other hand,
the concept of a scattering cross-section for a scattered beam of particles is
appropriate in quantum mechanics, and (7.5 .29) does in fact remain valid.
Summary
For reference purposes we end this section with a summary of the formulae
involved in the study of orbits governed by the inverse square law. The number-
ing of the equations agrees with that used in the body of the chapter.
The eccentricity e and the semi-latus rectum 1are
2h2E)112
e= ( 1+-2- , (7.5.8)
Il
(7.5.20)
7.5] The Inverse Square Law 195
with the positive sign for J.1 > 0 and the negative sign for J.1 < O.
Elliptic Orbit
(7.5 .9)
(7.5.10)
(7.5.11 )
Parabolic Orbit
(7.5 .l7)
(7.5.18)
Hyperbolic Orbit
(7.5.22)
I I J.1 I
a=--=- (7.5.23)
e 2 - 1 2E'
_ I _ 2 1/2 _ h
b-(e2_1)1/2-a(e -1) -(2E)1I2' (7.5.24)
Example 7.5.1
In the following example the earth is regarded as a sphere of radius re and the
196 Central Forces [7.5
rotation of the earth about its axis is neglected. The resisting effect of the
atmosphere is also neglected.
(i) A particle is projected from the surface of the earth with speed Ve at an
angle (j to the vertical. Show that the maximum range is 2rea m , where
sin am =KI(2 -K) and K = v~/gre < 1. Show that the maximum range is
attained when tan {j = (1 - K) -1/2.
(ii) If at its maximum height the speed is increased so as to put the particle into
circular orbit, what is the additional speed required?
(i) The trajectory of the particle is an ellipse. The centre of the earth is the
centre of attraction O. The maximum height is achieved at the apse which is
furthest from 0 (the apogee, see Fig. 7.5 .5).
It is convenient to take the polar axis along the radius vector from the
centre of the earth to the apogee. The equation of the trajectory is then
I/r= 1 - e cos e.
Since the orbit is symmetrical about the apogee, the range is 2re a, where () = a
when r = re (see Fig. 7.5.5). The maximum range is 2re a m , where am is the
maximum value of a as a function of {j. The equation of the trajectory gives
J.l./,~ = g, the gravitational acceleration, and this serves to determine J.l. = gr~.
The angular momentum h and the total energy E are determined by the
initial conditions, which give
1 - K sin 2 {3
cos a = {l _ K(2 _ K)sin2 (3}/2 '
whence
K tan (3 K ~
tan a = ------'----,-
1 + {1 - K)tan 2 (3 (l_K)1/21+e'
where ~ = (1 - K)1/2 tan (3. Now the maximum value of ~/(1 + ~2) is !,
attained when ~ = 1. It follows that
K . K
tan am = 2{1-K)112 or sma =--
m 2-K
(li) Now let us consider the problem of putting the particle into circular orbit
'a
upon reaching the apogee. The radial coordinate there is given by
'a =l/{1-e)
198 Central Forces [7.5
or 2_2
Va - Ve - 2gre
( I-e)
I - K sin2 (3 •
2 J1 gre(1 - e)
=-=
'a
Vo
K sin2 {3
Since e has already been found to be {I - K(2 - K)sin 2 {3} 1 /2 , both Va and
Vo can be determined in terms of the initial conditions. The difference
Vo - Va is the additional speed required to boost the particle into circular
orbit.
Example 7.5.2
Two satellites, SI and S2, are describing circular orbits in the same plane, of
radii '1and '2 (> ,d, around a fixed centre of force O. Let us consider the
manoeuvre required to effect a rendezvous by bo~sting the velocity of SI . The
most effective way to do this is to increase the velocity of SI without change of
direction so as to put SI into an elliptic orbit whose apsidal distances are'l and
'2 (see Fig. 7.5 .6).
The orbit required will have the equation
1/, = I + e cos (J
where 1/'1 = I + e, 1/'2 = I - e
or '2 -'1
e=--
'2 +'1
This defmes the orbit. The speed in the elliptic orbit at (J = 0 is given by
Since 82 has angular speed (p./,D 1 / 2 , in this time it will have travelled an angular
distance
At the beginning of the seventeenth century, and sixty years before Newton
postulated his theory of gravitation, Kepler had enunciated the following three
laws, which he deduced empirically from the extensive series of observations of
Tycho Brahe:
(i) The radius vector drawn from the sun to a planet describes equal areas in
equal times.
200 Central Forces [7.6
(li) Each planet describes an ellipse with the sun at one focus.
(iii) The squares of the periodic times of the planets are proportional to the
cubes of the major axes of their paths.
From these three laws it may be deduced that the law of attraction between
the sun and the planets is a central force of attraction inversely proportional to
the square of the distance. For the first law implies, by (7.1.7), that r2 is e
constant. By (7.1.5) this implies that the force is a central force, since the
transverse acceleration is zero.
The second law states that the path of the planet has a polar equation of the
form
llr = 1 + e cos O.
At this stage we have deduced that for each planet the force varies inversely
as the square of the distance, but so far we have only shown that 11 is a constant
for a particular planet. The third law now shows that 11 is the same constant for
all planets since, by (7.5 .15),
and, by the third law, rl/T2 is the same for all planets.
"In point of fact, accurate observation indicates a small discrepancy in
Kepler's third law. This has a simple theoretical explanation in that gravity is
both a mutual and a universal force, and the model of a planet subjected to a
single force arising from the sun as a fixed centre is oversimplified. Consider the
mutual force between the sun, of mass m s , and a planet, of mass mp. The law of
gravitation implies that the planet is subjected to an acceleration Gms/r2.
However, since the force between sun and planet is a mutual one, the sun is
subjected to an acceleration Gmp/yl . Thus the acceleration of the planet relative
to the sun is G(ms + mp)jr2. The constant of proportionality, 11, is therefore
effectively G(ms + mp) instead of Gm s, and is hence in error by a factor
(1 + mp/ms) when calculated on the assumption of a fixed centre of force. One
7_61 Kepler's Laws 201
21ra3/2
T=------
/-L 1I2 (1 + mp/ms)1I2
and the constant of proportionality is now different for different planets_ The
ratio of mp/ms is, however, very small as can be seen by reference to Table 7 _6_1,
which summarises some of the more important data concerning the planets_ Its
largest value, for Jupiter, is 10- 3 _ For the earth it is 3 X 10-6 _
Of the information summarised in Table 7 _6_1, the periods of revolution and
periods of rotation are obtained by telescopic and radar observation_ The length
of the orbital axis is then deduced from Kepler's third law, which states that
T~ /T~ = a~ /a~ _
The eccentricity of the earth's orbit can be deduced from the fact that the
distance to the sun is inversely proportional to the sun's apparent diameter_
Sidereal
Sidereal period Semi-
period ofrevolu- major
Mean of axial tion about axis of Eccen-
Mass radius rotation the sun orbit tricity
Notes
1. The earth is used as the unit, but note that the sidereal period of axial rotation is relative
to the fixed stars. For the earth this is 365/366 times the apparent period relative to the
sun (1 day). The difference arises from the motion of the earth around the sun_
2. The last two rows give, for comparison, data for the earth and moon in metric units_
Data for the moon refer to its orbit about the earth_
202 Central Forces [7.7
From the geometry of the ellipse (Fig. 7.5.1) the maximum and rmmmum
distances to the sun are in the ratio (1 + e)/(I - e). This is equated to the ratio
of the maximum and minimum apparent diameters of the sun. The same
principle holds for the other planets, but the calculations are more elaborate.
If a planet has a satellite, the mass of the planet relative to that of the sun can
be calculated from observations of the period of revolution of the planet round
the sun, and the satellite round the planet. For example, if the suffix m refers to
the orbit of the moon round the earth (of mass me) and the suffix e refers to the
orbit of the earth round the sun (of mass 11Is) we have, from (7.5.15) with
11= Gm,
and so
If am/a e and TmiTe are known from observation, this relation will serve to
determine the ratio me/ms'
To determine the mass of a planet in terrestrial units requires a know-
ledge of G, and this in turn is only determinable from an experiment in-
volving known masses. The most recent of such experiments gives the value
6'67 x 10- 11 m3 kg- 1 S-2.
In the discussion so far the effect of resistance on orbits has been neglected. This
is quite legitimate in the case of a planet moving round the sun. However air
resistance will have a modifying effect near the surface of the earth, for example.
Compared with the radius of the earth, the surrounding atmosphere is only a
thin layer. At an altitude of 100 km above sea level the density of the atmos-
phere is reduced by a factor of 10-8 compared with its value at sea level.
Weather satellites, which are in circular orbits higher than this, are moving in a
vacuum for all practical purposes and resistance is negligible.
As a first example of the effect of resistance, we consider a satellite whose
elliptic orbit round the earth lies outside the earth's atmosphere except for a
small portion near the point of closest approach (the perigee) as in Figure 7.7.1.
Since the resistive effect acts over a relatively short interval, we idealise the
resistive force into an impulse, applied at the perigee along the line of flight. This
will not affect the position of the perigee, since the impulse will not alter the
radial distance rp and the velocity of the satellite will still be perpendicular to
the radius vector after the impUlse. But the speed vp at the perigee will be
changed by a small amount SVp. There will be a corresponding change Sa in a,
the semi-major axis ofthe orbit. Thus, by (7.5.13),
7.7] The Effect of Resistance 203
., ., .... -
I
/ '"
I
,
f
, allTlO!>pherc
\
\
,
" "- ,
' ....
--
Figure 7.7.1 Satellite in orbit round the earth.
(7.7.1 )
(v + OV )2 = JJ. (~ _ _
P P TPa + oa
1_) (7.7.2)
after the impulse. Subtraction of (7.7.1) from (7.7 .2) and neglect of second and
higher powers of oVp and oa gives
(7.7.3)
Since oVp is negative, so is oa and the semi-major axis of the orbit is decreased.
The periodic time, since it is proportional to a 3 / 2 , is also decreased. Also, forthe
elliptic orbit,
Hence there is a small change oe in the eccentricity given, to the first order, by
the relation
oa oe
-=-- (7.7.5)
a l-e
Thus e also decreases with a and vp , and each time the satellite passes through
the earth's atmosphere the effect, according to the above model, is to leave the
minimum distance Tp unchanged, but reduce the maximum distance a(1 + e)
from the origin. The orbit will therefore tend to become circular. This model,
204 Central Forces [7.7
however, will have ceased to be valid before a circular orbit is achieved since the
assumption that most of the orbit lies outside the earth's atmosphere will
eventually fail to hold.
We therefore consider a suitable model to take account of a persistent
resistance to the motion of a satellite. We shall use an empirical law of the type
discussed in Chapter 4. Of the two more usual assumptions, that of a linear
variation with velocity, rather than a quadratic law, is more appropriate for a
highly rarefied atmosphere. Since this is also much the easier to handle we shall
proceed on this basis.
Because the resistance is not a conservative force, we must retum to the
equations of motion, which will now be augmented by a force -kf per unit mass,
where k is a positive constant, to give
.• 0'2 J1 •
r-r =--;J.-kr, (7.7.6)
1 d. .
- - (rO) = -krO. (7.7.7)
r dt
(7.7.8)
where h is a positive constant. This relation shows that the angular momentum is
no longer constant but is subject to an exponential decay because of the
resistance. This decay will be slow if k is small but persistent and cumulative.
Equation (7.7.6) is not so readily integrated and to proceed further requires a
preliminary transformation, the reason for which is not immediately apparent. It
should be judged by the simplicity achieved in the final results. We first note the
following relation:
this relation is substituted in (7.7.6) and eis eliminated with the help of (7.7.8).
The result is
(7.7.10)
(7.7.11)
where 1 is a constant with the dimensions of a length and serves to give the order
7.7] The Effect of Resistance 205
(7.7.12)
e-3kt~ (e-3ktdP)
dt dt
=e-3ktdO.!
dt dO
{_!:.!(!)}
dO
[2 P
h2 d (1)P .
2
= - [4 p2 do2
(7.7.14)
(7.7.15)
This equation would be of the simple form already discussed in § 6.1 if it were
not for the awkward term on the right-hand side. We now consider the condi-
tions under which it would be appropriate to neglect this term, and this requires
an estimate of its magnitude. For this purpose it is convenient to write h, which
from (7.7.8) is the angular momentum apart from the decay factor, as f w. Here
[, as already introduced, is a typical measure of the radius vector in the orbit.
Then w is a typical measure of the angular speed of the satellite in the orbit. The
factor k 2 [4/h 2 which appears on the right-hand side of (7.7.15) becomes, with
the suggested substitution, k 2 /w 2 • This is a non-dimensional quantity since both
w -I and k- I have the dimensions of time. But whereas w -I is a typical measure
of the time associated with the motion of the satellite in its orbit, k- I is an
.appropriate time scale with which to measure the effects of resistive decay. For
example, equation (7.7.8) shows that significant changes in the angular momen-
tum, through the factor e -kt , occur when the time changes by an amount of the
order of k -I . In a typical practical situation the ratio k/w will be very small, and
neglect of the right-hand side of equation (7.7.15) amounts to the realistic
assumption that e /w 2 will be negligibly small. With this simplification the
solution is immediate and can be written
1 JJl
- = - + € cos(O + ex), (7.7.16)
P h2 '
where € and ex are constants of integration. By analogy with (7.5 .7) we now
206 Central Forces [7.7
choose I to be h 2/p., since its precise definition is still at our disposal. Also IX = 0
if the polar axis is chosen appropriately so that fmally we have
iJ =!!:
12
(1 + € cos 8)2 e3kt
'
(7.7.18)
or, on integration,
3k
e3kt = 1 + - /(8), (7.7.19)
w
where
o d8
/(8) = fo (1 + e cos 8)2 ' (7.7.20)
3k )2/3
Ilr = (1 + e cos 8) ( 1 + -;:; /(8) . (7.7.21)
The integral /(8) is expressible in closed form* though at the cost of the
desirable simplicity which so far has given qualitative insight merely by inspec-
tion of the results. If we assume that the orbit is nearly circular, again a property
often met in applications, then the analysis simplifies, for € ~ 1 and / = 8 + O(e).
Since klw is also assumed small, the error incurred when / is replaced by 8 in
(7.7.21) is a second order term and hence negligible on a linear theory. The
*f8
, dO = 2 tan-I
{(1--- € ) 112 0 }
tan- -
€ sin 0
iff <L
o(l+€COSO)2 (l_e 2 )3/2 l+€, 2 (l-€2)(l+€cosO)
Exercises 207
I
-; = (1
( 3k
+ e cos 6) 1 + w 6
)2/3 (7.7.22)
Further simplification on the grounds that k/w is small is not possible. The
required condition is strictly k6 / w ~ 1, and this depends on the magnitude of 6
as well as that of k/w.
EXERCISES
7.7 Two uniform solid metal spheres, each of diameter a and density p, are
placed at rest with their centres a distance 2a apart. Show that if they are
subject only to their own mutual attraction, they will come into contact in
a time (l + rr/2)(3/Grrp)1I2 , where G is the gravitational constant.
Show that for iron spheres, of density 7'9 x 10 3 kg m -3, the time is
about an hour.
It may be assumed that the gravitational force of attraction between
the spheres is equivalent to that between particles of the same mass
situated at their centres.
7.8 A particle describes a circle of radius a under the influence of a central
force varying inversely as the square of the distance. Show that if the
direction of motion is suddenly changed, without change of speed, the
new orbit is an ellipse with semi-major axis a and the same periodic time as
the circular orbit.
7.9 A particle describes an ellipse of semi-major axis a under the action of a
force JJ./r2 per unit mass towards a fixed focus. When the particle is at one
end of the minor axis it collides and coalesces with a second particle
which is initially at rest. Prove that the new orbit is an ellipS"e and
that, if the periodic times are Tl and T2 before and after the impact,
1 < T 1 /T2 < 2 312 •
7.10 A particle describes an ellipse under the central force -JJ.r/r 2 per unit mass.
The minimum and maximum speeds are u, U, respectively.
Show that the eccentricity of the orbit and the speed of the particle are
given by e = (U - u)/(U + u), v2 = 2JJ./r - uU.
7.11 Two satellites are travelling in the same direction in the same circular orbit
of radius ro and initially are some distance apart. One of them, in an
attempt to reach the other, increases its speed without change of direction
by an amount u. Given that k = u(ro /JJ.)1/2, where JJ./r2 is the force of
attraction per unit mass at a distance r, and that e· + 2k < 1, show that
the new orbit is an ellipse whose major axis is ro (1 - 2k - k 2 ) and whose
eccentricity is (2k + k 2 ).
Discuss the possibility of effecting a rendezvous by this means.
7.12 A particle moves in a bounded orbit under an inverse square law of force.
Prove that the time average of the kinetic energy is half the magnitude
of the time average of the potential energy.
7.13 A comet travels in a parabolic orbit round the sun. Its least distance from
the sun is kre(k < 1), where re is the radius ofthe earth's orbit. Show that
the time during which the comet's distance from the sun is less than re is
Show that the time is a maximum when k = t, and that it is then 2/3rr
years.
7.14 A satellite moves in a circular orbit of radius 4rm round the moon, which
is assumed to be a uniform sphere of radius rm' The gravitational force at
the surface of the moon is g m per unit mass. The satellite is to be
intercepted by a rocket fired horizontally from the moon's surface with
initial speed (2gr m )1/2 and launched from a point in the plane of
the satellite's orbit. Show that the rocket should be fired at a time
Exercises 209
IS·SS(r m /gm )1/2. after the satellite has passed vertically over the rocket
site.
h~unching
7.15 (i) A particle is subject to a central force -p.r/r 3 per unit mass, where r is
the position vector. It is projected at right angles to the radius vector
with speed U at a distance c from the centre of force. Show that
(a) For u 2 <,. J.l./c describe the motion of the particle and give the
periodic time.
(b) For p./c < u 2 < 2p./c, find the greatest and least distances from
the centre of force attained by the particle in its orbit. Deduce the
periodic time.
(c) For u 2 ;;;;. 2p./c what is the asymptotic value of the speed far from
the centre?
(i) The particle is projected at right angles to the radius vector from a
point at a distance C from the centre of force with speed nvo, where
Vo is the speed which would be required for a circular orbit. Show that
the path of the particle turns through an angle cP in the subsequent
. motion, where sin cP = (n 2 - 1)-1, provided n 2 > 2.
(ii) The particle is projected from infinity with a velocity of magnitude
nvo, where Vo is the speed required for a circular orbit of radius c, and
the velocity is initially directed along a line whose perpendicular
distance from the centre of force is c. Show that the path of the
particle is turned through an angle 2CP, where tan cP = n -2 .
7.17 A uniform spherical star, of mass ms and radius rs , is moving with constant
velocity u through a large cloud of particles. Show that all particles ahead
of the star and within a cylinder of radius
where G is the gravitational constant, will eventually collide with the star.
7.lS A particle is subject to an inverse square law central force of attraction. It
210 Central Forces
is projected from infinity with speed Vee along a line whose perpendicular
distance from the centre of force is b, and is ultimately deflected through
an angle rr/2.
Let O(x, y, z) be a system of cartesian axes such that the particle starts
from infinity along the positive x-axis and ultimately recedes to infinity
along the positive y-axis.
Show that the centre of force is at the point (b, b, 0).
A particle is projected from infinity with speed v"" along a line, parallel
to the x-axis, whose perpendicular distance from the centre of force is c.
Show that the orbit has asymptotes which intersect on the plane x = 0 and
that the particle is ultimately deflected through an angle 2 tan -1 clb.
Particles are projected from infinity with speed v"" along lines parallel
to the x-axis within the cylinder (y - b)2 + Z2 = c2 . Show that, if c ~ b,
all such particles ultimately recede to infinity within a circular cone of
semi-angle 2clb, whose axis is the line y =b, Z =0 and whose vertex is the
point (b /2, b, 0).
7.19 (i) Prove that
~(~)=~
dt r r3
(rxdr)xr.
dt
Show that if the central force is of the form _p.r/r 3 per unit mass,
d
p. dt
(r); = dt2r
d2
x h.
1
u=-
r
7.21 A particle, which moves under the influence of a central force, describes
the curve r(O + 1)2 =a, where a is a constant and (r,O) are its polar
coordinates. What is the law of force?
When the particle is in the position 0 = 0 it receives an impulse which
reduces its radial velocity to zero and doubles its transverse velocity. Show
that its subsequent path is given by
3a/2r= 1 +1 COS(3 112 0/2).
7.22 A particle is subject to a central force with potential energy -tJ./4r4 per
unit mass. The angular momentum is h per unit mass.
(i) Show that if the total energy is zero, the orbit consists of part of a
pair of touching circles, each of radius (tJ./8h 2 )1/2 .
(ii) Show th1t if the total energy is h4/4tJ., the orbit is part of the curves
-m(!!..? -!!.)
r4 r,
where tJ. and v are positive constants. It is projected with speed Vo from a
point at a distance a from the centre of force, at right angles to the radius
vector. Find the equation of the orbit and show that if vg < 2tJ./a - vla 2
then the orbit is bounded. What is the additional condition for the orbit to
be closed?
Prove that when the particle is next moving at right angles to the radius
vector, the latter has turned through an angle 1ravo/(v +a 2vg)1/2.
7.24 A particle of mass m is subject to the central force
the differential equation for the trajectory can be written in the form
u=- ,
r
(see Exercise 7.20). On the assumption that the term on the right-hand
side of the first equation is small compared with the individual terms on
the left-hand side, show that a solution of the form
is consistent with the equation provided that all terms of the second order
of smallness are neglected. Substitute the assumed form for u into the
equation and compare like terms to obtain the following relations:
fJ.l v
1 - 2" = -2- (1 + ~e2),
h h l
I - n2 = 2v/(h 2 l),
€( 1 - 4n 2 ) = ve 2 /(2h 2 l).
When fJ.V/h4 ~ 1, show that these relations can be solved to give, approx-
imately,
h2 fJ.V 1 2
I = - - 4 (l + 2 e ),
fJ. h
n =1 - fJ.V/h4 ,
For e < 1, show that the most significant terms describe an ellipse
whose axes rotate slowly with the angular speed wfJ.v/h 4 , where 21f/w is
the time to traverse the orbit.
For e ~ 1, show that the resulting nearly circular orbit agrees with that
of Exercise 7.24.
Note, however, that the result is not now confined to nearly circular
orbits. A similar modification of the inverse square law arises from the
theory of general relativity, with v = 3h 2 fJ./C2 , where c is the speed of light.
It is used to explain the slow rotation of the orbit of the planet Mercury.
7.26 A particle moves under the influence of a central force f(r)r/r per unit
mass, where r is the position vector of the particle.
When the orbit has two apsidal distances a, b, show that the apsidal
speeds are in the ratio b /a.
Show also that the speed v at a distance r from the origin is given by
~(b2 - a 2 )v2 = a 2 fr
a
f(r) dr + b 2 f r
b
f(r) dr.
Exercises 213
7.27 It has been suggested that the gravitational constant may in fact vary with
the time. Consider the equation
d2 r r
dt 2 =- p.;:s
in which p. varies with the time. Show that angular momentum is conserved
but that there is no equation of conservation of energy.
Suppose that the particle traverses a bounded orbit, and that the unit
of time is adjusted so that the period of revolution is of order unity.
Suppose also that p. is a slowly varying function on this time scale, so that
p. = p.(et), 0 < e ~ 1.
s =R(et)r, T = fo
t
T(et) dt,
provided that
T1I2 =R = 11,
(8.1.1 )
is called the mass centre of the system of particles (see Fig. 8.1.1). If c' is also
a point such that
then, by subtraction,
~mi(CPi -
I
C'P;) = ~miCC' = (~mi)CC' = O.
I I
o
Figure 8.1.1 Illustration of the notation for systems of particles.
216 Systems of Particles [8.2
Lmj(CO + OP j ) = 0,
j
or (8.1.2)
(8.2.1)
(8.2.2)
where me = ~m and r e , a e are the position vector and acceleration of the mass
centre. In (8.2.2) the suffix i has been omitted for economy of notation.
In the summation over the forces acting on the system we can distinguish
between two contributions. The first contribution consists of all those forces
originating from outside the system. Particles moving in the neighbourhood of
the earth, for example, would experience the earth's gravitational attraction and
a resistance arising from their motion through the air. Such forces are external to
the system of particles and are to be distinguished from the forces arising from
mutual interactions between the particles. The important dynamical distinction
is that each internal force has an equal and opposite counterpart within the
system, by Newton's law of action and reaction, and the internal forces will
therefore cancel each other in a summation over the whole system of particles.
Thus in equation (8.2.2) the summation on the left-hand side can be restricted
to the external forces. We then have
(8.2.3)
where Fe refers to the typical external force on the typical particle. Equation
(8.2.3) governs the motion of the mass centre, whose acceleration is determined
by the resultant external force. If the latter is zero, then the velocity of the mass
centre is constant.
8.2] Equations of Motion 217
where Ve is the velocity of the mass centre and [mevel denotes the change in
mcvc over the interval of integration. When the interval of time tends to zero,
the limit of the integral
(8.2.6)
However, the internal forces may do work during the motion of the system, and
this must be included in the summation on the left-hand side of (8.2.6). For
example the work done by the forces acting on two particles connected by a
spring is equal to ! x tension x extension.
Let mi' mk be the masses of two particles with position vectors rj and rk
respectively. Let there be mutual forces ±Fjk respectively on the two particles
arising from their interaction. The work done by these forces is
(8.2.7)
and so depends on the relative motion of mj and mk. Note, however, that if Fjk
is parallel to rk - rj, and if Irk -- rj I is constant so that the particles move as a
rigid body, no work is done by Fjk because it is perpendicular to the direction of
relative motion. In this important special case the work done is confined to that
of the external forces, and if they are conservative it can be calculated from the
potential energy in the usual way.
There is an alternative expression for the kinetic energy ~!mv2, which is
derived as follows. We denote by r' and v' = dr'jdt the position vector and
velocity of a typical particle relative to the mass centre C, so that r = re + r' and
by differentiation, v = Vc + Vi (see Fig. 8.1.1). We note from (8.1.1) that
218 Systems of Particles [8.2
(8.2.8)
since ~m =me and ~mv' =O. The first term of (8.2.8) is equal to the kinetic
energy of the whole mass moving with the velocity of the mass centre. The
second term is the kinetic energy of the motion relative to the mass centre.
Example 8.2.1
In this example we have a small ring, of mass m1 , which is connected by a string,
of length I, to a bob of mass m2' The ring is free to slide along a smooth, fixed,
horizontal rail, as in Figure 8.2.1. We shall investigate the oscillations of this
system in a vertical plane.
Let x be the displacement of the ring from some suitable origin, and let e be
the inclination of the string to the vertical. The ring then has a horizontal
velocity x. The bob also has the horizontal velocity x of the ring; in addition it
has a velocity relative to the ring equal to 18 perpendicular to the string; see
Figure 8.2.1.
Of the forces acting on the system consisting of the ring and the bob, the
tension in the string is an internal force acting mutually on the two masses. It
does not contribute to the resultant force acting on the system. Moreover,
provided the string remains taut and inextensible, the tension is always in a
direction perpendicular to the relative displacement of the masses. It will
therefore do no work during the motion. The external forces are the gravi-
o x
i
tational forces and the normal reaction N between the ring and the wire. These
are all vertical. Also N and mig act in a direction perpendicular to the displace-
ment of the ring and so do no work during the motion. The gravitational force
m2g is conservative and so is derivable from a potential.
Since there is no resultant horizontal force, it follows from (8.2.4) that the
horizontal component of the momentum is constant. We suppose that the
system is started from rest; the horizontal component of momentum is then zero
and we have
(8.2.9)
Since the only force which does work on the system is m2g, and since this
force has the potential energy -m2g1 cos 8, there is an energy equation of the
form
This equation governs simple harmonic oscillations with period 2rr/n, where
n 2 =(ml +m2)g/ml 1.
The word collision is often used to denote the actual physical impact of two
bodies, the effect of which is to produce impulsive changes in velocity. These
changes take place in a very short time interval, outside which the velocities of
the bodies are constant if no other forces are involved. In this discussion it is
convenient to widen the definition and describe a collision between two particles
as an interaction between them such that the initial and final velocities are
constant. A collision can then take place without actual contact of the particles;
an example we shall eventually discuss is the motion of two particles which are a
large distance apart initially and which attract or repel each other with a force
inversely proportional to the square of the distance apart.
220 Systems of Particles [8.3
Example 8.3.1
We consider the spontaneous disintegration of a particle into two parts.
Let u be the velocity of the particle before disintegration. Let ml ,m 2 , Vt , V2
be the masses and velocities of the two parts after disintegration (see Fig. 8.3.1).
From conservation of momentum we have
vi
e_ u
Since Vc = (miVI +m2V2)/(ml +m2) is the velocity of the mass centre after
diSintegration, conservation of momentum shows that Vc remains unchanged and
equal to u. Also, since
where
are the velocities relative to the mass centre, it follows that the momenta of the
two parts relative to the mass centre are equal in magnitude and have opposite
directions.
The total energy in this problem can be calculated from the sum of the
kinetic energy arising from the motion of the mass centre, the kinetic energy of
the motion relative to the mass centre, and the internal energy which we denote
by E i • Conservation of energy then gives
which serves to determine v~, and hence v;, in terms of the internal energy
released. The angles through which the particles are scattered, say () I and () 2, are
not deducible without further information. If v~ >u, any value is possible for
() I, but if v~ < u, there is a maximum value given by sin -1 (v~ /u).
Example 8.3.2
We now investigate the collision of two particles.
Let the masses of the particles be ml ,m2' Let the velocities be UI , U2 before
collision and VI, V2 after collision. From conservation of momentum, we have
(8.3.1)
(8.3.2)
(8.3.3)
222 Systems of Particles [8.3
---'---'----.u c
/
/
/
---"T-----.uc
If we assume that the collision is perfectly elastic then kinetic energy is also
conserved. Effectively this means that the kinetic energy of the motion relative
to the mass centre is conserved, since conservation of momentum ensures that
the energy associated with the motion of the mass centre is unchanged. We
therefore have the further relation
(8.3.4)
(8.3.5)
Hence the momenta of two particles relative to the mass centre are unchanged in
magnitude by the collision, the effect of which is to rotate both the relative
momentum vectors through an angle of scatter X, say.
Laboratory experiments are often such that particle 2 is initially at rest, and
then (8.3.1), (8.3.2) and (8.3.5) give, with Uz =0,
(8.3.6)
given in terms of X, the angle of scatter in the centre of mass frame, by the
relations
V'l sin X sin X
tan () 1 = , = ---..:..:....-- (8.3.7)
Uc + VI cos X (mdm2) + cos X
and
(8.3.8)
since v;
= U c . If ml < m2, any value for 8 1 is possible, but if ml > m2 there is a
maximum value given by sin -1 (m2/md.
The kinetic energy transferred to the second particle is
We compare this with the total kinetic energy !mlui. The ratio is
so that the maximum transfer of energy occurs for a head-on collision (()2 = 0).
Note that it is small for ml·~ m2 and for ml ~ m2' It can only be substantial if
ml and m2 are comparable. For a head-on collision of a proton and a-particle,
where mt/m2 = 4, the fraction of energy transferred is 64%. For a proton-elec-
tron collision it is about 0.2% (mt/m2 = 1836).
When bodies of macroscopic size collide impulsively there is invariably a loss
of kinetic energy. A head-on collision is often dealt with by using the empirical
rule
(8.3.9)
where e is called the coefficient of restitution and differs for different materials.
This relation used in conjunction with the conservation of momentum is suf-
ficient to give VI and V2 in terms of Ut and U2' By (8.3.2) the kinetic energy
relative to the mass centre is
Example 8.3.3
When dissipative processes, such as friction, are present, conservation of energy
can be invoked only if the energy converted into heat is taken into account.
However, on a molecular level, this heat energy can be identified with the kinetic
energy of the random motion of the molecules. We investigate this on the basis
of an idealised model of a simple gas consisting of molecules all having the same
mass m. The gas is assumed to be in equilibrium, which means that on average its
energy does not change. It is also assumed to be isotropic, which means that
there is no preferential direction of motion of the molecules. Let us consider the
force on the walls of the container enclosing the gas. Suppose that when a
molecule strikes one of the walls it is reflected elastically, so that its normal
component of velocity is reversed in sign and its tangential component is
unaltered. If the wall is parallel to the (y, z) plane, a molecule moving towards
the wall with velocity ui + vj + wk before impact will then have a velocity
-ui + vj + wk away from the wall after impact. The net change in momentum is
2mu normal to the wall. Let nu be the number of molecules per unit volume
having an x component of velocity u. The number of such molecules hitting the
wall per unit area per unit time is then un u • The rate of change of momentum
per unit area is 2mu 2 n u . This is now summed over all positive u to get
2m ~ u 2 n u = mnu 2
u>o
say, where n is the number of molecules per unit volume, and u 2 is the mean
value of u 2 over all values of u. Note that the molecules for which u is positive
are assumed to be moving towards the wall. Since there will be an equal number
of molecules for which u is negative, we must have
2 ~ u2 n u = ~ u 2 n u = nu 2 •
u>O aIlu
where c is the speed of a molecule. We now have the result that the rate of
8.3] Collision Problems 225
change of momentum per unit area of the containing wall, which is the force per
unit area, or the pressure, is given by
p=pRT,
where T is the absolute temperature and R is the gas constant. We now see that
the concept of temperature is closely associated with the kinetic energy of the
random motion of the molecules. In fact the kinetic energy per unit volume is
!mnC2 =~pRT.
The above result might seem at first sight to depend on the idealised
assumption that the molecules are all reflected elastically from the wall. Such
elastic reflection does not take place in practice, indeed for some molecules
there is no correlation between the velocity of approach and the velocity with
which they leave the wall. Nevertheless if the isotropy of the gas is to be
conserved, the average effect at the wall is the same as if such reflection had
taken place. The reflecting wall is the simplest mechanism which will maintain
the isotropy of the gas, but the mathematical result depends on this isotropy and
not the particular means by which it is achieved.
Example 8.3.4
A particle PI, of mass ml, is projected from a great distance with a velocity u
along a line whose perpendicular distance from a stationary particle P2 , of mass
m2, is d. We consider the motion of PI and P2 when subjected to a mutual
attracting force inversely proportional to the square of the distance (see Fig.
8.3.3).
The initial velocity of the mass centre Cis ml u/(ml + m2) and, since there is
no external force, C continues to move with this constant velocity. We therefore
consider the motion of PI relative to C as a convenient origin. The motion of P2
can be dealt with by a similar argument.
Let the force on PI towards C be
kmlm2 _ kmlm~
(P 1 P2 Y- (ml + m2)2r2'
where PI C =r. The value of k will depend on the particular mechanism from
which the force arises. A measurable effect in the laboratory is obtainable if, for
example, the particles are electrically charged. Particle PI is therefore subject to
226 Systems of Particles [8.3
~!~
m +m
1
4
2
/
/
/"
/
/
"/
/
,/
,/
,/
_ _ _ _ _ _ f.l __ L.~~ - -
U
an inverse square law central force towards C of magnitude /1/r2 per unit mass,
where
The relative path is a hyperbola and the details are deducible from § 7 .5. In
particular, the angle through which the path turns is, by equation (7.5.27),
The result is not immediately useful for comparison with observation, be-
cause in practice the deflection angle will be measured relative to axes fixed in
the laboratory. To calculate this angle we note that the ultimate velocity of PI
consists of two components, the velocity of C and the velocity relative to C. The
velocity of C is
The direction of this component makes an angle X'with the initial direction. The
resultant of these two components makes an angle 8 with the initial direction,
where
Or
sin X
tan 8 = = tan h
l+cosX
and so e = x/2.
Note that the scattering cross-section for this problem is still given by relation
(7.5.29) relative to the centre of mass frame. But to obtain a result which is
useful in the laboratory frame, the variation of (7.5.29) with e and de, rather
than X and dX, needs to be calculated using the above relation between 8 and X.
Example 8.3.5
A wedge, of mass m2, rests with one face on a horizontal plane and with the
inclined face at an angle a to the horizontal. It is hit on its inclined face by a
particle, of mass m1, moving along the plane with speed u in a direction
perpendicular to the edge of the wedge. Show that the speed of the wedge
immediately after impact is mtU sin 2 a/(m2 + ml sin 2 a).
The particle then slides up the face of the wedge and back to the horizontal
plane. Find the final velocities of the particle and the wedge.
It is to be assumed that there are no frictional forces and that the impacts
occurring are inelastic, in other words relative motion in the direction of the
impact is destroyed.
(i) Since the impact between the particle and the wedge is inelastic, there will
be no rebound and the particle will proceed to slide up the face of the
wedge with relative initial speed Ul, say. Let the initial speed of the
wedge be U2' The impact between the particle and wedge is perpendicular
to the wedge face. Hence the momentum of the particle parallel to the
wedge face is unchanged. This yields the relation
228 Systems of Particles [8.3
UI
ml'U
~
m~ex
/11/111 II I,fl II;,
Figure 8.3.4 Illustration for Example 8.3.5: collision of particle and wedge.
The impulsive reaction between the wedge and the plane is vertical. Hence
the horizontal momentum of the system consisting of the wedge and the
particle is unchanged. This yields the relation
(8.3.11 )
Equations (8.3.10) and (8.3.11) can be solved for U1 and U2' The result is
(8.3.12)
(li) In the next stage of the motion the particle slides smoothly up the face of
the wedge and returns to its original position. There are two conservation
equations which are sufficient to determine the final speeds VI of the
particle relative to the wedge and V2 of the wedge (see Fig. 8.3 .4). These
equations arise from the conservation of horizontal momentum, since all
external forces acting on the system are vertical, and from the conservation
of energy. Since there is no net work done on the system the latter
equation amounts to conservation of kinetic energy, and we have
for the downward acceleration ofthe particle relative to the wedge, and
where v is the speed of the particle after travelling a distance x up the plane.
In particular x = 0 when v = VI' Hence VI = UI and the particle returns to
the base of the wedge with the same relative speed, but with a reversal of
direction. From (2.3.7) the time taken is 2ut/at, and from (2.3.6) the
wedge in this time acquires a speed
(iii) Finally the particle leaves the wedge and returns to the horizontal plane.
This involves another inelastic collision which destroys the vertical com-
ponent of velocity of the particle. The final speed of the particle is therefore
V2 - VI cos ex = V2 - U I cos ex
u(mi sin 2 ex+mlm2 - m~ cos 2 ex)
(ml + m2)(ml sin2 ex + m2)
F=ma (8.4.l)
was introduced for a single particle, it was assumed that the mass of the particle
was constant. When the mass varies, the equation is modified, and the modifi-
cation depends on the details of the process by which the variation is effected.
For example, a raindrop may increase its mass by condensation as it falls
through a cloud, and a rocket decrease its mass by the ejection of burnt gases.
When such problems are examined from the point of view of the appropriate
system, it will be seen that the mass of the system as a whole is not changing,
230 Systems of Particles [8.4
(~-m----")
v
------- ( m Hm ) ~
v+bv v
e ·~u
@
. ---"'u
t +ot t + bt
(a) (b)
Figure 8.4.1 The changes which occur in a small interval of time when: (a) the
mass increases; (b) the mass decreases.
but that individual parts of the system are changing their momentum. For
example, the condensed water vapour is moving with the cloud before conden-
sation, and with the raindrop after condensation, which implies a change in
momentum. Similarly the fuel of the rocket moves with the rocket before
ejection and with the velocity it acquires from combustion afterwards. Such
problems can be treated as dynamical systems in which the impulse of the
external forces is equated to the changing momentum of the system.
let us consider a mass m which, during the interval of time ot, changes its
velocity from v to v + ov and combines with a mass om. Let the velocity of om
before amalgamation with m be u (see Fig. 8.4.1 (a)). If Fe is the external force
acting on the system, it follows from (8.2.4) that
J
t+lit
Fedt=(m+om)(v+ov)-omu-mv (8.4.2)
t
or, equivalently,
1 jt+ot ov om
- F dt=(m+om)-+(v-u)-. (8.4.3)
Of t e Ot ot
The left-hand side of equation (8.4.3) represents the average value of Fe over the
interval of time 0 t, and will tend to the value of Feat time t as 0 t -+ O. Thus in
the limit ot -+ 0 equation (8.4.3) becomes
dv dm
Fe =m dt +(v-u)dt' (8.4.4)
Figure 8.4.1 (b) represents the situation when mass is ejected, rather than
acquired, as in rocket propUlsion. Because om conventionally denotes an incre-
ment in m, the ejected mass is denoted by -om. In problems where m is actually
decreasing, om and dm/dt are negative. The same equation (8.4.2) also governs
this situation and leads to the same limiting equation (8.4.4). Two particular
8.4 ] Variable Mass Problems 231
d
F =-fmv). (8.4.5)
e dt"
This is the equation governing the fall of a raindrop through a cloud which is at
rest.
The second case is u = v, when
Fe = mdv/dt (8.4.6)
and the equation of motion (8.4.1) for a particle of fixed mass is recovered. The
motion of an object from which particles are being released would be governed
by such an equation. But if the particles are ejected with a finite relative
velocity, the more general equation (8.4.4) must be used.
Consider, for example, the rectilinear motion of a rocket in free flight, that is
with no external forces. We assume that the exhaust velocity relative to the
rocket is constant and equal to c. The velocity of the rocket is v = vk say, and
u - v = c = -ck since the gases are ejected in the direction opposite to v.
Equation (8.4.4) then becomes, with Fe = 0,
dv dm
O=m-+c-. (8.4.7)
dt dt
where R (= mo/md is the ratio of the initial to the final mass. Note that the
change in speed depends only on the exhaust speed c and the mass ratio R. It
does not depend on the rate at which the fuel is consumed.
Example 8.4.1
In a two stage rocket, the first stage is detached after exhausting its fuel in order
to minimise the mass before the second stage is fired. Such a rocket is to
accelerate a 100 kg payload to a speed of 6000 m S-I in free flight. The exhaust
speed is 1500 m S-I relative to the rocket. For structural reasons, the mass of
either stage, without fuel or payload, must be 10% of the fuel it carries. We
consider the best choice of masses for the two stages so that the initial mass is a
minimum
Let the masses of fuel carried in the first and second stages be x and y kg
respectively as shown in Figure 8.4.2, so that the initial mass is
mo=100+1·1(x+y)
232 Systems of Particles [8.4
0'1 x + x
and the mass after the first stage has exhausted its fuel is
ml = 100 + l'ly +0·1 x.
VI = 1500 log(mo/md.
One immediate deduction is that the desired speed of 6000 m S-I cannot be
achieved by a single stage rocket. For with y = 0, mo/ml cannot be larger than
11, and 1500 log 11 is only 3597.
If we proceed to the second stage, after first jettisoning the burnt-out first
stage, the initial and final masses are, respectively,
11mom2
------=e4
(mo + 1Om2)m3
Exercises 233
(11
e ) =m
m ---
o m3
4
m2
(11
~ 100+0·ly
- e4
100+ 1·ly
) =10e 4
.
_10_0_+_I_·I,;:...y = e2
or y = lOOO(eZ - 1) = 1769.
100 + O·ly 11 - eZ
Example 8.4.2
A raindrop falls vertically from rest at time t = 0 through an atmosphere
containing water vapour at rest. The mass of the raindrop increases by con·
densation, so that the mass is mo + At at time t, where mo and Aare constants.
The motion of the raindrop is resisted by a force equal to kv per unit mass, where
v is the speed and k is constant.
Since the cloud is at rest, the equation of motion of the raindrop is
d
- {(mo + At) v} = (mo + At)g - k(mo + At)V
dt
or
which integrates to
and so
EXERCiSES
that the greatest inclination of the string to the vertical after it is free from
the pulley is
f= fo1f/2
(l + 2 sin 2 8)112 d8,
where E{k} is the complete elliptic integral of the second kind, namely
E{k} = fo 1f/2
{I - k 2 sin 2 8}1/2 d8.
8.3 A smooth semi-circular thin wire of radius a has two rings fixed at its ends
A and B. The total mass of the rings and the wire is mt. The rings can slide
freely on a. fixed, smooth, horizontal rod. A bead of mass m can move
freely on the wire and starts from rest at the end A of the wire. The
motion takes place in a vertical plane, and 8 is the angle which the radius
through the bead makes with the horizontal radius through A.
State, with reasons, whether or not the following quantities are con-
served:
. 2g(mt + m)sin (J
(J2 = - - " - - - - -
a(mt + m cos 2 (J) .
Hence show that the time for one complete oscillation of the system is
I(a/g)1/2, where
1= Jo
1r(2(m
t
+ m cos 2 (J))1I2
(mt + m)sin (J
d(J.
8.4 Two particles have masses ml, m2 and position vectors rl, r2 respectively.
They influence each other by gravitational attraction and there are no
other forces.
Show that if re is the position vector of the mass centre and G is the
constant of gravitation,
(i) re is constant,
d2 (rl- r2)
(ii) -d2 (r1 - r2) = -11 1 13 '
t r1 - r2
(iii) (r1 - r2) x (il - f2) is constant,
d2 Gm~ (rl - Ie)
(iv) -d = - (ml + m2 )2 1
2 (II - Ie)
t II - re 13 '
8.7 Two particles, each of mass m, are connected by a light elastic string of
236 Systems o[ Particles
(dtdx)2 = n (x -
2 1) 7 (x + 1) - (x -
(exI2 1)
) ,
Show that if ex ;p I the length of the string varies between the limits I
and l(ex l /2 - 1) approximately.
8.8 A smooth block has mass mb and its upper and lower faces are horizontal
planes. It is free to move in a groove in a horizontal plane and a particle of
mass m is attached to a point on its upper face by an elastic string of
natural length I and modulus X. The string is stretched parallel to the
groove to a length (ex + 1) times its natural length and the system is
released from rest. The particle always remains in contact with the upper
surface of the block.
Show that the motion is periodic with the periodic time
( "A(mlmmb)
+ mb)
1I2( 21T + -;;4) .
8.9 Three particles, each of mass m, are free to move on a smooth horizontal
plane. The mutual force of attraction between each pair of particles is of
magnitude m 2 J1 times their distance apart. At time t = 0 the particles are at
points P, Q, R and have velocities "AQR, XRP, APQ respectively. Show that
their centre of mass remains at rest, and that each particle describes an
ellipse with periodic time T = 21T/(3IJ.m) I /2.
Show that the area of the ellipse is 2M T /3, where A is the area of the
triangle PQR.
8.10 Two gravitating particles, of masses ml and m2, start from rest an infinite
distance apart and are allowed to fall freely towards one another. Show
that, when their distance apart is a, their relative velocity of approach is of
magnitude {2G(ml + m2 )/a}1I2 , where G is the gravitational constant.
Impulses of magnitude F are applied to the particles in opposite
directions perpendicular to their line of motion at the instant when their
distance apart is a. Show how to determine b, their distance of closest
Exercises 237
approach during the subsequent motion, and prove that when b ~a, it is
given approximately by the relation
b Fla ml +m2
a 2G mim~ .
8.11 A train consists of an engine, of mass me, and n trucks, each of mass m.
The train is initially at rest and each coupling has a length I of slack. Show
that if the engine starts to move forward and exerts a constant pull F, the
speed Uj of the train just after the jth truck has been jerked (inelastically)
into motion satisfies the difference equation
8.12 A shell of mass 2m is fired vertically upwards with velocity v from a point
on a level stretch of ground. When it reaches the top of its trajectory it is
split into two equal fragments by an internal explosion which supplies
kinetic energy equal to mu l to the system. Show that the greatest possible
distance between the points where the two fragments hit the ground is
2uv/g if U ~ v or (u 2 + v2 )/g if u ;;;. v.
8.13 A shell of mass m explodes when in flight into n fragments, whose masses
are ml, m2, ... , m n . Show that the energy developed by the explosion is
such that the kinetic energy associated with the mass centre is unchanged.
Show that the kinetic energy relative to the mass centre is
I 2
- ~ ~ mimj Vi,',
4m i j
of radius a, and the particles are scattered without loss of energy. Show
that the effective scattering cross-section is
1 1 1
- - - =- (1 - cos 8).
k2 kl cmo
8.19 The mass m of a moving particle is a function of the speed v, and the force
required to accelerate the particle is equal to the rate of change of
momentum. Show that the energy equation can be written
I I Iv
- mv 2 + - 2 -dm dv + V = constant,
2 2 dv
m = mo ( 1 - V2c )-1/2
2
V2)-lf2
( 1 -;?
moc 2 + V = constant.
Suppose that the particle moves along a straight line with V =! J,l m ox2 ,
where x is the distance of the particle from a centre of attraction O. Show
that the motion is oscillatory with the periodic time
Exercises 239
8.20 (i) Material is fed continuously from a hopper at rest on to a moving belt
at a mass rate m. Find the force required to keep the belt moving at a
constant speed v. Show that the power required is twice the rate of
change of kinetic energy. Why is the extra power required?
It may be assumed that the material sticks to the belt without
sliding.
(ii) An endless belt moves on rollers at each end. The axes of the rollers
are horizontal and parallel. The upper part of the belt is plane and
inclined at an angle (J to the horizontal. It is fed from a hopper, at a
constant rate 1m, with material which sticks to the belt until it reaches
the lower end, where it falls off. The distance travelled by the material
before it falls off is I. There is a constant force R resisting the motion.
Show that the steady state speed of the roller is
{R2 + 4m 2 gl sin (J}1I2 - R
v=
8.21 A uniform layer of snow, whose surface is a rectangle with two sides
horizontal, rests on a mountain slope of uniform inclination Q, the ad-
hesion being just sufficient to hold the snow while at rest. At a certain
instant, the uppermost line of snow starts to move downwards and collects
with it the snow it meets on the way. Kinetic friction between the snow
and the slope is negligible.
Prove that if v is the speed when a distance x of the slope has been
uncovered, then
d 2 2
- (x v ) = 2gx2sm
·
Q.
dx
Hence show that the moving snow has a constant acceleration
of!g sin Q.
8.22 A boat, of total mass m, is propelled at constant speed u through water
which offers a resistance equal to k times the speed.
At a given instant the boat springs a leak and subsequently takes in
water at a constant mass rate A. The boat sinks when the mass of water in
it is m. Show that the boat travels a distance
-mu (k
- -+-A0- r k / A) )
k+A A k
dv
(m - At)- =F-R.
dt
240 Systems of Particles
Given that the engine does work at the constant rate P, that there is a
resistance Uv, and that v = Vo when t = 0, show that
8.24 A cloud of water vapour moves with constant velocity u. A spherical drop
in the cloud increases its mass, by condensation, at a rate k 1 S, where S is
its surface area. It is subject to a constant gravitational acceleration g and a
resistance equal to k 2 S times the velocity of the cloud relative to the drop,
where kl and k2 are positive constants. Given that the drop is initially
very small, show that its trajectory is similar to that of a particle subject to
a uniform gravitational acceleration equal to kl g/( 4kl + 3k 2 ).
8.25 A rocket, of initial total mass mo, is fired from rest vertically upwards. It
propels itself by ejecting mass at a constant rate k with constant speed c
relative to itself. When all the fuel has burnt, the mass of the rocket is mi'
Show that if kclmog> I, the speed of the rocket at the instant of burning
out is c log(mo/md approximately.
A two-stage rocket is constructed from two rockets with the same
exhaust speed c. They each have a mass ratio mOlm1 equal to 4, but the
initial total mass of the first is twice that of the second. The second is
mounted on top of the first in such a way that it begins to fire just when
the first is burnt out and at this instant it is automatically released from
the first. Show that the speed attained by the second rocket is the same as
that of a single-stage rocket of mass ratio 8.
8.26 A rocket is fired vertically, starting from rest, and is propelled upwards by
projecting propellant at a constant relative speed c downwards. The
propellant burns for a time T and the ratio of the initial mass of the rocket
to the mass when the propellant is all burnt is R.
Write down the equation of motion and deduce that the rocket will not
rise from the ground initially unless
2
c log2
---=-- - R cT(R logR - )
1 .
2g R-I·
Show that the minimum value of u for which the rocket will escape
from the earth's gravitational field is (2a)1/2{gl/2 _ (kc)1/2}.
8.28 A ram-jet aeroplane, of constant mass m, moves horizontally with speed
vet). It takes in air, which is at rest, and ejects an equivalent mass of air at
the back with constant speed c relative to itself. The mass rate at which air
is taken in and ejected is mav per unit time, where a is a constant. There is
a resistance to the motion equal to mkrl' .
Calculate the change in momentum during a small increment in time,
and hence show that
dv
dt + av(v - c) + kv n =o.
For the case n =2, show that the speed v after the aeroplane has
travelled a distance x is given by
v = w - (w - vo)e -(:Ix,
where ~=a+k, vo«w) is the speed when x=O and w is the terminal
speed.
8.29 A pail, of large cross-section and mass m p , is made to move vertically
upwards with constant speed u. Water falls into it from a tank, a mass m
leaving the tank per unit time. The initial speed of the water is zero and
the first of the water reaches the pail when it is a distance h below the
tank. Show that if, after a further time t, the pail is at a depth x below the
tank, it then contains water of mass
Determine the rate at which mass is entering the pail and hence show
that the force required to maintain the speed of the pail is
9.1 EXISTENCE
The purpose of this chapter is to introduce a vector associated with rigid body
displacements, and to discuss its properties and applications.
Definition
A system of vectors ri(i = 1, 2, ... , n), which are displaced subject to the
constraints
(9.1.1 )
ri . ri = constant. (9.1.2)
for all i,j, including i = j. Geometrically the vectors are all of constant magnitude
and the angle between any two of them is constant. For future reference we
note that the vectors ri x ri can be added to the system in the sense that the
augmented system will still satisfy (9.1.3). This is so because the magnitude of a
typical vector ri x rj will depend on the constant magnitudes of ri, ri and the
angle between them. Furthermore the constant angles between three vectors ri,
rj, rk of the system are sufficient to determine the angle between rk and ri x rj,
which is therefore constant (see Example 1.5.4).
The vectors ri need not be interpreted as position vectors, but if they are it
will be seen that they are displaced according to the intuitive notion of a rigid
body, for which the distance between any two points remains invariant. The
effect of the additional constraint (9.1.2) is to restrict the displacement to a
rotation about an origin 0, which is fixed.
Let each ri be displaced to ri + ori in a way compatible with (9.1.3). Let the
displacement ori be of the form
(9.1.4)
9.1] Existence 243
fj • v; + f; • Vj =0 (9.1.5)
and in particular
fi • Vi = O. (9.1.6)
Theofem
If the rigid body displacements are of the form (9.1.4), and if the fi are not all
parallel, there exists a unique vector w such that
(9.1.7)
(w - w') x fi = 0 (9.1.8)
for all vectors of the system. Since the vectors fj are not all parallel, (9.1.8) can
only be satisfied if w - w' is a null vector, so that w' = w.
Let ft, f2, f3 be three linearly independent vectors of the system. Such a
choice is always possible since, by assumption, we can choose ft and f2 to be
non-parallel. If an appropriate third vector does not already exist we can
augment the system with the vector f3 = f t x f2 which satisfies all the required
conditions. The associated vectors Vt , V2 and V3 cannot all be parallel since, by
(9.1.6), this would imply that ft, f2 and f3 are all perpendicular to Vt and
therefore are not linearly independent. Without loss of generality we assume that
Vt and V2 are not parallel.
If there is a vector w satisfying (9.1.7) for all i it must be perpendicular to all
the Vj' In particular it must be perpendicular to Vt and V2. and therefore of the
244 Angular Vectors [9.1
where relations (9.1.5) and (9.1.6) have been used. It follows that 00 x fl = Vt
and 00 x f2 = Vz provided that C = -(fl' V2)-1 = (r2 . vd- l or
VI X Vz VI X V2
00=----=---. (9.1.9)
rl.V2 f2'VI
(9.1.10)
Since we have shown that 00 x fj = Vj for j = 1,2, it follows that (9.1.10) is zero
for j = 1, 2 and for all i. It is also zero for i =j since both 00 x fi and Vi are
perpendicular to fi' We can therefore deduce that (00 x r3 - V3) . fj = 0 for j =
1, 2, 3. Now a vector whose scalar products with three linearly independent
vectors are zero must be a null vector (see Example 1.5.5). Since fl , f2' f3 are
linearly independent we must have oox f3 = V3' We have now shown that oox fj =
Vj for j = 1, 2, 3. It follows from (9.1.10) that (00 x fi - va . fj = 0 for j = 1,2,3
and for all i. Hence 00 x fj - Vi = 0 and the theorem is proved.
The situation when the fi are all parallel is that a linearly independent
reference triad of vectors, on which the above argument depends, cannot be
defined within the system. However, we can augment the system with vectors
which do have the desired property, for example if f2 and f3 are added such that
they form a mutually orthogonal triad with a vector f1 of the system, then an
angular vector 00 can be defined for the extended system and hence for the
original system. Because of the flexibility in the choice of the added vectors
there is not a unique vector w associated with the original system and it is clear
that, for any scalar "11., (00 + Ar1) x fi =Vi if 00 x fi =Vj for any vector fj parallel to
f1 . The component of 00 parallel to the system is therefore arbitrary and can be
chosen, for example, to be zero by appropriate choice of the displacements V2
and V3 associated with the added vectors f2 and f3'
To interpret 00 geometrically, let fj refer to the position vector of a point Pj,
and consider a representation of 00 passing through the origin. Then 00 x fi St
represents a displacement perpendicular to 00 and fj with magnitude equal to
w 8t multiplied by the length of the perpendicular PiN from Pi on to 00 (see
Fig. 9.1.1). The rigid body displacement is thus eqUivalent to an angular dis-
placement of magnitude W ot about ON, at least to first order. For this reason an
alternative presentation of a rigid body displacement about a fixed origin 0 is by
the introduction of a vector oe n whose magnitude is the angular displacement
Of! , and whose direction is that of ON, the axis of rotation. The rigid body
9.1] Existence 245
(9.1.11)
for all points of the rigid body. However, the relation is correct only to
first order. It is not possible to defme vectors for which (9.1.11) is true
exactly when 081n1 and 08202 are small but non-zero displacements. In
spite of this we shall see that the vector W plays a significant role, especially
when we consider the continuous motion of rigid bodies.
So far the analysis applies only to a restricted rigid body displacement for
which both conditions (9.1.1) and (9.1.2) are satisfied. When (9.1.2) is relaxed,
we choose a reference vector, say r 1 , from the system. Then the relative vector
system r; = ri - rl will satisfy the stronger conditions, including (9.1.2), and an
angular velocity WI can be defined such that
and so (9.1.12)
When ri refers to the position vector of a point Pi, the more general
displacement ori of Pi consists of the displacement orl of PI, which is a
translation of all the Pi as a rigid body, together with the displacement or;
relative to PI, which is a rotation WI cSt to first order. In fact WI is independent
246 Angular Vectors [9.1
since rz and 8r2 satisfy (9.1.12). Subtraction of (9.1.13) from (9.1.12) gives
(9.1.14)
(9.1.15)
for all vectors ri of the system. Provided the vectors ri are not all parallel,
(9.1.13) can hold only if WI = W2 = W say. Thus w is a property of the
displacement independent of the reference vector.
Example 9.1.1
In this example we consider a vector defined by a finite angular displacement
and its implications.
Let a rigid body be capable of rotation about a fixed point O. In an angular
displacement (J I about an axis in the direction of the unit vector nl, let the
point P of the body, with position vector r, move to the point PI, with position
N~---+--'f------~
o
Figure 9.1.2 Illustration for Example 9.1.1: finite rotations.
9.1 ] Existence 247
(9.1.16)
and serves to determine rl in terms of rand n,., at least implicitly. To see that
relation (9.1.17) is satisfied, we first note that
where M is the mid point of PP1 (see Fig. 9.1.2). Also nt x OM has the direction
of PP t = rl - r and the magnitude of MN, where MN is the perpendicular to M
from the axis. Finally
and so the two sides of (9.1.17) have the same magnitude and the same
direction.
Note that, when 0 t is small so that I rl - r I is also small, we have
(9.1.18)
(9.1.19)
r2 - r = ! n 12 x (r + r2 ). (9.1.20)
The relationship between n 1 ,.$lz and n t2 is not the simple law of addition for
vectors, that is to say
(9.1.21)
and that
(9.1.24)
*
Thus $1 1Z £12 1 unless Al x £12 = 0, so that even the order in which the
displacements occur affects the result. Again, however, we see that the simple
vector relation
does hold approximately for small angular displacements, provided that the
second order terms are neglected.
Let r;(t) (i =1, 2, ... ,n) be a system of differentiable vector functions of t. Let
Then Vj, as defined in (9.1.4), is just the derivative of rj(t) since, by definition,
dri . lirj .
- = hm - = hm (Vj + Ej) = vi. (9.2.2)
dt or->'O li t or->' 0
dri drl
- =- + W x (rj - rl ) (9.2.3)
dt dt
for an appropriate vector 00. To give (9.2.3) a physical interpretation, let t be the
time and let the rj be the position vectors of points Pj moving as a rigid body.
Then (9.2.3) gives the velocity of Pi as the vector sum of the velocity of a
reference particle PI and the velocity relative to PI. The latter is determined by
a rate of rotation vector 00, now called the angular velocity. When drl /dt = 0 and
00 is a constant vector, Pi describes a circle with 00 as axis and with constant
angular speed w. The speed of Pi is wPiN, where PiN is the length of the
perpendicular from Pi on to the representation of 00 passing through PI. When 00
varies with time, it defines at time t an instantaneous axis of rotation, about
whi~h Pi instantaneously turns with speed wPiN. Note that relation (9.2.3) is
exact, since it is obtained by a limiting process in which lit -+ 0, and the term
00 x (ri - rd gives the relative velocity exactly. The angular velocity of a rigid
body, as opposed to an angular displacement, is definable as a vector without
approximation.
As before, the angular velocity 00 is independent of the particular choice of
reference vector rl since, with a second reference vector r2, we have
dri drl
- =- + 00 x (r· - rd
dt dt I
without slipping along a fixed surface. The point of contact of the body with the
surface is then instantaneously at rest since there is no relative motion. An
instantaneous centre can always be found in two-dimensional problems, where
dri/dt has no z component for all i, and w= wk. Strictly speaking the instan·
taneous centre is not unique, for if one point of the rigid body has zero velocity,
it can be regarded as rotating about that point and all other points on the axis of
rotation will also have zero velocity. However two-dimensional problems can
usually be regarded mathematically as problems in the (x, y) plane, and the
instantaneous centre is then taken to be in that plane. Given two points P1, P2
of the body in the (x, y) plane, with non-parallel velocities v 1, V2 , the instan-
taneous centre is at the intersection of the lines through P 1 and P2 perpendicular
to V1 and V2.
~xar.nple 9.2.1
We consider the motion of a rod AB in a vertical plane such that its ends are
always in contact with the ground and with a vertical wall, as in Figure 9.2.1.
Let the instantaneous centre be I. Since the velocity of A is horizontal and
that of B is vertical, it follows that I must be at the intersection of the vertical
through A and the horizontal through B. Let the length of the rod be l and let its
inclination to the vertical be 8. Then the velocity of A is lAO = lO cos 8 along
le
OA and the velocity of B is IBe = sin 8 along BO. The point of the rod which
has minimum speed is the foot of the perpendicular from I onto AB.
B ------lI
I
(}
I
I
I
I
I
I
I
I
I
0 A
lIxan1ple 9.2.2
A circular cylinder of radius at is enclosed inside a coaxial hollow circular
cylinder of radius a2 (>a 1 ). The space between them contains ball bearings of
radius Haz - ad. The inner and outer cylinders are.made to turn with constant
angular speeds Wt and W2 respectively. Show that if there is no slipping the
angular speed 0 f a ball is (W2 a2 - W tat )/ (a2 - a 1 ) and that the centre 0 f a ball
moves in a circle with angular speed (Wtat + W2a2)/(at +a2)' What is the length
of the circular arc on either cylinder with which the ball is in contact in unit
time?
In Figure 9.2.2 the rotations are assumed to be clockwise. In this particular
example the results hold for clockwise or anti-clockwise rotations provided that
the angular speeds are allowed to take both positive values (for clockwise
rotations) and negative values (for anti-clockwise rotations).
Let Wb, We be the angular speeds of the ball bearing and of the radius OC.
Let vq , Vr be the speeds of the points of the ball bearing in contact with the
inner and outer cylinders at Q and R (see Fig. 9.2.2). We have
Figure 9.2.2 Illustration for Example 9.2.2: the motion of a ball bearing
between coaxial cylinders.
252 Angular Vectors [9.3
Hence
Also
Let the radius OPt of the inner cylinder and the radius CP2 of the ball
bearing be vertical at zero time, when P t and P2 both coincide with P (see Fig.
9.2.2). In unit time the radius OPt will move through an angular distance Wt.
The radius OQ, where OQC is the line joining the centres of the coaxial circles
and the ball bearing, will move through an angular distance We. The arc on the
inner cylinder with which the ball bearing is in contact in unit time is
relative to Fj be CJ>;j, so that W;j =-Wj;, and W;j =0 if i =j. The question to be
investigated is whether or not the angular velocities combine in such a way that
(9.3.1 )
This relation is, in fact, satisfied and enables us to attach a meaning to the
statement that a rigid body or frame of reference can have simultaneously two
or more angular velocities which are additive in the usual vector sense.
To verify (9.3.1) we consider three vector functions of time rl , r2 , r3, which
are fixed relative to F 1 , F 2, F 3 respectively, and such that
Now ri moves as a rigid body with Fj, and Fi has an angular velocity W'ik relative
to Fk , so that
d:
( dr.) k =Wik x r. (9.3.3)
W ik x r = Wij x r + Wjk x r
or (Wik - Wij - W jk) x r = O.
r =xi +yj + zk
. dx. dy. dz
we have r=-I+-J+-k (9.4.1 )
dt dt dt'
dr dx. dy. dz di dj dk
but -=-I+-J +-k+x-+y - +z-. (9.4.2 )
dt dt dt dt dt dt dt
Note that although the components (x, y, z) of a vector r depend on the axes
chosen, there is no ambiguity about their derivatives provided it is understood
that they are referred always to the same set of axes.
Since i,j, k are three vectors which move as a rigid body with angular velocity
w, di/dt = w x i etc., and so (9.4.2) may be written
dr/dt=r+wxr. (9.4.3)
dw/dt = w+ W x W =W
so that the rate of change of w is the same in both frames. If dr/dt = v, we have
dv/dt = v+ W x v. (9.4.4)
But, by (9.4.3),
v=f+wxr+wxr, (9.4.5)
w x v = w x r + w x (w x r), (9.4.6)
9.4 ] Rotating Axes 255
(9.4.7)
The first term i' in (9.4.7) gives the acceleration of the particle as interpreted by
an observer using the rotating frame of reference. The second term wx r arises
only when the angular velocity varies, but the third and fourth terms, called
respectively the Coriolis acceleration and the centripetal acceleration, are present
even when the angular velocity is constant.
In Chapter 3 it was stressed that Newton's laws of motion are formulated
with respect to an inertial frame of reference, and such a frame has always been
assumed in previous applications. This does not mean, however, that we are
necessarily limited to such frames in all problems, and sometimes it is convenient
to use a frame which is not inertial. In such a case one must make appropriate
allowance for the acceleration in the equations of motion. If the position vector
r of a particle of mass m is defined relative to a frame of reference whose origin
o has an acceleration ao relative to an inertial frame, the equation of motion for
the particle is
where F is the force. If the frame of reference also rotates with angular velocity
00 relative to the inertial frame then, by (9.4.7), the equation of motion will be
that the second observer would attribute to the particle. Whether these terms are
regarded as modifying the acceleration or the force is a question of point of
view. For axes fixed on the earth's surface, gravity is regarded as a force of
attraction between the earth and a neighbouring mass. The earth's rotation also
produces a centripetal acceleration, which is often regarded as a small modi-
fication to the earth's gravitational force. It is then called a centrifugal force.
256 Angular Vectors [9.4
Example 9.4.1
A smooth wire is bent into the form of a circle, of radius a, and rotates about a
vertical diameter with a prescribed angular velocity w(see Fig. 9.4.1). A smooth
bead slides freely on the wire.
Let r be the position vector of the bead relative to the centre of the circular
wire, and let i be the rate of change of r in the plane of the wire. The equation
of motion is then
i + wx r + 2w x i + w x (w x r) =m -I F + g, (9.4.9)
where F is the reaction of the wire, which will be perpendicular to the wire. Of
the terms on the left-hand side of the equation, Wx r and 2w x i are normal to
the plane of the wire, and serve to determine the appropriate component of F
when r and i have been found. The component of the equation of motion
perpendicular to r in the plane of the wire gives
Equation (9.4.11) can be dealt with using the general theory for an energy
equation with the pseudo-potential
mg
(9.4.12)
where a is the value of () when (j = O. This equation also has oscillatory solutions,
although they are not harmonic. The periodic time is
where K is the complete elliptic integral with argument rI/2 (see Example
6.5.2).
d2 r dr
m-=-eE- e-xB. (9.4.13)
dt2 dt
The terms in (9.4.14) which involve f can be eliminated by the choke w = eB/
2m and then
.. eE.
r=---w.xr+wx ( )
wxr.
m
If the magnetic field is substantially constant and sufficiently weak for the
second and third terms on the right-hand side to be neglected compared with the
first, the equation approximates to
r= -eE/m.
It follows that the effect of a weak magnetic field is to cause the trajectory
obtained by its neglect to precess slowly, with angular velocity w. This is known
as Larmor precession. If, for example, the electron is in orbit about a nucleus of
charge e', then E = e'r/r3 and so
.. -ee'1-
r= r mr.
3
This is an inverse square law of force and the orbit of the electron is an ellipse.
The addition of a weak magnetic field causes this ellipse to precess slowly with
angular velocity w about an axis through the nucleus.
The conditions required for the validity of the approximation follow from
the equation of motion. They are
where r is a typical length defining the orbit. For an ellipse this can be taken as
the semi-major axis a, and then the conditions are
where
Reference to (7.5 .15) shows that T is the periodic time of the electron in orbit,
so that 21T/T is the mean angular speed.
9.5] Axes on the Earth '8 Surface 259
For many practical problems it is convenient to use a set of axes fIxed on the
earth's surface. The acceleration of a particle, as it appears to an observer using
such a frame of reference, will differ slightly from the acceleration relative to an
inertial frame because of the motion of the earth. The most significant effect
arises from the rotation of the earth about its axis, which amounts to 2ft radians
per sidereal day. This is equivalent to an angular velocity of magnitude
2n __
- - = 7·29 x 10 5 rad s 1. (9.5.1)
86 164
To define the frame of reference, we choose an origin 0 fixed on, or near, the
earth's surface at a distance BO from the earth's axis of rotation (see Fig. 9.5.1).
Let BO = b and let r be the position vector of a particle P relative to O. Let i,j, k
be a mutually orthogonal right-handed set of unit vectors such that i is hor-
izontal and points south, j is horizontal and points east, and k points vertically
upwards. Here horizontal means perpendicular to the vertical, and vertical means
the direction defined by a plumb line, which is a string fIxed relative to the
Figure 9.S.1 Illustration of the notation associated with axes on the earth's
surface.
260 Angular Vectors [9.5
earth's surface and supporting a particle suspended from one end. Because of the
rotation of the earth, the direction of the vertical differs slightly from the
direction defmed solely by the earth's gravitational attraction. We therefore
denote the latter by mg', where m is the mass of the plumb bob, and distinguish
it from mg = -mgk, which is equal and opposite to the force measured by the
tension in the string of the plumb line.
If we assume that the point B has no acceleration relative to an inertial frame
of reference, the acceleration of the particle Pis
(9.5.2)
(9.5.3)
where m now denotes the mass of the particle. The force F is the resultant of all
the forces except the earth's gravitational attraction mg'. The acceleration
d2 bjdt 2 of 0 arises from the earth's rotation with angular velocity w. It may
therefore be written
(9.5.4)
where the latitude A is the angle of elevation of the earth's axis of rotation above
the horizontal (see Fig. 9.5.1). With this substitution, equation (9.5.3) becomes
(9.5.5)
Let us now apply equation (9.5.5) to the bob of a plumb line fixed at the
point 0, so that the bob has zero acceleration relative to O. The force F will be
the tension in the string, which is defined to be -mg = mgk. Hence
where g~ is the value g' at O. In general the force of attraction g will vary from
point to point, partly because of the oblateness of the earth and partly because
it depends on the distance from the earth. In the general neighbourhood of a
fixed point near the earth's surface these variations are small (see §2.4). For
problems where they are small enough to be ignored, so that I g' - g~ I is
negligible, the equation of motion (9.5.5) can be written
(9.5.7)
for the general motion of a particle near the earth's surface. Since it is mgk
9.5] Axes on the Earth's Surface 261
which is readily measurable and which has in general the more immediate
practical significance, equation (9.5.7) is the form usually to be preferred.
To get some idea of the difference between g' and g, we can assume the earth
to be a sphere of radius re and that, approximately, b = re cos "A. Then, from
(9.5.6),
The maximum difference in magnitude occurs at the equator ("A = 0), where
, 2
go - g = re W
g g
d2 r/dt2 = f + 2w x r + W x (w x r),
Of these terms it is customary to neglect the centrifugal force -mw x (IW x r).
The ratio of its magnitude to mg is of the order rw 2 /g. Reference to equation
(9.5.8) giving the value of rew2/g shows that rw 2/g will be negligible in all
problems where r ~ re. The approximate equation of motion is then
On the other hand the Coriolis force, -2mw xi, gives rise to substantial
meteorological effects. When a pressure gradient is set up in the atmosphere the
air will initially tend to move along the gradient. But the Coriolis force, which is
always perpendicular to the velocity, tends to give the motion of the air a
component perpendicular to the pressure gradient. This in tum gives the Coriolis
force a component opposing the pressure gradient, and equilibrium is ap-
proached as the two forces tend to cancel each other. The air is then moving at
right angles to the pressure gradient rather than along it. This is called a
262 Angular Vectors [9.5
geostrophic wind. When a localised centre of low pressure is set up, the air will tend
ultimately to circle round it, giving rise to a cyclone. Consideration of the earth's
angular velocity shows that the air will circulate anti-clockwise in the northern
hemisphere, clockwise in the southern hemisphere, as viewed from above. At the
equator cyclones do not occur.
On a larger scale, the heating of the air at the equator causes it to rise, and to
be replaced by cooler air flowing from the poles. Thus a circulation is set up in
which the air flows from the poles to the equator, rises and returns in the upper
atmosphere. Because of the Coriolis force, the air does not flow due south or
north; it has a component of velocity towards the west, which causes it to blow
from the northeast in the northern hemisphere giving rise to the northeast trade
winds. Their counterpart in the southern hemisphere are the southeast trade
winds. The effect is more readily appreciated from the point of view of an
observer using an inertial frame of reference. Underneath the air flowing from
the poles to the equator is the earth rotating from west to east, and because of
this rotation the air has a component of velocity from east to west relative to the
earth.
Example 9.5.1
We consider the free trajectory of a particle under gravity, whose equation of
motion will be, from (9.5.9) with F = 0,
Now the effect of rotation is small, and the terms on the right-hand side of
(9.5.11) are in decreasing order of significance. The crudest approximation is
therefore obtained by neglecting rotation completely and gives
The integral of (9.5.13) will now include a first approximation to the correction
for the earth's rotation and we have
(9.5 .15)
The term -w x (ut 2 - ~gt3k) represents the small correction to the position
vector. For a particle which falls from rest relative to the surface of the earth,
9.5] Axes on the Earth's Surface 263
when Z =-h. For h = 100 m this gives y =0·022 cos A. m. It is very small, and
difficult to detect because of extraneous disturbances in the atmosphere.
It is instructive to consider such problems using an inertial set of axes. For
simplicity we consider the case of a particle at the equator dropped from relative
rest at a height h above the surface of the earth. Let re be the radius of the earth
and let oW be its angular velocity. We use inertial axes and specify the position of
the particle by polar coordinates (re + r, 8) based on an origin at the centre of
the earth. Initially r = h,;= 0, 8 = 0, iJ = wand the equations of motion are
from which
gt 2 ) w
. ( 1+-;:-
8= or 8= (t +gt3re )w
3
approximately. In this time the earth has rotated through an angle wt. The
relative angular displacement is therefore wgt 3 /3r e' This gives a relative linear
displacement on the surface of the earth of magnitude
as before.
For a projectile given a large initial speed in a nearly horizontal trajectory, as
in the flight of a rifle bullet, the term ut2 will dominate the term gt 3 k/3, at least
initially, and equation (9.5.15) approximates to
(9.5.17)
where I is the position vector of the bob relative to the point of suspension and
T is the tension (see Fig. 7.4.1). We now see, however, that this equation is in a
form suitable for use v,.'ith an inertial set of axes. Relative to axes fixed on the
earth's surface, equation (9.5 .17) becomes
where the small centripetal acceleration has been neglected. The vertical com-
ponent of (9.5.18) gives
- Tl . klml = k . (gk + i + 2w x i)
where -zk is the vertical component oft and r is the horizontal component. For
small perturbations about the equilibrium position, the most significant term on
the right-hand side of (9.5.19) is g, the other terms being small in comparison.
Also I. k = -1 approximately and so, if only the most significant terms are
retained, we have, as before,
Tim =g.
since, on a linear theory, only the most significant approximation for Tim is
required. For the linear theory, then, it is sufficient to regard the inertial axes as
rotating relative to the earth with the effective angular velocity -kw sin A. Since
the bob performs elliptic harmonic motion relative to inertial axes, it follows
that, relative to the earth, the ellipse rotates with an angular velocity -w sin A
about the vertical.
Alternatively one can argue directly from equation (9.5 .20). In polar co-
Exercises 265
The substitution
e=~-wsinA (9.5.22)
gives, when, as usual, terms involving w 2 are ignored,
- ' ' 1 ' = -gr;/1 ,
r··..),2 r~+2r~=0. (9.5.23)
These are just the equations which describe elliptic harmonic motion using the
polar coordinates (r, ¢). Equation (9.5.22) implies that the resulting elliptic
trajectory will, in the (r, 0) plane, rotate about the vertical with angular speed
-w sin A. A more direct interpretation of the result is to regard the pendulum as
oscillating in space with the earth rotating underneath it with the effective
angular velocity kw sin A, which gives the pendulum an equal and opposite
relative angular velocity.
Foucault first suggested that the effect could be used to demonstrate the
rotation of the earth. Care is required in setting up such a demonstration,
because w is so small that the effect can be concealed by other small higher
order corrections.
EXERCISES
9.l A rigid body is free to rotate about a fixed point 0, which may be taken as
the origin of a set of cartesian axes.
Angular displacements, each of magnitude n/2, are performed succes-
sively, first about Ox and then about Oy. Show that the result is equiv-
alent to a single angular displacement 2n/3 about an axis in the direction
i + j - k.
9.2 A rigid body is given an angular displacement 0 about an axis through the
origin 0 in the direction of the unit vector n. If the point with position
vector r moves to the point with position vector rt, show that
9.3 The vectors rl, r2, r3 =rl xr2 are three vectors of a system which is
displaced as a rigid body subject to the constraints ri . rj = constant for all
i, j. The displacement ori of ri is of the form ori = (Vi +IEi) ot,where Vi is
independent of ot and I Ei I ~ 0 as ~ O. ot
(i) Show that
266 Angular Vectors
where r = OP.
Hence show that, for some scalar A,
r = W -2 (v P - vo ) x w + AW,
I w x ·w 12r = w 2{(ao - a p ). w}w+ {(a o - a p ). w}W
+(a o -ap)x{wxw)}.
(i) Find the angular velocity of the disc, and show that it is horizontal if
(c 2 + 4a 2 )112 - C
cos Q =
2a
(U) Find the acceleration of the highest point of the disc, and show that it
is vertical if cos Q = a/c.
1 d 2' '2
- - (r (J) - rsin (J cos (J ,{..
r dt 'I' ,
1 d 2 2 •
- .- - (r sin (J cp).
r sm (J dt
Deduce from this result that a particle whose position vector is r describes
a circle of radius a with angular speed w.
9.16 A beam of particles, each of mass m and charge e, is emitted from a point
source. The particles are subject to a uniform magnetic field B and they
have initial velocities whose components parallel to B are all of magnitude
v. The equation of motion for a particle is
Show that the effect of the field is to focus the beam at a point whose
distance is 2rrmv/eB from the source.
This result is used in electromagnetic focusing.
9.l7 A bead slides on a smooth circular wire which rotates in its own horizontal
plane with constant angular speed w about a vertical axis passing through a
point A of the wire. Let 0 be the centre of the circular wire and let (J be
the angular displacement of the bead from ~A. Show that
where ex = g/aw2 •
9.19 A particle is constrained to move in a straight line due north at latitude A
with constant speed u. Calculate the force on the particle.
(i) Show that if a train of mass m travels due north at 100 km h -I, there
is a small horizontal force on the eastern rail of about 2 x 10-4 mg sin A.
(w=7'29x 10- 5 rads- l ,g=9'8l ms- z ).
(ii) Show that if the point of suspension of a plumb line is constrained to
move due north with constant speed u, the line will be inclined at an
angle tan -I (2uw sin A/g) west of the downward vertical.
(iii) The free surface of a liquid moving in the same way sets itself
perpendicular to the direction of the plumb line, so that there is no
apparent tangential force on the liquid. Show that if a tidal current
runs due north in a channel of breadth b, the height on the east coast
exceeds that on the west coast by an amount 2buw sin A/g.
For the Irish Channel show that this is about 1'6 m. (b = 90 km,
u= 1·5ms- l ,sin'A=0·8).
9.20 A particle is projected with speed u on a smooth horizontal plane at
latitude A. Show that, approximately, the particle will describe an arc of a
circle of radius u/(2w sin 'A), where w is the angular speed of the earth.
9.21 Relative to axes rotating with angular velocity w, the equation of motion
of a particle subject to a constant gravitational force can be written in the
form
r= -gk - 2wx i - w x (wx r).
On the assumption that the effect of rotation is small, show that the
third approximation to the trajectory of a particle falling from rest at
r = 0 is
In the case of the earth the final term is negligible. Indeed the
refinement is not strictly justified. It is comparable in magnitude with
errors already incurred in the neglect of local variations in g.
9.22 A projectile is fired due north, from a point on the earth's surface whose
latitude is A, at an angle ex to the horizontal, where ex> 0, 'A < n/2. Show
that the point where it strikes the earth will be east of the vertical plane
through the point of projection if tan ex < 3 tan 'A.
9.23 A particle moves freely near the surface of the earth. Derive an approxi-
mate equation of motion for the particle in the form
f+ 2wx i = g.
of 0 relative to the centre of the earth. Let the force on a particle of mass
m, arising from the attraction of the earth, be mg'(r) when the position
vector of the particle relative to 0 is r.
Show that the equation of motion of the particle is
r + 2w xi + w x (w x r) + w x (w x a) = g'(r).
r + 2w x i - 3w2 xi = O.
Assume that r ~ a and that the mutual attraction between the particle
and the satellite is negligible.
10 Moments
(10.1.1)
or, equivalently,
(10.1.2)
Another special case is the geometrical centre or centroid, for which each point
is given equal weighting. Then we have
nOC =~OPj.
j
*
associated mass centre is being discussed, this will be made clear in the context.
Note that the position of C is well defined by (10.1.1) provided Ljmj O.
The following general results are of practical use in the calculation of the
position of the weighted centroid.
Let a distribution mj at Pj (i = I, 2, ... ,n) have the weighted centroid C1 •
Let a second distribution mj at Pj (j = n + 1, n + 2, ... ,p) have the weighted
centroid C2 • Then the two distributions taken togeth~r have a weighted centroid
which coincides with that of Ljmj at C, and Ljmj at C2 • For we have that
from which the result follows. The result clearly generalises to a distribution
involving several subdivisions. As a simple example, the distribution m j , m2 , m 3
at the points P j , P 2 , P 3 is such that m2 at P2 and m3 at P3 have a weighted
centroid Cj on P2 P 3 , where m2P2 C1 = m3Cj P 3 • Then mj at Pj and m2 + m3 at
C1 have a weighted centroid Con PjC j , where m j Pj C=(m2 +m3)CC j . The
weighted centroid of the original system coincides with C.
An immediate corollary of the above result is that if a distribution can be
divided in such a way that the weighted centroids of the separate subgroups are
at a given point, or on a given line, or on a given plane, then the weighted
centroid of the whole distribution coincides with the given point, or lies on the
given line, or on the given plane. Thus, for example, a distribution of pairs of
points, each pair having the same weighting m and equal and opposite position
vectors relative to some point C, has its weighted centroid at C. Such a
distribution is said to have central symmetry. Again two equally weighted points
which lie on either side of a given line (or plane) and are equidistant from the
line (or plane) have a weighted centroid on the line (or plane). Any distribution
of such pairs of points will also have a weighted centroid on the line (or plane).
Note that it is not necessary for the pairs of points to be mirror images.
Equation (10.1.1) is readily generalised to include a continuous distribution.
This extension will be required when dealing with the dynamics of rigid bodies.
Let each point P of a given region have associated with it a density per), which is
a function of the position vector r of P (see Fig. 10.1.1). If an elementary
volume l57 containing P has a weighting om, the density at P may be regarded as
the limit
om dm
p= lim - = -
IiT->O Or dr'
on the assumption that such a limit exists. In particular, if om is the mass of the
element or, then p is just the mass density. It follows that the mass of the given
region is
jp(r) dr,
o
Figure 10.1.1 The position vector of a point P contained in an elementary
volume or of a continuous distribution.
10.1] First Moment 273
where the integration is taken over the volume of the region concerned. The first
moment is defined as
Jrp(r) dr
and the weighted centroid is the point e whose position vector is given by
oe = frp(r) dr. (10.1.3)
fp(r) dr
oe =frdr (10.1.4)
fdr'
With an appropriate interpretation of dr, the results can be used to calculate the
first moment and the position of e for a continuous distribution in one, two, or
three dimensions (line, surface, or volume distributions).
Example 10.1.1
Let the line density along the x-axis be proportional to xn. It follows, from
(10.1.3), with dr = dx, that the weighted centroid of a line distribution OA, of
length a, along the x-axis is a point e 1 such that
oe l =
r~ax n n dxdx =--2
+1 n +1
a. (10.1.5)
OX n+
The calculation is similar for a surface density distribution p = poxn over the
rectangular surface with adjacent sides OA, oflength a, and OB, of length b (see
Fig. 10.1.2). Relation (10.1.5) then gives the x coordinate of the weighted
B
- ---------------l
I
1
b
I
I
I
jY~..u-"
I
I
---...
o x dx C1 A
a - - - - ' - - -.....
- IUg rPox n dx dy
OCz =
IUob PoX n dx dy
(i I~ x n+ 1 dx IS dy + jJ~ xn dx IS y dy)
=
I~ xn dx IS dy
n+1
= - - ai+!bj. (10.1.6)
n+2
Example 10.1.2
Let us consider the centroid of a uniform density distribution over the following
regions (see Fig. 10.1.3):
o o
A
Figure 10.1.3 Illustration for Example 10.1. 2.
10.2] Second Moment 275
(ii) When the strip refers to an element of area, the centroid C1 of the strip
clearly lies on the median of the triangle OAB, and C2 is then the mid point of
AB. It follows that the centroid of the triangle lies on OC 2 • To find the position
of C on OC 2 we note that, as the position of C 1 varies along OC 2 , the weighting
of an elementary strip varies in proportion to x = OC 1 since, by similarity, the
length of the strip parallel to AB varies in proportion to x. We therefore consider
a line density distribution along OC 2 proportional to x. With n = 1 in (10.1.5)
this gives OC = ~ OC 2 •
(iii) For a distribution over the surface of a cone, the elementary slice refers
to a closed contour, whose thickness at a point P on the circumference is
proportional to OP, the distance from the vertex O. The weighted centroid C 1 of
this slice is that of a line distribution around the circumference, with a density
proportional to the distance from 0 of the points on the circumference. By
similarity, C1 will lie on OC 2 , where C2 is the weighted centroid of the
circumference of the base, weighted according to the distance from O. This part
of the calculation will depend on the details of the base profile.
When C2 is known, the position of C on OC 2 is obtained from a line
distribution along OC 2 with density varying in proportion to x = OC 1 since, as x
varies, the circumference of the ring, and hence its weighting, varies in propor-
tion to x. This gives OC = ~ OC2 .
(iv) For a solid cone, C2 is clearly the centroid of a uniform density
distribution over the surface of the base of the cone. To calculate the position of
C on OC 2 , a line distribution along OC 2 with weighting proportional to the
cross-sectional area of an elementary slice is required. Since the variation of this
area is proportional to Xl , we have, from (10.1.5) with n =2, OC =~OC2.
The second moment about a given axis, of a set of scalar quantities mi associated
with a set of points Pi, is defined as
(10.2.1)
pr = rt sin2 e = (n x ri?
=rr - rr cos2 (} =rr - (n . ri? ,
and so 1= "EmiPr = "Emi(n x ri)2 = "Emz{rr - (n . ri)2}. (10.2.2)
i i i
276 Moments [10.2
Pi
The analysis may be carried still further with the choice of a set of cartesian
axes based on 0 such that rj has components (X{,Yi, Zi) and n has components
(nx, ny, nz), with ni + n~ + n~ =1. Then we have
The coefficients lxx, Iyy, I zz are just the moments of inertia about the
respective coordinate axes Ox, Oy, Oz. The coefficients Iyz, Izx, Ixy are called
products of inertia. It follows that, with a knowledge of these six coefficients,
the moment of inertia about any line through 0, with direction cosines
(nx, ny, n z ), may be calculated using (10.2.3).
Although we shall not prove it here, it is in fact true that there always exists a
set of axes based on 0 for which I yz =Izx =Ixy =O. They are called principal
axes and the associated moments lxx, I yy, I zz are called principal moments of
inertia. Relation (10.2.3) for the moment of inertia about a line through 0 with
direction cosines (nx, ny, n z ) simplifies to
(10.2.5)
when principal axes are used. The simplicity resulting from their use makes such
axes a desirable choice in practical applications.
10.21 Second Moment 277
the moment of inertia is the same for all axes through 0 without restriction and
the products of inertia will vanish for any set of axes. This is referred to as total
kinetic synunetry. It is well to remember that for this to be so it is not sufficient
that Ixx =Iyy =Izz . It is also necessary that the axes are principal ones so that
the products of inertia vanish.
It is often convenient to introduce a length k, called the radius of gyration,
defined by
(10.2.6)
(10.2.7)
radius of gyration about a parallel axis through the weighted centroid C of the
system Then
(10.2.8)
Since Limiri = 0, by (10.1.2), the last term is zero and so, finally, we obtain the
result k 2 = d 2 + k~ .
Yi
o x
Figure 10.2.3 Illustration for the perpendicular axes theorem.
radii of gyration about the axes Ox, Oy, Oz respectively. It follows at once that
(10.2.10)
where Pi is the distance from the given axis, and p is the density at the field
point associated with the continuous distribution.
Example 10.2.1
As in Example 10.1.1 and Figure 10.1.2, let the line density along Ox be
proportional to xn. The radius of gyration k oy about an axis Oy perpendicular
to Ox is given by
(10.2.11)
(i) About an axis inclined at an angle a « n 12) to Ox, the radius of gyration is
koy sin a, because the distance from the axis of every point in the distribu-
tion is reduced by the factor sin a compared with the distance when
a = n12.
(ii) From Example 10.1.1, the weighted centroid C is at a distance (n + 1)al
(n + 2) from O. Therefore, by the parallel axes theorem, we have
2 -_ key
koy 2 + (n+I)22
- - a,
n+2
and so
2
key -
_ {nn ++ 3I - (nn ++ 2I )2} a 2 _ en + l)a
- (n + 2)2 (n + 3) ,
2
(l0.2.12)
where key is the radius of gyration about an axis through C parallel to Oy.
In particular, for a uniform density distribution (n = 0), we have
k oy
2 =a 2 /3 , cky
2 =a 2 /12 . (10.2.13)
(iii) The calculation is essentially unchanged for a surface density P = poxn over
the rectangular surface with sides OA=a, OB=b (Fig. 10.1.2). The formal
calculation is
k2 - r:..:o=-"'0-i:-_
fb px 2 dx dy
_"-
oY - fUS p dx dy
_ °
f a0 x n + 2 dx fb dy
- f~xn dx fS dy
n +1 2
=--a
n+3
Ix = -
y
f° f°
a b
pxy dx dy =- J f° poxn+ly dx dy =-Po 2(n +b2) .
0
a b an + 2 2
(iv) For a uniform density distribution over the rectangle, the radius of gyration
about the centroid C, where OC = ! (OA + OB), is given by k~y = a 2 112 for
an axis parallel to Oy. By analogy, we have k~x =b 2 /12 about an axis
through C parallel to Ox. By the perpendicular axes theorem, the radius of
gyration about an axis through C perpendicular to the plane Oxy is there-
10.21 Second Moment 281
fore given by
(10.2.14)
The principal axes at C are parallel to Ox, Oy and their mutual perpen-
dicular Oz, since the products of inertia vanish by symmetry. The radius of
gyration k about any other line through C with direction cosines
(nx, ny, n z ) is given, with the help of (10.2.5), by
(10.2.15)
Note that, for a uniform square of side a, the radius of gyration about any
line through C in the plane ofthe square (so that n z = 0) is such that
(10.2.16)
(l0.2.17)
since n is a unit vector. A uniform cube therefore has the same moment of
inertia about any line through its centroid.
Example 10.2.2
(i) The radius of gyration of a uniform line density around the circumference of
a circle, of radius a, is equal to a about an axis Oz through the centre 0 and
perpendicular to its plane. This follows immediately from the fact that
every element of the distribution has this radius of gyration.
About perpendicular axes Ox, Oy in the plane of the circle, the radii of
gyration are equal by symmetry. From the perpendicular axes theorem,
they must therefore be equal to a/2112. It is clear from symmetry that this
is the radius of gyration about any axis through 0 in the plane of the circle.
(ii) For a uniform surface density over a circular disc, of radius a, the radius of
282 Moments [10.2
a
where m= fo 2rrpr dr =rra 2 p.
Hence
and so
10.3] Vector Moment 283
Hence, by symmetry,
(10.2.18)
ia
and k 2 = 2 for any axis through O.
(iv) For a uniform volume density throughout the volume of a sphere of radius
a, we can calculate the radius of gyration about any axis through 0 from
that of elementary spherical shells of radius r. The volume of the elementary
shell is 41Tr2 or, its moment of inertia is
where
Hence (10.2.19)
(v) The radii of gyration calculated in (i) - (iv) are also appropriate for, respec-
tively: (i) a semicircle; (ii) a semicircular disc; (iii) a hemispherical shell;
(iv) a solid hemisphere. This follows from the fact that the two halves which
constitute a complete circle or sphere have, by symmetry, the same radius
of gyration. It must therefore be the same radius of gyration as that of the
complete configuration. Note that the semicircular distributions have the
same radius of gyration about any coplanar axis through O. The hemi-
spherical configurations have the same radius of gyration about any axis
through O.
o r
since pp' and F are parallel vectors with zero vector product.
The combined moment of several line vectors F 1, F 2 , ••• all of whose lines of
action pass through the point P is
r x F 1 + r x F 2 + ... = r x (F 1 + F 2 + ... ),
which is the moment of a line vector passing through P equal to their resultant.
For a general system of vectors the notion of a resultant is still applicable, in the
usual sense of a free vector. But for a system of line vectors it is useful to
introduce the following definition of equivalence.
Definition
Two systems of line vectors are said to be equivalent if: (i) they have the same
resultant; (li) they have the same moment.
It is sufficient that they have the same moment about one point; it can then
be shown that they have the same moment about all other points. For let
Fj, Fj be typical members of the two systems, and let Pj, pj be points on their
respective lines of action. If we are given that
= 0'0 x ~F~
• J
+ ~OP~
• I
x F~I
J I
10.4 ] Parallel Vectors 285
= ~(O'O
•
+ op)
J
x F~J
J
- , x F'j.
= ~O'Pj
j
and this defines a point C, the weighted centroid of the Fi associated with the
points Pi. A necessary condition here is that the resultant R must not be zero,
and provided this is satisfied the system is equivalent to a single vector Rn whose
line of action passes through C. Note that the position of C is independent of the
direction of n.
A particular application of this result is to the gravitational forces acting on a
system of particles (or, in the limit, to the continuous distribution of parallel
forces acting on the elements of a continuous distribution of mass). Here
Fi = mig, R '£imig is the total weight, n is in the direction of the downward
=0
r
P 1 r - - - - - - 7 I P2
F
We return to the general system of parallel vectors, and to the special case
R = 0 for which there is no single equivalent vector. The most that is possible by
way of reduction is to separate the vectors into two groups, those for which
Fi > 0, and those for which F j < O. Since 'LiFi =1= 0 for each group, there is a
single equivalent vector for each group separately. These two vectors will be
equal in magnitude, opposite in direction and each will have a definite line of
action. Apart from the trivial case in which the two lines of action coincide and
the system is equivalent to a null vector, the equivalent system consists of two
equal and opposite parallel vectors a definite distance apart. Such an entity is
called a couple and no further reduction is possible. We list below three
important properties of couples.
(i) A couple has the same moment about any point. Let PIP 2 be the common
perpendicular to the lines of action of equal and opposite line vectors F and
-F (see Fig. 1004.1). The moment about any point 0 is
Example 10.5.1
We have already seen that, in general, a system of line vectors is not equivalent
M+O'O xR
Figure 10.5.1 Equivalent systems at 0 and 0'.
288 Moments [10.6
to a single line vector. One might, however, consider the possibility that such a
system is equivalent to two line vectors, one of which acts along a prescribed
line.
Let 0 be a point on the prescribed line, and let n be a unit vector defining its
direction. We assume that the given system is equivalent to the force and couple
combination (R, M) at 0, and look for an equivalent system consisting of
F 1 = Fl n at 0 and F2 = R - Fl n along a suitable line. Let the position vector of
a typical point on the line of F 2 be r relative to O.
The couple M can arise only from F 2 , since F 1 has no moment about O.
Hence we have
rxF 2 =rx(R-Fln)=M.
M.R
Fl = - - .
M.n
Note that this can only be done provided M . n =f= 0 and so the prescribed line
cannot be chosen in a completely arbitrary fashion. It must not be perpendicular
toM.
With Fl and hence F2 determined, the relation r x F2 = M is, in effect, an
equation giving the line of action of F2 . To derive an explicit expression for r,
we write it in the form
r =a +AF2 ,
F2 xM
a=--2-'
F2
Let a system of line vectors Fi be localised at points Pi with position vectors rj.
Let the system be eqUivalent to
at Pt. Let the rj suffer a small rigid body displacement ~ri defined by a
translation ~rt of P t and an angular displacement 66n about Pt. We then have
This expression is the work of the system correct to the first order, which is
sufficient for our purposes. Note that only R, M and the small rigid body
displacement are required to evaluate (10.6.2). Since R, M are the same for all
equivalent systems, it follows that the work done in any given rigid body
displacement is the same, to first order, for all equivalent systems.
We note in particular that the work done in any small rigid body displace-
ment is zero, to first order, for a null system in which R = M = o. Conversely
(10.6.2) can only be zero for all small rigid body displacements if R = M = 0 and
the system is equivalent to a null vector. These results form the basis of the
discussion of virtual work in Chapter 12.
If the riare differentiable functions ofa parameter t and, as in §9.1,
where 1Ei 1-+ 0 as cSt -+ 0, we can divide (10.6.2) by cSt and take the limit as
cSt -+ 0 to get, exactly,
1::Fi . Vi =R . VI + M . w. (10.6.3)
i
EXERCISES
10.1 The lines QS, RT intersect at 0, and P is a point such that PT = RO.
Show that the centroids of uniform density distributions over the areas
QRST and PQS coincide.
10.2 Establish the following rule, known as Routh's rule.
For the bodies listed below, the moment of inertia about an axis of
symmetry of a uniform density distribution with total mass m is
m
-(sum of squares of the perpendicular semi-axes),
n
290 Moments
where
10.3 An isosceles triangle ABC is such that AB = AC, BC=a and the
perpendicular height from A to BC is h. Show that a uniform density
distribution over the area ABC has a radius of gyration equal to
(h 2 /2 + a2 /24)1/2 about an axis through A perpendicular to the plane
of ABC. What is the interpretation of this result as: (i) h -+ 0; (ii) a -+ O?
Deduce that the radius of gyration of an n sided regular polygon,
about an axis through its centroid perpendicular to its plane, is given by
(i) mh2/12 about an axis through the centroid parallel to one pair of
sides, where m is the total mass and h is the perpendicular distance
between the pair of sides;
(ii) m(a 2 + b 2 )/3 about an axis through C perpendicular to the plane of
the distribution, where 2a, 2b are the lengths of the adjacent sides.
10.6 The two sets of axes Oxy, Ox'y' have a common origin 0 and lie in the
plane of a two-dimensional distribution of matter. The angle between Ox'
and Ox is a. Show that
, .
x' = x cos a + y sin a, y = y cos a - x sm a.
Show that
Deduce that the two values of a for which Ix'y' = 0 differ by Ti/2 and
hence that the principal axes are perpendicular.
10.7 A plane distribution of matter has the same amount of inertia about two
lines which lie in the plane and meet at a point O. Show that the
Exercises 291
'"
~myz = (~m)yczc + ~my I z , ,
10.10 A circle in the (x, y) plane has radius a and centre at the 'Origin of
coordinates. A second circle has radius a/2 and centre at (a/2, 0).
Find the centroid of the area between the two circles.
Find the principal radii of gyration at the 'Origin of coordinates and at
the centroid.
(C )112 a,C~4y/2
56 a, (~r2 a).)
292 Moments
10.12 Three uniform line distributions OX, OY, OZ, each of length i, lie along
the coordinate axes. Show that any axis through 0 is a principal axis with
associated moment of inertia 2mP /9, where m is the total mass. Show
that the principal moments of inertia at the centroid Care
Show that relative to axes parallel to the original axes, but with the
origin at C, the associated moments of inertia are all equal to mi 2 /6 and
the products of inertia are all equal to mZ 2 /36.
10.l3 Three line vectors have lines of action passing through Pi, P22 3
respectively. They are represented in magnitude and direction by OQi ,
OQ2, OQ3, where Qi' Q2, Q3 are the mid points of P2 P 3 , P 3P i , P i P2
respectively. Show that they are equivalent to the system represented by
OP i , OP 2 , OP 3 passing through Qi, Q2, Q3'
10.l4 The system of line vectors
i x (F x i), j x (F x j), k x (F x k)
RxM
r=--+AR
R2
for arbitrary A.
10.l7 Show that a system of line vectors is equivalent to a null vector if their
moments about two points are both zero and their resultant has no
component in the direction of the line joining these two points.
10.18 Investigate the possibility of equivalence between a general system of line
vectors, and a combination of two line vectors acting at two prescribed
points.
10.19 A typical mass m of a system has a position vector r relative to the mass
Exercises 293
=-G -
rp
(
~m+ - 3
1 (Ixx+1yy+l -31)+'" ) ,
2rp
zz
where Ixx + Iyy + I zz = ~mr2 and I is the moment of inertia of the mass
system about CPo
Let there be a particle mp at P. Show that the moment about C oft~e
forces on the system arising from the attraction of mp is approximately
_3Gmp
M - - - 5- :I; mer x rp)(r . rp).
rp
Show that, relative to the principal axes at C of the system, this expres-
sion can be written
V=-.!. i iGmjmj .
2i=lj=1!rj-rt!
i*j
Let the three moments of inertia referred to coordinates at 0 be lxx,
Iyy,/zz. Derive the Lagrange-Jacobi identity, namely
d2
dt 2 (Ixx+Iyy+Izz}=4(T+E),
We start from the fundamental equation F = rna for the motion of a single
particle. It follows immediately that
where the summation is over a system of particles. For economy of notation the
suffix i labelling a particular particle is omitted in (11.1.1), so that F represents
the resultant of the external and internal forces on a typical particle of mass rn,
position vector r and acceleration a. When the internal forces form a null system
(that is, a system with zero resultant and zero moment), as is the case when they
occur in equal and opposite pairs in the same line, they do not contribute to the
left-hand sides of either of equations (11.1.1) and the summation is then
confined to the external forces only. The same is true of a rigid body if it is
regarded as a system of particles obeying the same rules. There are, however,
logical difficulties associated with this assumption, one of which is that there are
limiting processes involved in the mathematics since rigid bodies are modelled as
continuous distributions of matter. Another difficulty is that a model in which
the internal forces occur in equal and opposite pairs acting along the same line is
not realistic, since there are electromagnetic forces involved which do not have
this property. Indeed the internal forces in real solids are not adequately
described by classical mechanics. However, we can proceed with the dynamics of
rigid bodies if we assume that equations (11.1.1) are valid, with the summation
confined to the external forces only. Since this assumption is justified under
conditions which are weaker than those invoked for a system of particles, it is
better to accept the equations themselves as basic postulates and to regard their
derivation from the motion of a system of particles as a motivation rather than a
proof. This point of view makes the minimum necessary assumptions, and
no discussion of the internal forces of cohesion is required. The ultimate
justification, as always, depends on the correspondence between theory and
observation.
We use the notation
~F = R, ~r x F = M, (11.1.2)
where F now refers to a typical external force. Note that the resultant R
depends only on the external force system but that, in general, the couple M
depends also on the reference point 0 relative to which r is measured. When this
11.1 ] Equations of Motion 295
d dr
= mere x ao + dt ~r x m dt ' (I 1.1.4)
where the suffix c in general refers to the mass centre C and, in particular,
me = ~m is the total mass. Note that the second term on the right-hand side
depends only on the motion relative to O. It is the moment of the relative mass
acceleration and can be written in the alternative form
d2 r d dr
~rx m- =-~rx m-
dt 2 dt dt
o 30
because dr/dt x m dr/dt = O. Now k(r x m dr/dt) is just the moment of the
relative momentum about 0, more usually called the relative angular momen-
tum. It follows that the moment of the relative mass acceleration is equal to the
rate of change of relative angular momentum.
We use the notation
p= kmV=meVe (11.1.5)
dr
L=krxm- (11.1.6)
dt
for the relative angular momentum. Note that L, like M, depends on the
reference point 0 and will be written Lo when this dependence needs to be
stressed.
We can summarise the foregoing notation and results in the equations
*'
(ii) The general case re x 80 0.
In most applications it is safer and advisable to take moments about a point 0
for which re x ao = 0, othelWise this term makes it awkward to handle (11.1.8)
since ao is not known a priori. If there is a compelling reason to take moments
about a general point for which re x 8 0 =#:'0, it is better to proceed as follows.
We first note, from case (i) with moments taken about the mass centre C, that
where Me and Le now refer specifically to the moment of the external forces
an(the relative angular momentum about C. Next we appeal to §10.5 to infer
that the system of external forces, which is equivalent to the vector couple combi-
nation (R, Me) at C, is therefore equivalent to (R, M = Me + re x R) at 0
(compare Fig. 11.1.2 with Fig. 10.5.1). The corresponding results for the mass
acceleration system are that it is equivalent to (dp/dt, dLc/dt) at C and
(dp/dt, dLe/dt+ re x dp/dt)at 0.1t follows immediately that, without restriction,
R=dp dL dp
M=_e +r x- (11.1.11)
dt' dt e dt'
where M is now the moment of the external forces about O. The rate of change
of angular momentum, however, is calculated relative to C with a correction
equal to the moment about 0 of dp/dt localised at C.
It is also possible to deduce the moment equation given in (11.1.11) from
first principles, and it is perhaps instructive to see what lies behind the analysis
leading to this equation. The basic relation is M = ~r x rna and we rewrite this
relation in the form
ctp
R 01
ctp
R dl
The second tenn ~rc x ma is, by (11.1.7), rc x dp/dt and the third tenn
~ {(r - rc) x mac} is zero because ~mr = (~m )rc. The final result agrees with
the second of equations (I 1.1.11 ).
One useful application of equations (11.1.11) is that for which the system
divides naturally into several component subsystems, for example two or more
rigid bodies moving under constraints. In these circumstances the notion of a
single mass centre for the whole system is not so convenient, and it can be more
useful to preserve the identity of each component in the calculation of the mass
accelerations and their moments. For this purpose we observe that, although
(11.1.11) was derived for the system as a single entity, the arguments are still
valid if the system is split into several components. Equations (11.1.11) are then
replaced by
(11.1.12)
where the notation implies that each tenn is calculated from the linear momen-
tum and the relative angular momentum of each subsystem separately. Note that
the calculation for a particular subsystem is referred to its own mass centre.
The results so far are applicable to any system of particles. We now specialise
to the case of a rigid body and consider the point 0 to be moving with the body.
Then the velocity relative to 0 of any particle of the body is just w x r, where w
is its angular velocity. The relative angular momentum about 0 is
dr
L= ~r x m dt = ~mr x (w x r) = ~m{r2w - (r. w)r}. (11.1.13)
Equations (11.1.14) simplify when the axes are chosen as principal axes,for
then the products of inertia all vanish. To denote that principal axes are being
used, we introduce the notation
(11.1.15)
(11.1.16)
The components of the angular momentum are now proportional to the cor-
responding components of the angular velocity. Note, however, that L and 00 are
not in general parallel vectors. For this to be so the body must rotate about a
principal axis.
The simplest application of the foregoing theory is to a rigid body which is at
rest. Then the linear and angular momenta are both zero and so, by (11.1.7) and
(I 1.1.8), the external forces form a null system. Conversely, if the 'externaf
forces form a null system, it follows from (11.1.7) and (11.1.8) that the linear
and angular momenta are constant. Thus the velocity of the mass centre is
constant and, by (11.1.14) or more directly (11.1.17), the angular velocity is
constant. In particular if they are zero at one particular instant, they remain zero
and the body is permanently at rest.
The dynamical applications will be confined initially to two-dimensional
motion, in which all particles move parallel to the (x,y) plane and ,00= wk, say.
Usually only the component of relative angular momentum perpendicular to the
plane of the motion is required, namely L z = /zzw, where /zz = ~m(x2 + y2) is
the moment of inertia about the axis through 0 perpendicular to the plane of
the motion. If M is the moment of the external forces about this axis we have
d
M=-/
dt zz w (11.1.18)
Example 11.1.1
A uniform cube, of mass ml and side 0, stands on a rough plane. A uniform
circular cylinder, of mass m 2 , diameter d and length d, rests with its curved
surface in contact with the plane and with one of the vertical sides of the cube.
The plane is gradually tilted about an axis parallel to the line of contact of the
cube and cylinder. The coefficient of friction for every contact is p.
Show that if p < 1, equilibrium is broken by the cube slipping down the
plane and the cylinder rolling down the plane. The angle of inclination 0: of the
300 Rigid Bodies [ ILl
tan a= pml .
ml + (1 - /l)m2
The system of forces is shown in Figure 11.1.3. Note that the frictional forces
Fl and F2 must act up the plane to prevent sliding. The tangential component
F3 of the contact force between the cube and the cylinder must then be in the
direction shown in order that the force system should have zero moment, for
eqUilibrium, about the axis of the cylinder. In fact since all the forces on the
cylinder act through its mass centre except F2 and F 3 , we must have
F2 = F3 = F (say).
Resolving along and perpendicular to the plane for the cube and the cylinder,
we get
Accordingly we have
and so N2 > N3 if tan a < 1. Since F2 = F3 this means that F 2/N z < F31N3 if
tan a < 1. Since /l is the same at all contacts, it follows that slipping will occur
between the cylinder and the plane before it occurs between the cylinder and
the cube if tan a < 1.
We now consider how eqUilibrium is broken when the cube is about to move.
Let us assume that tan a < 1 and that the cube slips. Then
ILl ] Equations of Motion 301
F =J.JN3 =--E-
l+IL
m2g sin a
'
mig sin a + N3 = FI = J.JNI = IL(mlg cos a + F).
or tana= .I!ml
ml + (1 - IL)m2
and tan Q < 1 if IL < 1, which is consistent with our original assumption.
It remains to consider the condition for which equilibrium is broken by the
cube toppling about its lowest edge. When this is about to happenNl will have
no moment about this edge, and so the moment equation about it gives
so that
tan Q = (1 + Il)ml
----'--~-"---
(1 + Il)ml + {l - 21l)m2
Hence equilibrium will be broken by slipping between the cube and the plane if
or
For 0";;; Il";;; 1, the maximum value of the left-hand side of this inequality is
attained when IL = 2 - 3 1 / 2 and is (2 - 3 1 / 2 )2/(3 1 / 2 - 1)2 < 1. It follows that
the above inequality is always satisfied for 0 .,;;; Il";;; 1 and so the cube will slip
before it topples.
Example 11.1.2
A uniform rod AB, of mass m and length 21, is suspended by two strings OA, OB
of equal length attached to a fIXed point O. The rod is at rest in a horizontal
302 Rigid Bodies [ ILl
B
cr'~x
10
Figure 11.1.4 Illustration for Example 11.1.2.
position, and each string makes an angle Q with the horizontal (see Fig. 11.1.4).
If the string OB is cut, show that the tension in OA is instantaneously
reduced in the ratio 2 sin 2 al(l + 3 sin 2 a).
When OB is cut, the acceleration of the mass centre C of the rod is the
acceleration of A plus the acceleration relative to A. The end A moves in a circle
with the string as radius but, since the initial angular velocity is zero, it will have
no radial acceleration initially. Let the transverse acceleration of A be x. Let the
initial angular acceleration of the rod be e.Since its initial angular velocity is
zero, the acceleration of the mass centre has components x perpendicular to OA
and 10 perpendicular to the rod.
The equations of motion are therefore
IT sin a =!mI2e.
The last equation is obtained by taking moments about C and using the result
that the moment of inertia of the rod about a perpendicular axis through C is
t ml2 . Elimination of 0 from the last two equations gives
T= mgsina
1 + 3 sin 2 a
11.1 J Equations of Motion 303
This compares with the value mg/2 sin a when the rod is at rest supported by
both strings. There is a reduction in the ratio 2 sin 2 ,a1(l + 3 sin 2 a).
Example 11.1.3
A light, inextensible string passes over a pulley of mass m, radius a and radius of
gyration k about its axis. Particles, of masses m 1 and m2 (m2 > m d, are sus-
pended from the ends of the string. The pulley is sufficiently rough to prevent
the string from sliding and motion is resisted by a frictional couple M at the axis.
Find the acceleration of the masses when the system is released from rest.
Let x be the displacement of either particle. The equations of motion are
M«m2 - ml)ga
Example Il.1.4
We consider the problem of a body of circular cross-section rolling and slipping
down an inclined plane. We shall assume that the body is symmetricai about a
circular cross-section, and that the mass centre coincides with the geometrical
centre of this circular cross-section. This enables us to treat the problem as a
two-dimensional one. The uniform sphere and the uniform cylinder are special
cases.
Let a be the radius of the body in the plane of symmetry, x the displacement
of the mass centre C,8 the angular displacement of the body, F the frictional
force and N the normal reaction (see Fig. 11.1.6). The equations of motion are
x - a8 = 0 or x - aiJ = 0
which gives mg sin a - F = F a2 /k 2 ,
k2
or F = -2--2 mg sin a,
a +k
.. a2
and so X =a8 = ---gsin a.
a2 +k2
Figure 11.1.6 Illustration for Example 11.1.4: circular body rolling and slip-
ping on an inclined plane.
11.1 ] Equations of Motion 305
Note that, by taking moments about the instantaneous centre, we can obtain
directly the relation
The right-hand side here is the rate of change of relative angular momentum
because the acceleration of the instantaneous centre is towards C (ao parallel to
rc in the notation of equations (11.1.9)). To see this we can calculate the
acceleration of the instantaneous centre as the resultant of the acceleration of C,
which is x down the plane, and the acceleration relative to C, which has the
components aO up the plane and a0 2 towards C. The resultant has components
x - aO = 0 down the plane and a(;2 towards C. *
If Il is the coefficient of friction we must have F'::;; pN. If there is no
slipping this gives k 2 tan IX .::;; ll(a2 + k 2 ). This is a necessary condition to be
satisfied if there is to be no slipping, but if the motion is to start as a rolling
motion it is also necessary that the initial velocity Vo of the point of contact
between the body and the plane be zero. The condition for this is Vo = Xo
- aBo = 0, where Xo and Bo are the initial values of x and O. In this case, if the
condition k 2 tan IX .::;; ll(a 2 + k 2 ) is also satisfied, the body will continue to roll.
If, however, Vo is positive, the body will start by slipping down the plane. The
frictional force F will be limiting and will act up the plane. Conversely if Vo is
negative F will be limiting and will act down the plane. To take the former case,
Vo > 0, the kinematical condition of no slip is replaced by F = JJN = 1J.mg cos IX as
the fourth relation. This gives
x - aO = Vo +g { sin IX - ( I+ ~: ) Il cos IX } t.
Unless k 2 tan IX .::;; ll(a 2 + k 2 ), x - ad will be positive for all positive t, which
verifies our previous conclusion that the inequality is a necessary condition for
rolling. If k 2 tan IX =ll(a 2 + k 2 ), rolling is possible only if Vo =0, but if
k 2 tan IX < ll(a 2 + k 2 ) there is a value of t;> 0 for which x - ad is zero. The
motion for later times is one of rolling.
If Vo < 0, then F = -1J.mg cos IX and
x- aO = Vo + g { sin IX + (1 + ~: ) Il cos IX } t.
* Note that when taking moments about the instantaneous centre, the unrestricted
relation (11.1.11) will in general be required. The restricted relation (11.1.9) is valid here
only because of the special properties of the kinematics associated with a rolling body of
circular cross-section.
306 Rigid Bodies [ Il.I
Example 11.1.5
In this example we investigate the dynamics of a car moving in a straight path
without slipping. The mass of the car body, excluding the wheels and axles, is m
and its mass centre C is at height h above the ground and a horizontal distance d
from the rear axle. The distance apart of the front and rear axles is I, and the
radius of the car wheels is a. The mass centre of the front wheels and axle is C1 ,
the mass is m I and the radius of gyration of the system about the axle is kl . The
corresponding quantities for the rear wheels and axle are C2 , m2, k 2 . respec-
tively. It is assumed that the car has a vertical plane of symmetry defined by the
plane containing C, C I and C2 .
We consider first the forces involved and we distinguish between those which
are external to the car as a whole and those which are external to the wheels and
axles. For the car as a whole there are the gravitational forces mg, mIg and m2g
acting vertically through C, C I and C 2 in the plane of symmetry. There are also
the reactions at the points of contact of the wheels with the ground. For either
pair of wheels these reactions can be combined into a single equivalent resultant
in the plane of symmetry. Let the resultants have components in the horizontal
and vertical directions equal to F I , NI at the front and F2 , N2 at the rear, as
shown in Figure 11.1.7. It is assumed that the car is being driven forward by the
rear wheels, which means that the frictio l1 ;J;1 force F2 must be in the direction of
motion to prevent the slip that would otherwise occur if the rear wheels rotated
~mg
c. _x
1
t h
t
0 2 - F2 1';-01
N2i:::=======__d__-_-_-_~~--··__________~.~iNI
Figure 11.1.7 Illustration for Example 11.1.5: the dynamics of a car.
11.11 Equations of Motion 307
with the car at rest. In fact F2 is the force which moves the car forward.
Conversely the front wheels are pulled along by the car and so F1 opposes the
motion of the car to prevent slip.
Internal to the car as a whole, but external to the rear wheels and axle, is the
torque or couple assumed to be applied to the rear wheels by the engine. This is
denoted by T. There are also reactions where the car body is mounted on the
axles. These are assumed to have no moment about C1 or C2 •
Let the horizontal displacement of C be x, and let the angular displacement
of the wheels be (). The kinematical condition of no slip gives x - a(} = O. The
equations of motion for the car as a whole are
(11.1.19)
(11.1.20)
The moment equation for the whole system requires some care. It is possible to
define a mass centre for the system as a whole, but this would not be a
convenient procedure here since we are dealing with three systems. These are: (i)
the front wheels and axle; (ii) the rear wheels and axle; and (iii) the car body;
they do not move as a single rigid body because the wheels can rotate relative
to the rest of the car. This is one of the exceptional situations where it is
preferable to choose a convenient reference point and apply the unrestricted
result in its more general form (11.1.12). We choose to take moments about the
point in the plane of symmetry which is equivalent in the two-dimensional
problem to the point of contact of the rear wheels with the ground, denoted by
O2 in Figure 11.1.7. The advantage of this is that Fl ,F2 ,N2 have no moments
about O2 and the resulting relation can be used to determine Nt. The car body
has no angular momentum about its mass centre C, so that dLe/dt is zero and
the only contribution to the moment is through the term fe xdp/dt. Since the
linear momentum of the body is mX horizontally, and since the height of C
above O2 is h, this contribution is hmi (the vector character of the moment is
suppressed since the direction of all moments is perpendicular to the plane of
the problem). For the front wheels and axle, the relative angular momentum
e,
about C1 is mkj so that dLe/dt for this component is of magnitude mkjO.
The linear momentum is mIX horizontally and contributes a term amlx to the
moment about O2 , Similarly the contribution from the rear wheels and axle is
m2 k~ (j + am2X. The resulting equation is
For the wheels and axle systems, moments about C1 and C2 for each system
separately give
(11.1.22)
308 Rigid Bodies [ 11.2
Note that, by assumption, the reactions between the car body and the axles do
not contribute.
We now have, from (11.1.19) and (11.1.22),
or
This equation serves to give the linear acceleration directly in terms of the
torque of the engine. When x is known we can obtain Nt from (11.1.21), which
gives
Here the signs of Fl and F2 will change but the expression for F2 - F J is
unchanged. Since x depends only on F2 - F t ,it follows that the expressions for
Nt and N2 are unchanged. Now slipping begins at the front wheels when
I Fl I = JJ.Nl and at the rear wheels when I F21 = JJ.N2, and so it depends on the
magnitudes of the normal reactions.
There are two factors affecting the magnitudes of Nl and N 2 . The
first is the position of the mass centre C. In a front-engined car C will
tend to be nearer the front, and so d> /- d. This factor will tend to in-
crease Nl relative to N2 and improve the adhesion of the front wheels.
On the other hand acceleration has the opposite effect, since it increases
N2 and decreases Nt. When the acceleration is substantial it will tend to be the
dominant effect so that the rear wheels are able to sustain a larger frictional
force than the front wheels before slipping. This is a reason in favour of using
the rear wheels to drive the car. Conversely the act of braking when x < 0 tends
to increase Nt relative to N2 and more reliable braking is obtained through the
front wheels.
We shall discuss equations (11.2.1) only in the limit of idealised impulses when
the time interval tends to zero but
~r x JF dt = ~r x fma dt
and equations (I 1.2.1) become, in terms of F,
* Remember that any force which does not become correspondingly large as the time
interval becomes small will not contribute to the system of impulses.
310 Rigid Bodies [ 11.2
(11.2.5)
Example 11.2.1
Suppose a rigid body suffers an impulsive blow F in a plane of symmetry, so that
the motion is two dimensional. We show that although the body is instan-
taneously set in motion, there can be points for which the induced velocity is
zero.
Let F act along a line whose perpendicular distance from the mass centre C is
x. Let the instantaneous angular speed of the body be w_and let the instan-
taneous speed of C be Vc (see Fig. 11.2.1). We have
F= mvc , Fx = mk2 w.
If P is a point on the perpendicular, through C, to the line of F and if PC = y, the
instantaneous speed of P is Vc - yw. This is zero if y = vclw = k 2 Ix. For
example a uniform rod, of length I, will begin to turn about one end if given an
impulse perpendicular to its length at a distance 11 from that end. A uniform
sphere of radius a resting on a horizontal plane will begin to roll without slipping
if it receives an impulse at a height 7al5 above the plane. This is the correct
height for the cushion of a billiards table.
Example 11.2.2
Two uniform rods AO, OB, each of mass m, length 21 and radius of gyration
x
F
1
t
l L,,~o
tf
y
~A~____________~~__~8r ____~______________-=B
x~H~x
Figure 11.2.2 Illustration for Example 11.2.2: impulsive motion of two
jointed rods.
1/3 1 / 2 about an axis through its mass centre, are smoothly jointed together at 0
and are at rest in a straight line. An impulse F is applied at A perpendicular to
the line of the rods. We wish to determine the instantaneous motion of the two
rods.
The problem is two dimensional in the plane defmed by the initial line of the
rods and the line of F. There will be an internal impulsive reaction at the hinge
producing equal and opposite impulses on the rods, say with components X
along the rods and Y perpendicular to the rods in the directions shown in
Figure 11.2.2.
Since the impulses on the two rods parallel to AB, namely iX, are equal and
opposite, any velocity components acquired by the mass centres of the rods
parallel to AB must be equal and opposite. This is not consistent with the
constraint imposed by the hinge unless X = 0 and the rods move perpendicular
to AB initially. Let the velocity of 0 be Vo perpendicular to AB and let the
angular velocities of the rods be WI, W2 perpendicular to the plane of the
problem, as shown in Figure 11.2.2. These quantities are sufficient to determine
the motion of the rods.
The impulse equations can be written down separately for the mass centres of
the rods. We have
and, by taking moments about the mass centres of the rods, we get
nor a fixed point it is necessary to use the general result (11.2.5), which gives
Example 11.2.3
A uniform inelastic sphere of radius a rolls without slipping along a horizontal
plane with speed u. It strikes a kerb, of height a/S, which is at right angles to its
path. Show that the conditions for the sphere to tum without slipping until it
has surmounted the kerb are
Figure 11.2.3 Illustration for Example 11.2.3: rolling of a sphere over a kerb.
11.2] Impulse and Momentum 313
where sin 0 = 4/5, cos 0 = 3/5. If the sphere is not to slip initially, it is necessary
that F < pR, or
6u _ 4u <3/lu
7 5 5'
In the subsequent motion, we continue to use 0 for the angle made by the
radius vector to the kerb with the horizontal, and we denote by F,N the
tangential and radial components of the reaction at the kerb. The equations of
motion are
F - mg cos 0 = maO,
. ·2
N- mgsmO =-maO ,
2
. 0 - -4 ) = -i'oma 2 ( 0.2
mga ( sm - 36u
-- )
5 49a 2 •
In order that the sphere surmounts the kerb, (j2 must remain positive until
o = n/2, so that
36u 2 /49a 2 > 2g/7a, or 18u 2 > 7ga.
In order that there should be no slipping during this motion, it is necessary that
F<pN,andso
The most critical case occurs initially, since cos 0, 0, jj all decrease, and sin 0
increases. Substitution of the values
gives
90U 2 )
or 21</l ( 98- ga .
The inequality required for no slip during impact is clearly covered by this.
314 Rigid Bodies [ 11.3
It has already been noted that the system of vectors consisting of the external
forces F and the system of vectors consisting of the mass accelerations rna are
equivalent in the sense of the defmition introduced in § 10.3. Hence the work
done by these two systems in a rigid body displacement is the same (see § 10.6).
The work done by the mass acceleration system is the increase in kinetic energy
since
Thus we arrive again at the energy relation, namely that the work done by the
external forces is equal to the increase in kinetic energy. Note, however, the
slight change in the argument compared with that used in §8.2. Here there is no
mention of the internal forces. The equivalence of the two systems of vectors is
itself sufficient to ensure that they do the same work in any rigid body
displacement.
The result can be generalised to a system consisting of several particles and
rigid bodies by summation over the constituent parts. However, the calculation
of the work done by the forces must then include the contribution, if any, from
the forces of constraint. Such a contribution might arise, for example, from the
stretching of a string, or the sliding of two rough bodies in contact. On the other
hand there are a number of situations for which the constraints are workless. For
example the net work done by the reactions ±X between the two surfaces in
contact is
fX. dr,
where dr refers to the relative displacement. If the contact is smooth X has no
component in the direction of relative displacement, so that the work done is
zero. If the contact is rough, but there is no sliding, the relative velocity is zero
and so
JX . dr = f X . v dt =o.
Another example arises from the tensions at the two ends of a light inextensible
string, which are equal in magnitude. The displacement of the two ends is
composed of two components, along and perpendicular to the string. The
component perpendicular to the string is also perpendicular to the tension, and
therefore makes no contribution to the work integral. The component parallel to
the string will make equal and opposite contributions to the work integral from
the two ends and so these contributions cancel. This will be the case if the string
is free or if it passes over a smooth peg but not if the peg is rough (see Example
5.3.1 ).
11.3 ] Work and Energy 315
To calculate the kinetic energy of a rigid body, let 0 be a point moving with
the body and let (a) be the angular velocity. A point of the body with position
vector r relative to 0 has velocity (a) x r relative to O. The kinetic energy of the
motion relative to 0 is therefore ~ !m(oo x r)2 . With the help of (11.1.13), this
is seen to be equal to ! L . (a), since
with the help of the relation (a). (r x v) = v . ((a) x r). The most general expres-
sion for the kinetic energy of the motion relative to 0 is therefore, from
(11.1.14),
It simplifies to
(11.3.3)
where I is the moment of inertia about the axis defmed by (a). This is readily
demonstrated by taking instantaneous axes such that (a) = (w, 0, 0). It also
simplifies to
(11.3.4)
when the axes are principal axes, so that the products of inertia vanish.
In applications, the above results are most useful when 0 is a fixed point, so
that the relative kinetic energy is the total kinetic energy, or when 0 is the mass
centre. In the latter case the total kinetic energy T can be written, with the help
of (8.2.8), (11.1.5) and (11.3.3), as
(11.3.5)
316 Rigid Bodies [ 11.3
where Le is the relative angular momentum about the mass centre C and Ie is the
moment of inertia about an axis through C parallel to w.
Example 11.3.1
A uniform circular disc turns about a horizontal axis through its centre 0, the
motion being resisted by a frictional couple whose magnitude is constant and
equal to M when the disc rotates. A particle is attached to a point P on the rim
of the disc. The system is released from rest with OP horizontal and swings
through an angle (!1T + 8) before coming to rest (see Fig. 11.3.1).
Show that: (i) the disc will not move from rest again unless (hI' + 8 )tan 8 > 1;
(li) if this condition is satisfied, OP will not reach the vertical again unless
The loss in potential energy of the particle is mga cos 8. The work done by
the frictional couple M is -M(! 1T + 8). Hence
The particle will not move unless the moment of the force mg can overcome the
couple M, or mga sin e > M. Therefore we must have
If OP is to reach the vertical, the loss in potential energy must not be less
than the work done by the frictional couple, since the kinetic energy cannot be
negative. This means that
mga(1 - cos e) ~ Me
1T + 28
or cos e ...;;;-- .
11'+48
mg
Figure 11. 3.1 Illustration for Example 11.3.1.
11.3 ] Work and Energy 317
Example 11.3.2
A uniform solid sphere is slightly displaced from its position of unstable
equilibrium at the topmost point of a fIxed sphere and it then rolls on this
sphere (see Fig. 11.3.2). Show that, irrespective of the magnitude of the
coeffIcient of friction, slipping must occur before the angle 0 which the join of
the centres of the spheres makes with the vertical attains the value cos- l (l0/17).
I.et the radii of the spheres be al, a 2 • The centre of the rolling sphere
describes a circle of radius (a 1 + a2). Its acceleration therefore has radial and
transverse components -(al +a2)(P and (al +a2)8 respectively. The equations
of motion are
·2
mgcosO-N=m(al +a2)0 ,
mg sin 0 - F= meal +a2)8.
The velocity of the centre of the rolling sphere is (a 1 + a2)8. It is also equal to
al W, where W is the angular speed of the sphere, since the sphere is instan-
taneously rotating about the common point of contact. Hence w =(a 1 + a2)8 /al
and the energy equation gives
N = mg ( cos 0 - 7"
10 (1 - cos 0) ) = 7"
mg
(17 cos 0 - 10).
Accordingly N is zero when cos 0 = 10/17. Slipping must occur before this.
318 Rigid Bodies [ 11.4
11.4 OSCILLATIONS
Example 11.4.1
A rigid body, allowed to oscillate about a horizontal axis through 0, is called a
compound pendulum (see Fig. 11.4.1). Moments about the fixed axis give
where e is the angular displacement, k is the radius of gyration about the mass
centre C, and OC = 1. The integral of the above equation gives the energy
equation, namely
The equations are similar to those for the simple pendulum (see Example 6.5.1)
and can be integrated in similar fashion. For small oscillations we have
2 gl
n =---
k + 12 '
2
When M arises from solid friction, it is constant in magnitude (but its sign
changes with that of dO (dt). The analysis for small oscillations corresponds to
that of Example 6.1.3. For viscous damping M is proportional to de (dt. The
analysis for small oscillations corresponds to that of § 6.2. For quadratic damp-
ing M is proportional to dO/dt 1dO/dt I. The equation in this case is essentially
non-linear evert when the amplitude is small. It is discussed in Chapter 13.
Example 11.4.2
We consider the equilibrium of a cylinder resting on a fixed cylinder. The
generators of the two cylinders are horizontal, their perimeters are assumed to
have continuous curvatures, otherwise their cross-sections are arbitrary. The
contact is assumed to be sufficiently rough to prevent sliding.
We can treat this as a two-dimensional problem in the vertical plane through
the mass centre C of the movable cylinder. Let P denote the point in this plane
on the generator of contact, lnd let the common normal at P make an angle 1/1
with the upward vertical. The mass centre C of the first cylinder must be
vertically above P, say at a height h, for otherwise the moment of the gravi-
tational force through C about P will be non-zero and prevent equilibrium.
Let us consider a small displacement from such a position of equilibrium, in
which the point of contact moves to Q. Let P t , P2 be the points on the two
cylinders which started together at P in equilibrium. Let the normals at P t and Q
meet in Ot, and let the normals at P2 and Q meet in O 2 as shown in Figure
11.4.2. We use the notation
°2
Figure 11.4.2 Illustration for Example 11.4.2: equilibrium of one cylinder
resting on another.
320 Rigid Bodies [ 11.4
cos1/J>.!..+.!..,
h rl r2
(In the critical case of equality the increment in potential energy is of a higher
order and a more .accurate calculation is required.)
If equilibrium is stable the cylinder will oscillate when disturbed. The oscil-
lations can be described with the aid of the energy equation, for which we need
to calculate the kinetic energy. From (11.3.5) this is !mv~ + !Ie(d¢/dt)2, where
Ie == mk2(say) is the moment of inertia about an axis through C, and Ve = QC dcp/
dt since the cylinder is instantaneously rolling about Q. Now dct>/dt is small and
QC differs from h by a small quantity of the first order. It follows that the
kirletic energy is !m(k 2 + h2 )(dr/>/dt)2 correct to second order, and so the energy
equation takes the form
dct»2
( dt + n2 ct>2 = constant,
11.5] Three-Dimensional Problems 321
This is the standard form of the energy equation for simple harmonic motion.
If one of the cylinders is locally concave, the formulae are still applicable
with a change of sign for the appropriate radius. If one of the cylinders
degenerates into a plane, so that the appropriate radius tends to infinity, we have
1/1 =0,
n2 = 8 g 0·6518g
97T - 16 r r
If the cross-section of the cylinder is an ellipse, with major and minor semi-axes
a and b «a), it will rest in stable equilibrium on a horizontal plane with the
major axis horizontal. In this case we have
(a 2Ib) - b a2 - b 2 4g
Ha 2 +b 2 )+b 2 g=a 2 +5b 2 b·
Note that when the cylinder rests with its minor axis horizontal we have h = a,
r = b2 la. Hence r - h < 0 and equilibrium is consequently unstable.
dLe dp
M=~rxF=-+r x-
dt e dt
Example 11.5.1
A body rotates with an angular velocity (a) about a fixed point O. The forces on
the body have moment M about 0 and the angular momentum of the body is L.
11.5] Three-Dimensional Problems 323
dL/dt = L+wx L= M
with components
dL 1 d 2
L.M=L.-=--L
dt 2 dt
2a
~
o
Figure 11.5.1 Illustration for Example 1l.5.2: rotating rectangular plate.
are
Lc = Ii Wi i + 12 W2j + 13 W3 k
= tmw(ib 2 cos a - ja 2 sin a)
mwab
(11.5.1)
mwab ..
Lo = Lc + rc x p = Lc = (2 2)112 (bl -aJ). (11.5.2)
3 a +b
Let w' be the angular velocity immediately after 0 is fixed. The new angular
momentum relative to C is
and so*
Since the expressions calculated in (11.5.2) and (11.5.3) are equal, it follows
that w~ = 0 and
*The result can also be deduced from (11.1.14). With axes based on 0 parallel to the
original axes, we have
Example 11.5.3
A uniform circular disc is suspended from a point 0 on its circumference. It is
given an impulsive blow, perpendicular to its plane, at a point P of its cir-
cumference (see Fig. 11.5.2). We wish to fmd the instantaneous axis of rotation.
The principal axes at 0 are horizontal and vertical in the plane of the disc and
perpendicular to the plane of the disc. The principal moments of inertia are
where a is the inclination of the radius vector CP to the vertical. The moment of
The instantaneous axis of rotation is therefore in the plane of the disc. Let it
make an angle (J with OC so that
w =-w(i sin (J + j cos (J).
It is clear that (J remains small over a substantial range of 0:. In fact 0 ~ (J < rr/lS
for 0 ~ 0: ~ rr/2.
To fmd the impulsive reaction at 0, say X, we require the velocity of the
mass centre. This is
vc = w x (-aj) =Sm
4F (1 - cos o:)k.
Then from the equation for the motion of the mass centre, namely
Note that X is zero when cos 0: = -1 so that for this value of 0: the disc would
still rotate about the same axis if free from the constraint at O.
Example 11.5.4
We consider the motion of a spherically symmetric sphere rolling and slipping on
a horizontal plane. Let m, a, r denote respectively the mass, radius and position
vector of the mass centre of the sphere. Let -Fn denote the frictional force at
the point of contact of the sphere and plane, where n is a horizontal unit vector.
When the sphere slips we have F =prng and the velocity of slip will be in the
direction of n and so of the form vn (see Fig. 11.5 .3).
The calculation of the angular momentum is straightforward for the sphere
328 Rigid Bodies [ 11.5
--lin
Figure 11.5.3 Illustration for Example 11.5.4: a sphere rolling and slipping on
a horizontal plane.
because all sets of orthogonal axes through the mass centre are principal axes. If
the angular velocity of the sphere is w, the angular momentum about the mass
centre is
since Ii =12 =13 =I say, at the mass centre. Note that the symmetry properties
of the sphere allow a choice of principal axes which are fIxed in direction and
not rotating with the sphere.
The equations arising from the kinematical condition at the point of slip, the
motion of the mass centre and the motion relative to the mass centre are
dr
dt + W x (-ak) = vn,
-Jimgn = m d 2 r/dt 2 ,
ak x /lmgn = I dw/dt.
Note that the gravitational force is cancelled by the normal reaction, and neither
of these forces is included in the equations. Differentiation of the fIrst equation
and elimination of d'2 r/dt2 and dw/dt with the help of the second and third
equations gives
dv dn ma 2
- n + v - = -pgn - - Jig(k x n) x k
dt dt I
Ima
2
= -pgn - pgn,
where the fact that k and n are perpendicular has been used. Since n is a unit
11.5] Three-Dimensional Problems 329
vector, dn/dt is also perpendicular to n and so the above equation implies that
dn/dt = 0 or n is a constant vector. The equation then reduces to
from which it follows that v decreases uniformly to zero. Note that d 2 r/dt2 =
-pgn is a constant vector, so that
where ro, u are the values of r, dr/dt at t = O. The above expression for r
represents a parabolic path for the centre of the sphere.
The equation
shows that k . 00 and n . 00 are constant and that the horizontal component of 100
perpendicular to n is proportional to the time.
After slipping ceases, the equations become
dr
--a,ooxk=O
dt '
--Fn = m d 2 r/dt 2 ,
doo
I -=ak x Fn = -ak x m d 2 r/dt 2
dt
=_ 2{d,OO _ (k doo) k}
rna dt . dt '
Example 11.5.5
Two uniform rods OA I , OA2 , each of mass m and length 2a, are freely jointed
together at O. We shall derive the equations of motion when the system falls
freely under gravity.
Let Cl , C2 be the mass centres of OA I , OA2 respectively and let al = OC I ,
330 Rigid Bodies [ 11.5
A2
Figure 1l.5.4 Illustration for Example 11.5.5: two jointed rods falling under
gravity.
I d2
or R=-m- 2 (a1 -a2).
2 dt
The moment equation for OAt about Ct gives -at x R = dL/dt where L
depends on the angular velocity Wt say of OAt. Now the angular velocity for a
rod is not uniquely defined and we can in fact specify that Wt . at = O. Since
Wit x at = da 1 /dt, we then have
and so
1 d2 1 d( da 2 ) 1 d 2a2
-a2 x (-R)=-ma2 dt (a1 -a2)=-m-d
X -2 a2 X-
dt =-3ma2 x -2- ·
2 3 t dt
dk
and so we can write -w=kx- +sk (11.6.3)
dt '
332 Rigid Bodies [ 11.6
(11.6.4)
(11.6.5)
provided that 0 is either a fixed point or the mass centre. Equation (11.6.5) is
the general equation describing the motion of a spinning symmetrical body
about a fixed point. For more general motion it can be used to describe the
motion relative to the mass centre and it is then augmented by an equation for
the motion of the mass centre.
Note that the first two terms on the right-hand side of (11.6.5) have no
component parallel to k. In gyroscopic problems this is often true of M as well,
in which case (11.6.5) implies that s is constant and Simplifies to
(11.6.6)
o
Figure 11.6.1 Illustration for Example 11.6.1: the spinning top.
11.6] Gyroscopic Problems 333
Let k be a unit vector along OC and let k' be a unit vector in a fIxed
direction. We consider fIrst the conditions for which a solution is possible in
which the angle 8 between k and k' is constant and k rotates steadily about k'.
This is called steady precession and the constant angular speed of precession p is
defIned so that pk' is the angular velocity with which k rotates about k'. The
point on OC with position vector k then has a velocity
so that
It follows that a couple of constant magnitude (/3SP - Jp2 cos 8)sin 8 is re-
quired to maintain the steady precession. Its direction is normal to the plane of
k and k'. It is the direction of this couple which is surprising when fIrst
encountered. Suppose the couple is produced by a force Fk', say, acting at a
point ak on the axis of symmetry so that M =ak x (Fk'). The body then moves
perpendicular to the applied force rather than yielding to it, as might have been
expected. A familiar example of this effect is the fact that the rider of a bicycle
must lean over on one side in order to turn a corner. The axle of the front wheel
is then the axis of spin, and the gravitational couple acts about a fore and aft
horizontal axis. The effect of the couple is to give the wheel an additional
angular velocity of precession which has a vertical component and turns the
front wheel in the required direction (the rear wheel is of course not free to
rotate about a vertical axis relative to the frame).
The results are immediately applicable to the spinning top, as illustrated in
Figure 11.6.1, which has its mass centre C at a distance a along the axis of
symmetry from the fIxed point O. The fixed direction k' is then vertical, the
gravitational couple at 0 is ak x (-mgk/), and a steady precession about the
vertical is possible if
Apart from the degenerate case, when the axis of spin is vertical, the
condition is
(11.6.11)
This is a quadratic in p2 and so, for given values of S and 8, two values of p are
334 Rigid Bodies [ 11.6
, dZk dk
-mgakxk =Ikx- z +/3 sd- (11.6.l3)
dt t
has the solution k = k'. This solution refers to the case of a top spinning with
constant spin s about the vertical. To discuss the stability of this solution, we
assume that in a small perturbation of the above solution
k=k'+p,
This equation is best discussed with reference to axes rotating with constant
angular velocity wak', where Wa is a constant to be determined. Relative to such
a frame we have
since k' . p = o. The notation now being used is that introduced for rotating axes
in §9.4. Substitution of (11.6.15), (11.6.16) in (11.6.14) gives
(11.6.18)
11.6] Gyroscopic Problems 335
The nature of the solution of (11.6.18) depends on the sign of the coefficient
of p within the parentheses. If the coefficient is positive, say n 2 , then p
satisfies the equation
which represents elliptic harmonic motion relative to the frame rotating with
angular speed 1 3 s/21. It follows that any small initial disturbance will remain
small. On the other hand if the coefficient is negative, say _[12, the solution for
p has the form
where P1 and P2 are fixed vectors in the rotating frame. Since P2 will not be
zero for an arbitrary initial disturbance, this situation represents instability. In
the critical case, for which the coefficient of p in (11.6.18) is zero, we have
p = O. Hence p varies linearly with t and this case is also unstable. It follows that
the condition for instability is
Example 11.6.3
We now consider the motion of a uniform sphere rolling without slipping on the
surface of a second fixed rough sphere. The fixed sphere has centre 0, radius b.
The moving sphere has centre C, radius a and mass m. The reaction on it at the
point of contact with the fixed sphere is R. Unit vectors k' and k are defined in
the vertical direction and along the line of centres OC respectively (see
Fig. 11.6.2).
In this example k is not fixed in the moving sphere, so that the previous
analysis is not immediately applicable. However, it is clear that the position
R
kl b
o
V.
Figure 11.6.2 Illustration for Example 11.6.3: a sphere rolling and slipping on
a fixed sphere.
336 Rigid Bodies [ 11.6
a +bdk
---=wxk. (11.6.20)
a dt
This equation is the same as (11.6.1) apart from the factor (a + b )/a and the
analysis now follows similar lines. Analogous to (11.6.2) we have
a +b dk
--kx-=w-(w.k)k
a dt
a +b dk
and so (a) = - - k x - + sk (11.6.21)
a dt '
where I is the moment of inertia of the moving sphere about a diameter. The
component of the above equation parallel to k gives ds/dt = 0 and so s is
constant.
Equation (11.6.22) must now be augmented by the equation for the motion
of the mass centre, which is
d2 k
R - mgk' = mea + b) dt 2 • (11.6.23)
, a+b d2 k dk
or -mgak x k = (I + ma 2 )-a- k x dt 2 + Is dt . (11.6.24)
e
There are, for example, solutions for which the angle between k and k' is
constant and k rotates steadily about k' with constant angular velocity pk'.
From (11.6.11) and the appropriate substitutions the condition for this is,
a+b
(I + ma 2) - - p2 cos () - Isp + mga= O. (11.6.25)
a
(11.6.26)
The angular velocity Wa of the axes differs from We by a term -Oi and so
Wa = i(We sin A - 0) - jWe cos A sin e + kWe cos A cos (). (11.6.28)
N
N
s
s
(a) (b)
Figure 11.6.3 Illustration for Example 11.6.4: notation used in the discussion
of the gyrocompass.
338 Rigid Bodies [ 11.6
where I, I, 13 are the principal moments of inertia at the mass centre of the
flywheel. To maintain the axis in a horizontal plane, it is necessary to apply an
appropriate couple of the form .Mj through the bearings in which the axis
rotates. Now
.Mj = dL/dt = L+ wa x L
= -i/O - jIweO cos A. cos 8 + kI3s
+ iWe cos A. sin 8(Iwe cos Acos 8 - 138)
+j(we ~ A. - OXIwe cos A. cos 8 - 138), (11.6.31 )
and so there are two relations implied by the condition that the i and k
components of dL/dt are zero. The second of these implies that the spin 8 is
constant, and the first gives the equation
(11.6.33)
This is just the equation for the simple pendulum as discussed in Example 6.5 .1.
In particular,small oscillations are simple harmonic with a periodic time 2rr/n.
Since the axis of the flywheel oscillates about 8 = 0, such a device offers the
possibility of estimating true north on the earth's surface, as opposed to the
magnetic north registered by a magnetic compass. This principle is the basis of
the gyrocompass. Since we is small, s must be large in order to achieve a periodic
time which is not too small. If, by way of illustration, we take
13 cos 'A/I = 1
s =6000 rev min -1 =2rr x 100 rad S-1,
We = 1 rev(sidereal day)-1 = 2rr/86 164 rad S-1,
11.6) Gyroscopic Problems 339
21T (
I )112 =(861'64)1/2 =29·4 s.
13swe cos A.
Example 11.6.5
In this example we consider the motion of a uniform disc, of mass m and radius
a, which is rolling on a rough, horizontal plane. Let k', k,j, i, be unit vectors
such that k' is vertically upwards, k is along the axis of the disc, j is parallel to
k x k' and i = j x k (see Fig. 11.6.4). Let e be the inclination of k to the vertical
and let 4J be the angular displacement of the disc about the vertical. The angular
velocity tIla of the axes is
Wa = i~ sin e - j8 + k~ cos e. (11.6.34)
where I, I, 13* are the principal moments of inertia at C. The equations for the
motion of the mass centre, the motion relative to the mass centre and the
k'
ve - wxai=O, (11.6.39)
where R is the reaction at the point of contact of the disc and plane and ve is the
velocity of the mass centre.
Elimination of R from (11.6.38) using (11.6.37) gives
dVe)
•
-aIX
( I
mgk +m'dt =dL
dt' (11.6.40)
. (gk' +
-al x m d ( wx al.)) = dL
m dt dt . (11.6.41)
Substitution for wand L from (11.6.35) and (11.6.36) into (11.6.41) gives
This is the equation for the general motion of the disc. We consider a special case
by looking for a solution in which 8, ~ and s are constant. This case represents a
rolling motion with the plane of the disc at a constant angle to the vertical and
the mass centre C moving in a circle with constant angular velocity. For this
steady motion, the rate of change with time arises solely from the rotation of
the axes. Hence (11.6.42) simplifies to
(11.6.45)
Exercises 341
jmgax - ma 2Ux + kS) - ma2i x (Wa X js) = iI~ +jlx + kl3 s+ Wa X kl3 s.
(11.6.46)
The three components of (11.6.46) give, respectively,
I~ +I3sX=O, (11.6.47)
x
mgax - ma 2 + ma2s~ =IX - 13s~, (11.6.48)
-ma 2s=hs. (11.6.49)
EXERCISES
11.1 A rigid body has angular velocity 00 and a point 0 of the rigid body has
velocity vo' Show that the kinetic energy T(v o , (0) is such that
11.5 Two equal rough circular cylinders rest in contact with each other on a
rough inclined plane, and their axes are horizontal. The inclination of the
plane to the horizontal is a. The coefficient of friction at all contacts is
tan X.
The cylinders are prevented from moving down the plane by a force
applied to the lower cylinder along the line of centres, and this force is
gradually increased. Show that equilibrium is finally broken by the lower
cylinder slipping up the plane if a > i1T - X, and by slipping between the
two cylinders if a < i1T - X.
11.6 A uniform circular disc, of mass m, is suspended in a vertical plane by two
vertical strings attached at points A, B on the circumference. The points
A, B are on the same level and the arc AB subtends an angle 2a at the
centre of the disc.
One of the strings is cut; what is the initial tension in the other?
(Ans.: mg/O + 2 sin 2 a).)
11.7 Two equal uniform rods AO, OB, are smoothly hinged together at 0 and
the end A is smoothly hinged to a fixed point. The rods are held in a
vertical plane with A and B on the same level and the angle AOB a right
angle.
Show that when the rods are released the initial angular accelerations
are in the ratio 3 : 4.
Note that there are two possible initial configurations, with 0 either
above or below the level of AB. The answer is the same for each case.
11.8 A uniform rod is supported against a smooth fixed sphere by a horizontal
string fastened to its upper end and to the highest point of the sphere.
Show that if the string is cut, the reaction between the sphere and the rod
is changed instantaneously by a factor cos2 a/(l + 3 sin4 a), where a is
the inclination of the rod to the horizontal.
11.9 A flywheel rotates with an average angular speed w. It is acted upon by a
forcing torque T sin 2 wt and there is a constant torque! T opposing the
motion.
What is the least moment of inertia required to make the difference
between the greatest and least angular speeds less than w/l00?
(Ans.: 50T/w 2 .)
11.10 A rough uniform rod, of mass m and length 21, is freely pivoted at one
end and is initially supported at the other end in a horizontal position. A
particle, of mass am, is placed on the rod at the mid point. The
coefficient of friction between the particle and the rod is fJ.. Show that if
the support is removed:
(ti) the particle begins to slide when the rod is inclined to the horizontal
at an angle (J given by IJ. = (10 + 9a)tan (J.
11.11 Two rods, each of mass m and length l, are joined together to form a
single rod of length 2l. The combination rotates on a smooth horizontal
plane about one end which is fixed. The join is such that it cannot sustain
a tension greater than T. Show that the angular speed must be less than
(2T/3ml)1/2.
Describe the subsequent motion of the two rods if the linkage should
snap.
11.12 A straight beam, of mass mb, rests on two uniform circular cylinders,
each of mass me and radius a. The cylinders are at rest on a horizontal
plane with their axes parallel, and the length of the beam is perpendicular
to the axes. The mass centres of the beam and cylinders lie in one vertical
plane. The beam is pulled with a horizontal force F parallel to its length
acting through its mass centre. Show that if there is no slipping, the
acceleration of the beam is F(mb + ime) -1.
11.13 An axially symmetric yo-yo consists of two equal discs joined by a short
axle, of radius a. The radius of gyration of the yo-yo about its axis is k. A
light string is partly wound round the axle and held vertically by its free
end.
Show that if the free end of the string is stationary and the yo-yo is
ascending with angular speed w, it will climb a length 1 of the string
provided that w 2 > 2g1/(k2 + a2 ). Show that if this condition is not
satisfied it will still climb a length 1 of the string if the free end is given a
constant down ward acceleration not less than
where
To what height does the mass centre of the cylinder rise? What is the
final speed along the ground?
mk2(u +aw)2
2(a 2 + k2 )
where k is the radius of gyration about an axis through the mass centre.
346 Rigid Bodies
Show that the mass centre of the sphere (or cylinder) will return to its
initial position if u < k 2 wla.
11.29 A rigid body has radius of gyration k about an axis through the mass
centre C.
(i) Show that the frequency of small oscillation.~ about a parallel axis is
a maximum if the axis is at a distance k from C.
(ii) Show that the frequency of oscillation is the same for all parallel
axes at a distance cor k 2 Ie from C.
(iii) The points 0, C, 0' lie on a straight line in that order and such that
OC = c, CO' = k 2 Ie. Show that if a particle is placed on the line, the
frequency of oscillation of the complete system about an axis
through 0 is increased if it lies between 0 and 0', and decreased if it
lies outside this interval.
11.30 A uniform rod, of length 2b, is suspended by two vertical light strings,
each of length I, which are fixed to the ends of the rods and to two fixed
points a distance 2b apart.
Show that the energy equations for motions in which the strings
remain taut and: (i) each string moves through an angular distance e with
the rod moving in a vertical plane; (ii) the strings move through equal and
opposite angular distances e with the mass centre of the rod moving
vertically, are
(ii)
1 .
-10
(!b
2 2 * e) -
2(l + 2 sin 2 */2 sin4
2 • 2 e
e) + g(1 - cos e) = constant.
2 b - 1 sm
Deduce that the periods of small oscillations are: (i) 2rr(l/g) 112 ;
(ii) 2rr(I(3gi 12 •
11.31 A uniform rectangular lamina, of mass m and with vertices A, B, C, D, lies
in a vertical plane and is free to swing about the point A which is fixed. A
particle of mass ! m is fixed to the lamina at B. The system is released
from rest with AB horizontal and is next instantaneously at rest when AB
is vertical.. Show that BC = 2AB.
Show that the frequency of small oscillations about the equilibrium
position of the system is
(
13g )112
6.2 1/2 AB
11.32 A uniform solid hemisphere, of radius a, stands with its curved surface in
contact with a horizontal plane.
Show that the period of small oscillations about its equilibrium
position is 21r/n, where
2 120g
n =--
119a
k2
1J. ~ -2--2 tan Il,
k +a
where Il is the maximum angle which the plane joining the axes of the
cylinders makes with the vertical.
Show that the period of small oscillations about the eqUilibrium
position is then 21r/n, where
11.37 A uniform rectangular block has adjacent edges of length 20, 2b, 2c. It
rotates with angular speed W about an axis parallel to a principal diagonal
passing through a vertex not lying on the diagonal.
Show that the kinetic energy is
11.38 Two wheels, each of mass mw and radius a, are joined by an axle of mass
ma and length 1. Each wheel has moments of inertia equal to Ia about its
axis and I d about a diameter. The axle has moment of inertia I at its mass
centre. The system is axially symmetric.
The combination rolls on a horizontal plane and the angular speeds of
the wheels are w, w'.
Show that the kinetic energy is
11.39 A uniform circular cone, of semi-angle /3, vertex 0 and axis OC, rolls inside
a conical shell, of semi-angle (Q + /3), vertex 0 and axis OZ. The shell
rotates with angular velocity Ws about OZ. The axis OC of the cone
rotates with angular velocity We about OZ.
Exercises 349
sin a
Ws + (ws - we) --:--a k,
sm '"
11.40 A uniform pole is supported at its lower end, which is carried round in a
horizontal circle of radius a with constant angular speed w. Prove that it
can maintain a constant inclination a to the vertical provided that
where ko is its radius of gyration about its supported end, and c is the
distance of its mass centre from this end.
11.41 A uniform sphere rolls without slipping on an inclined plane. Show that
the path of the centre of the sphere is a parabola.
11.42 A spherically symmetric sphere rolls without slipping on a horizontal
plane, and the plane rotates about a fixed vertical axis with constant
angular speed wo. Show that the centre of the sphere describes a circle
with constant angular speed wo/O + ma2 II), where m is the mass, a is the
radius and I is the moment of inertia about an axis through the centre of
the sphere.
11.43 Two men support a uniform pole, of mass m and length 2/, in a horizontal
position. They wish to change ends without changing their positions by
throwing the pole into the air and catching it. Show that if the pole is to
remain horizontal throughout its flight, and if the magnitude of the
impulses applied by each man is to be a minimum, the required impulses
have components! m (11g11 6) 1 /2 vertically and !m(11g116)1 /2 horizontally.
11.44 A uniform square plate, of mass m and side 2a, is suspended from one
comerO. Let axes Oxyz be defined such that Ox is horizontal in the plane
of the plate and Oy is vertically downwards. Show that· these are principal
axes.
Let the plate be given an impulsive blow F at the corner with position
vector 21/2 a(i + j). Show that the angular velocity imparted to the plate
is
350 Rigid Bodies
11.45 A uniform rectangular lamina has sides of lengths 2a, 2b. It is given an
impulsive blow perpendicular to its plane at one of its corners. Show that
the lamina instantaneously rotates about a line parallel to a diagonal and
at a distance ab /3(a 2 + b 2 )1 /2 from it.
11.46 A uniform cube is spinning freely about a diagonal. Show that if one edge
is suddenly fixed, the kinetic energy is reduced by a factor 1/12.
11.47 The centre of a spherically symmetric sphere, of mass m and moment of
inertia I about a diameter, coincides with the origin. The sphere is free to
turn about the point ck. It is set in motion by an impulse Fi applied at
the point r = xoi + yoj + zok.
Show that the impulsive reaction on the sphere at the fixed point is
Xi, where
11.48 A body turns freely about its fixed mass centre C under no forces other
than the reaction at C. The principal moments of inertia at C are It, 12 ,
1 3 , where It > 12 > 1 3 . The body is set in motion so that the angular
momentum L and the kinetic energy T satisfy the relation L 2 = 212 T.
Obtain a solution of Euler's equations for the angular velocity and hence
show that the angular velocity tends to (0, L/I2' 0) as t --+ 00.
11.49 A body turns freely about its fixed mass centre C under no forces other
than the reaction at C. The principal moments of inertia at C are It, 12 ,
13 and the angular velocity has components (Wt , W2, W3) relative to the
principal axes.
Show that if Wi ~W3, W2 ~W3, then W3 is approximately constant and
that if the small variation in W3 is ignored both Wt and W2 satisfy the
equation x + vx = 0, where v = (13 - It )(h - 12 )w~/ItI2.
Hence show that a steady rotation about the third principal axis is
unstable unless (13 - Id(I3 .- 12 ) > o.
11.50 A body turns freely about its fixed mass centre C under no forces other
than the reaction at C. The principal moments of inertia at C are It ,12,
13 and the angular velocity has components (Wt , W2 , W3) relative to the
principal axes.
Show that if (I2 - II )(13 - II - 12 ) = 0 and 12 ;p II , then
(i) For a body with It = 12 = I show that the magnitUde of the angular
velocity is constant and that the direction precesses about the axis of
symmetry with a periodic time 2nI/(I3 - I)s, where s is the constant
spin about the axis of symmetry.
Exercises 351
where
where 1/>0 is the initial value of 1/>. Deduce that if 13 > 1 the direction of
the angular velocity tends to coincide with the direction of the axis of
symmetry. Show that during the motion a plane through these two
directions turns through a total angle 1 3 (13 - I)s/kl relative to axes fixed
in the body.
11.52 Two gyroscopes are fixed to a light axle, which is their common axis of
symmetry. The gyroscopes can turn freely and independently about the
axle and the whole system can turn freely about a fixed point 0 of the
axle.
Show that the gyroscopes each have constant spin about the axle.
Show that the motion of the axle would be the same if the whole system
moved as a rigid body with spin (/3S + l~s')/(/3 + I~) about the axle. Here
s, s' refer to the constant spins of the gyroscopes and 13 , I~ refer to their
moments of inertia about the axle.
11.53 A top, of mass m, spins about a fixed point 0 on its axis of symmetry.
Unit vectors k', k, j, i are defined so that k' is vertically upwards, k is
parallel to the axis of the top, j is parallel to k x k' and i = j x k. The angle
between k and k' is (}, and I/> is the angular displacement of the vertical
plane of k and k' relative to a fixed vertical plane. The principal moments
of inertia at 0 are I, I,I3.
Show that the angular velocity of the top and the angular momentum
relative to the fixed point are, respectively,
where Hand E are constants, and a is the distance of the mass centre
from the fixed point.
(i) A spinning top is released from an initial position in which its axis is
horizontal and at rest. Show that the axis will rotate about 0 until it
makes an angle {3 with the downward vertical such that
(ii) Given that the axis of the top passes through the vertical, and that
e e
= 0 when = 0, show that
.
y2 = -2mga
I_y(1)2 + 2ky _ y2),
where
y =1 - cos e, 1)2 = 0 21
mga'
Hence show that when I) is small, the axis of the top remains near the
vertical provided that k is small or negative.
11.54 The top of Exercise 11.53 is subjected to a couple Mk'. The couple is
adjusted to ensure that ~ = p is constant. Show that
let cp be the angular displacement of the vertical plane through the rod.
,p
Let 0 =n/4, iJ =0, =Wo when t =O. Show that
Z 3g
8z =W~ (2 - cosecz 0) + _(rI/Z - cos 0)
4 1
provided that the rod remains in contact with the floor. Show that the
minimum value of Wo below which the rod will maintain contact with
the floor during the subsequent motion is given by wg IIg =3(2 - 2 1 / 2 ).
11.56 A spherically sYmmetric sphere, of mass m, centre C, radius a and
moment of inertia I about a diameter, rolls without slipping inside a
hollow fixed sphere, of centre 0 and. radius b. Show that a steady
precession is possible in which OC is inclined at a constant angle a to the
downward vertical and rotates about the vertical with constant angular
speed p provided that
b-a
-(1+ maZ) _ _ p2 cos a+Isp + mga =0,
a
where s is the constant spin of the sphere about OC.
Show that the solution is consistent with the no slip condition
provided that
where we is the angular speed of the earth and l, l, 13 are the principal
moments of inertia at 0 of the flywheel.
11 .58 An axially symmetric circular disc, of radius a, is in motion with its edge
touching a smooth horizontal table. The axis of the disc is inclined at an
angle 0 to the vertical, the spin about the axis is s.
Show that in steady motion the two rates of precession p are given by
the roots of
where I, I, 13 are the principal moments of inertia of the disc at its centre.
Show that the particular solution 0 = n/2, s = 0, when the disc rotates
about the vertical, is unstable if ap2 ~ 4g.
11.59 An axially symmetric circular cone, of mass m and semi-angle a, rolls on a
354 Rigid Bodies
horizontal plane. The angular speed of the cone's axis about the vertical is
p. Show that the vertex of the cone tends to rise from the plane if
where d is the distance of the mass centre from the base of the cone, and
1,1,111 are the principal moments of inertia at the mass centre.
11.60 A spherically symmetric sphere, of mass m, radius a and moment of
inertia 1 about a diameter, rolls without slipping on the inner surface of a
fixed hollow circular cylinder of radius b.
Let 8 be the angular displacement of the plane through the axis of the
cylinder and the centre of the sphere, and let x be the displacement of
the centre of the sphere parallel to the axis.
(i) Assume that the axis of the cylinder is vertical. Show that iJ is
constant and that x satisfies the equation x+n 2 iJ 2 (x+xo)=0,
where Xo is a constant and n 2 = fl(I + ma 2 ).
(ii) Now consider the problem when the axis of the cylinder is hori-
zontal. Show that
d2 X• 2 • .. ma 2 g sin 8
- + n x=O 8+ = O.
d0 2 ' (I + ma 2 )(b - a)
12 Virtual Work and
Lagrange's Equations
LF :: Lma, Lr x F :: Lr x ma,
that the force of constraint may do work in such a displacement and thereby
become an applied force, even though it is workless in the constrained condition.
It is important to appreciate that the result is to be taken as it stands, and is
irrespective of what may happen if the system undergoes a real physical displace-
ment appropriate to the dynamics of the situation. In application it is important
that the displacement of the system can be arbitrary, subject to compatibility
with the workless constraints. For this reason the displacement is referred to as a
virtual displacement and the work of the applied forces is called the virtual
work.
We begin with a discussion of a system in equilibrium, for which the work of
the system of mass accelerations is clearly zero, and we have the following result.
Example 12.1.1
Three uniform rods AB, BC, CD, each of mass m and length 2/, are smoothly
hinged at Band C. They are suspended from A and a horizontal force F
is applied at D (see Fig. 12.1.1). It is required to find the equilibrium
configuration.
Let the inclinations of the rods to the downward vertical be denoted by () 1 ,
()2, () 3' We consider the work done by the forces in a small virtual displacement
in which () 1, ()2, ()3 change by small increments. The mass centre of AB is at a
depth 1 cos () 1 below A and so there is a contribution to the work done by
gravity of amount mg6 (J cos () 1 ) = -mglo () 1 sin () 1 arising from the displacement
of AB. We take the sum of the contributions from the displacement of each rod
together with the work of F to get
6 W = mg 6(l cos () d + mgo(21 cos () 1 + I cos ()2) + mgo(21 cos () 1 + 21 cos ()2
+ I cos ()3) + FO(21 sin ()l + 21 sin ()2 + 21 sin ()3)
= (2F cos () 1 - Smg sin () 1 )/o() 1 + (2F cos () 2 - 3mg sin () 2 )lo() 2
+ (2F cos () 3 - mg sin () 3 )/o() 3'
~21
I
o F
Figure 12.1.1 Illustration for Example 12.1.1: eqUilibrium of three jointed
rods.
358 Virtual Work and Lagrange's Equations [ 12.1
For equilibrium this must vanish for all displacements, which is only possible if
the coefficients of [j() 1, [j() 2, and [j() 3 are separately zero. Hence we have
2F 2F 2F
tan 8 1 =-5-' tan 8 2 =-3-' tan 8 3 = - .
mg mg mg
Note that because the result must be true for all displacements, we get from the
one calculation the inclination of all three rods. Furthermore, since the displace-
ments chosen are compatible with the constraints at A, B and C, the net work of
the reactions at A, B and C is zero, and there is no contribution to the work
equation.
Example 12.1.2
Four uniform rods AB, BC, CD, DA each of mass m and length 2/, are smoothly
hinged together at their ends to form a rhombus. The frame is maintained in a
rigid configuration by a light rod BD. The rods AB and AD rest on two smooth
pegs at the same level and at a distance 2d apart as in Figure 12.1.2(a).
We derive the condition for the stability of equilibrium of the symmetrical
position and we calculate the tension in the tie rod.
Let x be the height of A above the level of the pegs, let AB be inclined at an
angle a + e to the downward vertical, where BAD = 2a. We have
x tan(a + e) + x tan(a - e) = 2d
and hence
The centre of the rhombus is at a depth 21 cos a cos e - x below the level of
the pegs. The potential of the gravitational forces on the rods, measured from
the level of the pegs, is therefore
4mg
V= 4mg(x - 21 cos a cos e) = --:--2 (d cos 28 + d cos 2a - 41 sin a cos 2 a cos 8),
sm a
d2 V 4mg
d8 2 = - -'-2-
sm a
(4d cos 28 - 41 sin a cos 2 a cos 8).
o
D
C mg
(a) (b)
oW = -0 V - 10 (41 sin a)
= { 4mg(d cosec2 a - 21 sin a) - 41T cos a} oa.
For equilibrium this must be zero, and so
T =. mg ( ~ _ 2 sin3 a ) .
sm2 a cos a '1
Let us compare the above derivation with one working directly with the
equations for equilibrium. In this approach we can take advantage of the
symmetry and confine the discussion to one half of the framework. Because of
the symmetry the forces on either side of AC are mirror images; in particular,
the reactions Xa and Xc at A and C are horizontal, as shown in Figure
360 Virtual Work and Lagrange's Equations [ 12.2
12.1.2(b). Note that when the equilibrium of the right-hand half of the
framework is considered, Xa and Xc are external forces.
The relation ~F = 0, has the horizontal and vertical components
In order to fmd T, we now need to calculate Xa and Xc' These are obtained by
taking moments about D for the equilibrium of AD and CD separately. We have
T=NCOSCli.-~(
cos CIi.
I _ _d_ )
21 sin CIi. '
= mg ( ~_ 2 sin3 0: )
sin 2 CIi. cos 0: 1 .
Comparison of the two calculations shows how the virtual work approach avoids
the calculation of unknown reactions. If it is necessary to calculate these
reactions, the virtual work approach can also be used to do this. It merely
requires the relaxing of a constraint so that the desired reaction contributes to
the work equation. If this can be done in a way which involves only one
unknown reaction, the equation gives directly an expression for the reaction.
For example if it is required to calculate Xc, the constraint at C is relaxed and
the virtual work equation is derived for a displacement in which the rod CD
rotates through a small angle OCli. about D. We then have
for equilibrium. This gives Xc = !mg tan 0: which is the same as the result
obtained by taking moments about D for the equilibrium of CD.
by cartesian coordinates (x,y) or polar coordinates (r, 8). The latter description
has obvious advantages in central force problems and especially when the
particle moves in a circle, since its distance from the centre is fIxed and its
position can be defmed by a single angular coordinate 8. If we insist on using
cartesian coordinates (x, y), then we must also impose the constraint x 2 + y2 =
constant for circular motion. The common feature of all descriptions, whatever
coordinates are used, is that the number of coordinates less the number of
constraints is a constant, called the degrees of freedom. A particle moving on a
circular path has one degree of freedom. A particle free to move in three
dimensions has three degrees of freedom. A system of two particles moving
independently has six degrees of freedom, but if they move as a rigid body this is
reduced to fIve by the constraint that their distance apart is invariant. The
position in space of such a system can be described by three coordinates defIning
the position of one particle, and two angular coordinates defming the position
of the second particle relative to the fIrst. We then have fIve coordinates and
no constraints, giving the same number of degrees of freedom as three
coordinates for each particle together with one constraint. A rigid body
proper actually has six degrees of freedom, because in addition to the five for
the orientation of two of its points, the body can have a further angular
displacement about the line joining them.
These ideas can be extended to more complex systems of rigid bodies, either
moving independently or subject to certain constraints. Any set of quantities
which is used to specify the configuration of a system is called a set of
generalised coordinates and a knowledge of the generalised coordinates implies a
knowledge of the position of every particle of the system. When the generalised
coordinates are subject to constraints, it is not possible to vary all of them
independently without violating the constraints. But if the constraints are of the
form
(12.2.1)
where q 1, Q2, ••• ,qn are the generalised coordinates and t is the time, such a
relation can be used to eliminate one of the coordinates. If there are m
independent constraints, all of the form (12.2.1), it is possible in principle to
eliminate m of the coordinates in this way. Any configuration of the system can
then be specified in terms of the n - m remaining coordinates, the number of
coordinates now being equal to the degrees of freedom. There being no further
coordinate constraints, these n - m coordinates can be changed independ-
ently. When the configuration of a system can be defined by a set of generalised
coordinates which can be varied arbitrarily and independently without violating
any constraints, the system is called holonomic. This possibility does not always
exist, because relation (12.2.1) does not represent the only type of constraint
that can be encountered. For example, the condition of no slip when one body
rolls on another involves the derivatives of the coordinates, and such constraints
are of the form
362 Virtwll Work and Lagrange's Equations [12.2
f(q1> Q2, ••• , qno til> 42, ... , 4n,t) =O. (12.2.2)
Unless the expression can be integrated to the form of (12.2.1), the pOSSIbility
of eliminating one of the coordinates is not available and the use of more
coordinates than there are degrees of freedom is unavoidable. Systems involving
such constraints are called non-holonomic.
Example 12.2.1
Figure 12.2.1 represents a system consisting of two circular discs, each of radius
a, free to rotate at the opposite ends of a slender axle of length 21. The system is
free to roll on a horizontal plane with the axle horizontal and the discs vertical.
To specify the configuration of the system we first specify the position of the
" mid point C of the axle by cartesian coordinates (x, y) relative to horizontal axes
Ox, Oy. Relative to C the position of the axle is specified by the angular displace-
ment 8 o(the axle from the fixed direction Oy. Finally the displacements of the
. discs relative to the axle can be specified by the angles cp and", through which
they rotate about the axle.
The five quantities x, y, 8, cp, '" represent one possible choice of generalised
coordinates which will specify the configuration of the above system. If the discs
were allowed to slip on the plane, all these coordinates could be given indepen-
dent increments without violating any constraints, and so there would be five
degrees of freedom. If, however, the motion is confined to that which does not
involve slipping, certain constraints must be imposed to ensure that the velocity
at each point of contact is zero. One condition is that the velocity of C along the
line of the axle should be zero, so that there will be no slipping in the direction
of the axle. This gives
o x
In addition the velocity of each disc perpendicular to the axle should be zero at
the point of contact with the plane and this gives the two further conditions
These last two conditions are better discussed in the alternative forms
obtained by addition and subtraction of (12.2.4) and (12.2.5). There are in all
three constraints and five coordinates and so two degrees of freedom.
Equation (12.2.7) can be integrated to give
subject to suitable initial conditions. We now have the choice of eliminating one
of the variables, say e, from the system of generalised coordinates since it is seen
to be specified in terms of cp - 1/1. No such integration is possible with the
remaining pair of constraints, or any combination of them, and we cannot
reduce the coordinates to less than four with two non-integrable constraints.
If, however, we consider a motion with one degree of freedom by specifying a
°
further constraint that e = and the discs roll in a straight line, the constraints
simplify to
o = 0, y = 0, x - a~ = 0, x - a{J = 0.
These are all integrable and give
e = 0, y = 0, x - acp = 0, x - al/l = °
subject to suitable initial conditions. In this case one coordinate, say x, is
sufficient to specify the configuration and this is classed as holonomic since
there are no constraints on x.
The point has been made that there is considerable merit in a careful choice of
generalised coordinates, so that the configuration of a mechanical system is
specified in a way that avoids unnecessary complication. But if we wish to
discuss the dynamics of the system by direct application of Newton's laws of
motion, it is necessary to calculate the accelerations in terms of the chosen
coordinates. This is not always a straightforward matter. Even for the simple
364 Virtual Work and Lagrange's Equations [ 12.3
and the position of the bob depends explicitly on the time as well as on the
generalised coordinate II .
We write
(12.3.1)
(12.3.2)
12.3 ] Lagrange's Equations 365
Thus r, and its partial derivatives or/oqi and or/ot, are functions of the coordi-
nates and the time. The total derivative dr/dt, which takes into account the
variations of the coordinates qi with t as well as the explicit variation ofrwitht,
is a linear function of the generalised velocities (i; (i = 1, 2, ... , n), with coeffi-
cients which are functions of the qi and t. If it suits our purpose to do so, we can
look upon dr/dt as a function of the 2n + 1 variables q j, qi (i = 1,2, ... ,n) and
t, and operate on it as such. It should be stressed that this procedure is solely
mathematical manipulation of the functions involved, and is quite independent
of any mechanical considerations. On this basis, it follows from (12.3.2) that
o dr or
(12.3.3)
(12.3.4)
d or
(12.3.5)
On the assumption that r has continuous second order partial derivatives, so that
the order of differentiation is not significant, relations (12.3.4) and (12.3.5)
show that
(12.3.6)
(12.3.7)
say, where the summation is over all the forces which do work in the
366 Virtual Work and Lagrange's Equations! [ 12.3
or
Qj=~F '-0 (12.3.8)
qj
(12.3.9)
where the summation is over all particles of the system. It follows that
(12.3.10)
~m d2 r . ~ = i. ~m dr . ~ _ ~m dr . i. ~ . (12.3.11)
dt 2 oqj dt dt oqj dt dt oqj
(12.3.12)
where T is just the kinetic energy of the system. Again, with the help of
(12.3.6), we have
dr d or
~m-.--=~m-.--=-~!m-.-=-,
dr 0 dr 0 dr dr or (12.3.13)
dt dt oqj dt oqj dt oqj dt dt oqj
d or or
so that finally, Q.=---- (12.3.14)
! dt oQ.j oqj'
av
-ov=--oq·
aqi I'
av d aT
-=--
aT
hence (12.3.16)
Lagrange's equations in this form are particularly effective because all the
equations are deducible from a knowledge of T and V.
The above discussion is valid only for holonomic systems where all the
coordinates can be given independent displacements. This is necessary in order
to justify a relation such as (12.3.7), which is obtained on the basis of a virtual
displacement in which one coordinate varies independently of the rest. A
non-holonomic system of the type discussed in Example 12.2.1 can be brought
within the framework of the discussion by allowing all the coordinates involved
to be capable of independent displacements in spite of the constraints, which
will now be violated. In other words the constraints are not to be imposed
during the exercise of deriving Lagrange's equations. We therefore formulate the
equations of motion for a more general situation, but once they have been
formulated we can reimpose the constraints and integrate the equations of
motion subject to these added conditions. When this procedure is followed it
should be borne in mind that the forces brought into play by a constraint may
not be workless if the constraint is relaxed, even if this is so in a constrained
displacement. This would be the situation in Example 12.2.1, where the system
can be made holonomic by relaxing the no slip constraints. Then the contact
forces between the discs and the plane will contribute to the work even though
they are workless when the discs are rolling. It is further to be emphasised that
the work is calculated on the basis of the actual forces involved in the physical
situation when they are given a virtual displacement. If the conditions of the
problem require no slipping, it is the forces appropriate to such a situation that
are involved, even though the virtual displacement may violate the no slip
condition.
Example 12.3.1
The position of a particle in three dimensions can be specified by the length r of
the position vector, the inclination e of the position vector to a fixed axis (the
polar axis) and the angular displacement if> of the plane containing the position
368 Virtual Work and Lagrange's Equations [ 12.3
vector and the polar axis relative to a fIxed plane. The quantities (r,8,4J) are
called spherical polar coordinates. The components of the velocity in the
directions of increasing r, 8 and 4J are ;., riJ, r~ sin 8 respectively.
Let the force on the particle have components F r , Fe, F q, in the direction of
increasing r, 8, 4J respectively. The work of the force arising from virtual
displacements 8r, 88, 84J in the coordinates is Fr 8r, Fe (r88), Fq,(r84J sin 8).
These expressions are to be identifIed with Qr8r, Qe88, Qq,84J. Hence the
generalised forces are
daT aT (d 2· 2·
rFe = dt ae - a8 = m d/ r 8) - r 4J2 sin 8 cos 8 ,
)
. d aT aT d .
rFq, sm 8 = - - ; - - - = m -(r24J sin 2 8).
dt a4J a4J dt
These are the equations of motion for a particle in spherical polar coordinates.
A spherical pendulum consists of a particle free to swing at the end of a string
whose other extremity is fIxed, as in Figure 12.3.1. Spherical polar coordinates
are a natural choice for such a system because r is now constant and the first
mg
equation simplifies to
0= :/T2~sin2()).
The last equation verifies that the angular momentum about the vertical is
constant, since there is no moment of the forces about the vertical. Between the
pair of equations ~ can be eliminated to give a second order equation for () .
Note that in this problem the particle actually has two degrees of freedom,
but three equations of motion have been obtained by relaxing the constraint
that T is constant until the equations of motion have been formulated. We
therefore get more information at the cost of discussing a more complicated
situation. It is one of the strengths of Lagrange's theory that this choice is open.
Because the system with two degrees of freedom is still holonomic, we can
stipulate that T is constant from the beginning, and simplify the kinetic energy to
The effect of this simplification is to yield the equations in () and <p, but not in T.
But this is sufficient to determine the motion, and if the calculation of the
tension in the string is not of interest, the equation in T is superfluous.
Example 12.3.2
A ring 0, of mass ml, is free to slide along a fIXed horizontal smooth wire. An
inelastic string, of length I, is fixed to one point A of the wire, is threaded
through the ring and has a particle P, of mass m2, fastened to its other end.
-;
/-r A
._r
;/\.
p •
r()
Figure 12.3.2 Illustration for Example 12.3.2.
370 Virtual Work and Lagrange's Equations [ 12.3
If the string remains taut, the configuration of the above system is completely
specified by the three coordinates r, 8, fj), where r = OP, 8 is the inclination of
OP to the wire and cp is the angular dil\plac~ment from the vertical of the plane
through OP and the wire.
The velocity of the ring is ;- along the wire. The velocity of P consists of a
horizontal component f of the ring together with components relative to the ring
of magnitude ;-, riJ, rei> sin 8 in the directions of r, 8, rp increasing. The kinetic
energy is
If we neglect all frictional forces, only gravity will contribute to the work, and
so there is a potential energy given by
~ {m2r(re+ f sin 8)} - m2 {f2 sin 8(1 - cos 8) + reos OCre + f sin 0)
. g cos rp)
(cos 2 0 - 2k cos 0 + 1) (rrp2 +-.- = 0,
sm 8
where k = 1 + ml .
2m2
12.3] Lagrange's Equations 371
The equations are therefore compatible with the assumed form of solution if
r= g cot (J
so that the acceleration of the ring along the wire is approximately constant.
We shall discuss these equations when they are linearised on the basis of small
oscillations, so that
The linear equations describe coupled oscillations of the type discussed in §6.4.
We look for the normal modes of oscillation on the assumption that
We then have
This equation is quadratic in n 2 and so has two roots. The left-hand side is
negative when n 2 = gill or gl12 and is positive when n = 0 or n ~ 00. Hence there
are two positive roots for n 2 , one less than gill or gl12 and one greater. The
corresponding oscillations are such that 0 1 10 2 is positive and negative respec-
tively. When the two roots nl, n2 are known, the general solution can be written
as a linear combination of the two normal modes.
Example 12.3.4
The position of a rigid body which is free to turn about a fixed point 0 may be
specified by the three coordinates 0, 4>, tJ;, where 0 is the angle between an axis
OC fixed in the body and an axis fixed in space, 4> is the angular displacement of
OC about the fixed axis and tJ; is the angular displacement of the body about
oc. The quantities 0, 4>, tJ; , are called Euler's angles (see Fig. 12.3 .4).
If we consider unit vectors i, j, k with k along OC, j perpendicular to the
plane of OC and the fixed axis, and i = j x k, the angular velocity of the body is
o
Figure 12.3.4 Illustration for Example 12.3.4: Euler's angles.
V= mga cos e,
d . .
13 dt (1/1 + et> cos e) = o.
~ + ~ cos e = s,
where s is the constant spin about the axis of the top. The second equation gives
Example 12.3.5
For the system represented in Figure 12.2.1, let m be the total mass, let I be the
moment of inertia of the system about a vertical axis through the mass centre
and let Iw be the moment of inertia of either disc about the axle. It is assumed
that the discs are equal and symmetrical about the axle, and that C is the mass
centre of the system. We then have
mi = -(Fq, + FIjJ )cos e + (Gq, + GIjJ)sin e, my = (Fq, + FIjJ )sin e + (Gq, +GIjJ )cos e,
10 = -1(Fq, - FIjJ), Iw~ =aFq" Iwij; = aFIjJ.
These equations could have been written down directly, and the power of
Lagrange's method is not fully apparent in this example. It does illustrate,
however, that a system with non-integrable constraints can be dealt with if the
constraints are ignored until the equations of motion have been derived. The
equations are then available for the more general situation of slipping or rolling.
For rolling, the above equations are to be solved in conjunction with the
constraints given by (12.2.3), (12.2.6) and (12.2.7). Since (12.2.7) is an inte-
grable constraint it could have been used to eliminate e, say, from the above
discussion, but a little care is then required in the calculation of the generalised
forces in cp and 1/1. The neatness of the equations of motion and the ease of
calculation of the generalised forces when e is retained illustrate a situation in
which the extra degree of freedom actually simplifies the discussion.
The solution of the equations of motion, subject to the rolling constraints,
will not be discussed in detail. It may be verified that they are consistent with
the solution ~ = constant, 0 = constant. Equations (12.2.7) and (12.2.6) then
determine the constant values of I:i and the speed of the mass centre respectively.
12.4] The Lagrangian and Hamiltonian 375
L=T- V, (12.4.1)
(12.4.2)
This follows because V does not depend on the generalised velocities, so that
H =~ oL .
"" ;-:- qj - L.
i uqj
(12.4.4)
and (12.4.2) has been used. The variables Pi are called generalised momenta. At
this stage we change our point of view and choose to regard H as a function of
the qi, Pi, t, regarded as independent variables. Note that in order to evaluateH
it is necessary to eliminate the tii in (12.4.4) in favour of the Pi defined by
(12.4.6). When the differential
aH aH) aH
dH(qi,pj, t) =~ ( ;- dqi +;- dpi + -;- dt (12.4.7)
i uqj UPI ut
376 Virtual Work and Lagrange's Equations 12.41
(12.4.9)
since each coefficient of the differentials on the right-hand sides of (12.4.5) and
(12.4.7) must separately correspond. For i = 1,2, ... ,n, relations (12.4.9) form
a system of 2n first order equations which are equivalent, and can be used as an
alternative, to the n second order Lagrange's equations (12.4.2).
One of the useful properties of (12.4.9) is that if a coordinate qi does not
appear in the Hamiltonian, so that aH/aqi = 0, the corresponding moment';1m Pi
is a constant. The 2(n - ],.) Hamilton's equations for the other coordinates and
momenta then form a system of n - 1 degrees of freedom, and the coordinates
not appearing in H can be ignored. For this reason they are called ignorable
coordinates. Further, if H (or L) does not depend explicitly on the time, we have
that
T= ~~aijiliqj.
I I
~ aL.
'"-'-a·
~ aT .
qi= '"-'-a· qj=2
T
,
i qi i qi
and that
H= ~ aa~ qj - L =2T - (T - V) = T + V.
qi
12.4] The Lagrangian and Hamiltonian 377
Example 12.4.1
The successful exploitation of ignorable coordinates depends on the choice of an
appropriate coordinate system. The motion of a single particle in two dimen-
sions can be described, using cartesian coordinates, with the aid of the functions
. aH av. aH av
Px = - ax =- ax' Py =- ay =- ay .
and the problem is now reduced to one degree of freedom. The remaining
equation gives
. ..
Pr = mr =- a;:
=7 - V '()
aH mh2
r
but in this problem it may be more convenient to use the energy equation
1 mh2
H=-p Y2 +-+ Vi(r)=E
2m 2r2 .
Example 12.4.2
In Example 12.3.4 the motion of a spinning top is described in terms of Euler's
378 Virtual Work and Lagrange's Equations 12.4 ]
Here
Po = aL/aO =10,
Pq, =aL/a~ =I~ sin2 0 + 13 cos 0 (~+ ~ cos 0),
PIjJ= aL/a~ = I3(~ + ~ cos 0),
{) = Po/I,
• _ Pp - PI/! cos 8
cp - I sin2 () ,
~ + ~ cos () = PI/;/I3.
2 (p P cos O? p2
H= T+V=P
-.!l. + q, - IjJ + ~ + mga cos O.
21 21 sin 2 () 213
Both cp and I/J are ignorable coordinates in this Hamiltonian, implying constant
values of PI/> and PI/! which represent the constant angular momenta about the
vertical and about the axis of symmetry respectively. The reduced problem,
having one degree of freedom, is described by the Hamiltonian
2
H=PO + U«())
21
2
~~ + U(O) =tI0 2 + U«()) =E,
where E is constant.
EXERCISES
Ilr =1 - e cos 8,
a + 5b
pg 2.3 112
12.5 A tripod is composed of three equal uniform rods OA, OB, OC, smoothly
hinged at O. The ends A, B, C are joined by three equal strings forming an
eqUilateral triangle, and stand on a smooth horizontal plane. Show that
the tension in each string is (3 1/2 mg tan 8)/6, where m is the mass of
each rod and 8 is the inclination of the rods to the vertical.
12.6 A long, uniform, flexible chain, of mass me and length I, hangs over a
pulley consisting of an axially symmetric disc, of mass md and radius a,
which can turn freely about a horizontal axis through its centre. A
particle of mass mp is attached to the rim of the pulley. The particle is at
the lowest point of the pulley when the lengths of the chain hanging on
either side are equal. The angular displacement 8 of the disc and the
380 Virtual Work and Lagrange's Equations
potential energy V of the system are both measured from this position.
(ti) Show that if the particle is removed, and the system is displaced
slightly from its position of unstable equilibrium, the angular displace-
ment after a time t is given by
12.7 A smooth circular cylinder, of radius a, is fixed with its axis horizontal
and parallel to a smooth vertical wall. The axis is at a distance d from the
wall. A smooth uniform rod, of length 2/, rests on the cylinder with one
end on the wall, and in a plane perpendicular to the wall. Show that its
inclination (J to the horizontal is given by
d + a sin (J - I cos3 (J = o.
tancp=i +tan8.
12.9 A uniform sphere, of mass ms and radius a, lies at the bottom of a fixed,
rough, hemispherical bowl, of radius a(1 + a), where a > O. A particle, of
mass m p , is attached to the top of the sphere.
Show that if the sphere rolls without slipping to a position in which
the join of the centres of the sphere and the bowl is inclined at an angle (J
to the vertical, the increase in potential energy is
°
natural length 2l, is attached to the end A and to a fixed point B
vertically above and at a distance 2l from A.
Show that, if 0 is the angle BOC, stable equilibrium is possible when
o= 1T for 0: ~ 1,
0:
or sin!O = - - - for 0: > 1.
20: - 1
12.11 A particle of mass m moves in a plane. Its cartesian coordinates are (x, y)
and the cartesian components of the force on the particle are (Fx,Fy )'
Generalised coordinates qt, q2 are defined so that x=X(qt,q2) and
y =y(qt,q2).
Derive the generalised forces Q 1, Q2 and the kinetic energy in terms
of 4 t , 42 and the partial derivatives of x and y.
(i) For x = qi - q~, y = 2q 1 q2 show that Lagrange's equations give
12.12 A conservative dynamical system with two degrees of freedom has kinetic
and potential energies given by
Show that F(8)~ is constant for any motion of the system. Show that
if the steady motion with 0 = 0:, ~ = W is slightly disturbed without
altering the value of F(O)~,
+7
.. (2F'2
a~ + F - F "F'V") w2 ~ = 0
12.13 Calculate the kinetic and potential energies of a simple pendulum using in
turn: (i) the angular displacement; (ii) the horizontal displacement; (iii)
the vertical displacement as the generalised coordinate.
Hence obtain the forms of the equations of motion in these coordi-
nates.
12.14 A uniform, smooth rod AB, of mass m 1 , hangs from two fixed supports
C, D by light inextensible strings AC, BD each of length I. The rod is
horizontal and AB = CD. A bead, of mass m2, can slide freely on the rod
and the system moves in a vertical plane.
Let 8 be the inclination of the strings to the vertical, and let x be the
distance of the bead from a fixed point of the rod. Write down Lagrange's
equations and show that
e
where 8 = Q when = O.
12.15 A ring, of mass ml, slides on a smooth straight wire inclined at an angle Q
to the horizontal. A particle, of mass m2, is attached to the ring by a
string of length l.
Show that a motion is possible in which the string is inclined at a
constant angle Q to the vertical and the ring slides down the wire with
constant acceleration g sin Q.
Show that a small perturbation of this motion is simple harmonic with
frequency
12.16 A uniform rod AB, of mass m and length 2a, is attached at A to a vertical
wire. The rod is free to rotate about the end A, and A itself is constrained
to move down the wire and is subject to a vertical force mf. Use as
generalised coordinates the vertical displacement x of A, the angle 8
between the rod and the upward vertical and the angular displacement
cp of the vertical plane through the rod and the wire. Derive the Lagrangian
and the generalised forces and hence obtain Lagrange's equations. Show
that ~ sin 2 8 = p is constant and that
3 2)" 3.
( 1 - - sin 8 8 - - 8 2 sin 8 cos 8 -
p2 cos 8 3f
3 + - sin 8 = O.
4 4 sin 8 4a
3 ) . 3f
( 1 - - sin2 8 8 2 + p cosec 2 8 - - cos 8 = constant.
4 4a
horizontal plane initially at the points (0, i), (0,0), (0, -I) respectively
relative to a set of axes Oxy in the plane.
At time t = 0 forces Yj, Xi, -Yj begin to act on A, B, C respectively.
Use the displacement x of B and the angular displacement 8 of either
string (assumed to remain taut) as generalised coordinates. Hence show
that the generalised forces are
Qx=X, Qo=-2Ylsin8.
r
Hence show that
where V varies only with the relative distances between particles, that is
with
foralli,j.
Show that the transformation
.2 2g h2
8 = - cos 8 - 2 +c1,
a sin 0 + I/ma 2
where C2 is a constant.
12.20 Two discs, each of radius a, are joined by an axle to form an axially
symmetric system which rolls without slipping on a plane inclined at an
angle a to the horizontal. The mass of the system is m and its radius of
gyration about the axle is k. A pendulum is suspended from the mid
point of the axle and can swing freely in a plane perpendicular to the
axle. The mass of the pendulum is m p , the distance of its mass centre
from the point of support is 1 and its radius of gyration about an axis
through its mass centre parallel to the axle is k p . The motion is two
dimensional with each disc rolling down a line of greatest slope.
Use the displacement x of the axle and the angular displacement e of
the pendulum from the vertical as generalised coordinates. Write down
the kinetic energy and potential energy and hence obtain Lagrange's
equations.
(i) Show that a motion is possible in which the system rolls down the
plane with constant acceleration provided () is constant and given by
tan a mk2
tan () = ----::-- where" = 2 .
1 + " sec 2 a ' (m + mp)a
(ii) If a = 0 and e is small, show that the system oscillates with frequency
2rr/n, where
12.21 A uniform circular drum, of radius a and moment of inertia I about its
axis, is free to rotate about its axis, which is horizontal. A light inelastic
string is wound round the drum and has one end fixed to the drum. A
light spring, of modulus A. and natural length I, is fixed to the free end of
the string. A particle, of mass m, is attached to the spring. The system is
released from rest in which the string and spring are held in a vertical
position with the length of the spring equal to I.
Write down Lagrange's equations in terms of the angular displace-
ment of the drum and the extension of the spring. Hence show that the
acceleration of the particle is
where n 2 =A
- (1+ma-2 )
ml I
where
12.23 A double pendulum consists of two uniform rods OA, AB, each of mass
m and length l, smoothly hinged at A. The system is suspended from 0
and performs small oscillations in a vertical plane. Show that the fre-
quencies of the two normal modes are
Write down the kinetic energy and the potential energy of the system
in terms of ql, q2, q3 and their derivatives. Hence obtain Lagrange's
equations, and deduce that the frequencies of the three normal modes of
oscillation are equal.
12.25 A uniform rectangular plate rests on four equal springs at the corners.
Show that for small vibrations, in which each corner is displaced
vertically to first order, the three normal modes have frequencies in the
ratios 1 : 3 1 / 2 : 3 1 / 2 .
12.26 A uniform rectangUlar plate has sides of length 2a, 2b. It is suspended by
386 Virtual Work and Lagrange's Equations
two light vertical strings attached at adjacent corners of the plate. The
strings are of length 1 and their distance apart is 2a.
Find the characters of the four normal modes. For small oscillations
about the eqUilibrium position show that the frequencies are
( !)1/2
1 '
(K)1/2 (2g)1/2{(I+..!)
31 ' 1 4b
± (1+31+~)112}1/2
4b l6b 2
where the summation on the left-hand side is over all the impulses which
do work in the virtual displacement and the right-hand side represents the
change in kinetic energy of the system.
12.28 The motion of a mechanical system is defined by generalised coordinates
qi. If the motion undergoes an impulsive change, show that
The left-hand side represents the change in oTjoqi arising from the
impulses. The generalised impulse Qi is such that Qi oqi is the work of the
impulses in a small virtual displacement oq i.
12.29 Two uniform rods AB, BC, of masses ml, m2 and lengths 11 ,12 respec-
tively, are jointed at B and lie in one straight line on a horizontal table.
The system is given a horizontal impulsive blow F at A at right angles to
ABC.
Formulate the kinetic energy in terms of the initial velocities Yt, Y2,
Y3 of A, B, C respectively. Deduce the generalised impulses and hence the
velocities.
Show that the kinetic energy acquired by the system is
(4ml + 3m2)P
2ml(ml + m2) .
12.30 Two uniform rods AB, BC, each of mass m and length 1, are smoothly
jointed at B and lie on a horizontal table with AB perpendicular to BC. A
horizontal impulse F is applied at C in the direction AC.
Show that the kinetic energy produced when A is fixed is 15/16 times
the kinetic energy produced when A is free.
12.31 A dynamical system with one degree of freedom has the Hamiltonian
H = q2(!p2 + 1). Write down Hamilton's equations and show that q =
sech 21/2 q
t given that = 1, q = 0 at t = O.
Exercises 387
where (r, 8) are the polar coordinates of the particle and JJ.. is the constant
strength of the dipole. Derive the generalised momenta Pr, Po, write
down the Hamiltonian in terms of r, 8, Pr , Po and hence obtain
Hamilton's equations. Show that p~ = Q - 2JJ.. cos 8, where Q is a con-
stant, and hence that ,.2
+ Q/r2 = 2E, where E is the constant total energy.
The particle is projected from the point r = a, 8 = 0 at time t = 0 with
a speed v perpendicular to the radius vector. Show that r2 = a 2 +
(v 2 + 2JJ../a 2 )t 2 .
12.33 Find the Hamiltonian for the system whose Lagrangian is given by
where the Euler angles 8, cp, 1/1 are the generalised coordinates. Determine
the generalised momenta Po, P</>, PljI. Show that if L 1 , L 2 , L3 are
defined by
n (3F 3G 3F 3G)
[F G] = ~ - - - - -
, i= 1 3qi 3Pi 3Pi 3qi '
388 Virtual Work and Lagrange's Equations
where the summation is over all the degrees of freedom. Note that
[F, FJ = 0 and that [G, FJ = -[F, G].
Evaluate [L 1 +iL 2 ,L 1 -iL 2 1 and [L 1 ±iL 2 ,L 3 J. Hence show that
[L 2 ,L 3 J =L t , [L 3 ,Ltl =L 2 , [L t ,L 2 J =L 3 •
13 Non-Linear Problems
(13.1.1)
where !x 2 is the kinetic energy, V(x) is the potential energy, and E is the total
energy, per unit mass. This equation has already been discussed within the
context of non-linear oscillations in §6.5. We have also seen in Chapter 7 that
even a two-dimensional problem can sometimes be reduced to the form of
(13.1.1), as in (7.2.3). This possibility arises when there are two conservation
equations, thereby allowing one of the two dependent variables to be eliminated
in favour of the other. Equations of the form (13.1.1) are not difficult to
integrate although in complicated problems this may have to be done numeric-
ally. We summarise here some of the main qualitative features of problems
governed by (13.1.1).
Motion is physically possible in any interval for which E - Vex) is positive,
since (13.1.1) gives a real value for x in such intervals.
390 Non-Linear Problems [ 13.1
Example 13.1.1
With a little experience, qualitative information about the motion of a particle
subject to a conservative force is quickly deduced from a graph of the potential
energy. Figure 13.1.1 shows the variation of
is negative. Hence, although the particle may initially move in the positive x
direction, it will ultimately move in the negative x direction with ever
increasing speed.
(ii) V(a)~E~ V(-a).
Case (i) is still possible, but now there is a second possibility in which
the particle oscillates between the two largest roots of the relation
V(x)=E.
If E = V(a) there is equilibrium since V'(a) = O. The equilibrium is
stable since V"(a) = 2n2 > O. For a value of E slightly greater than V(a), the
particle will perform small oscillations with the approximate periodic time
21T(V"(a)y 1 / 2 = 2 1 / 2 1Tln.
The point x = -a is also a point of equilibrium but it is unstable, and if
the particle starts from rest at a small distance from x = -a, it will move
further away, at least initially.
(iii) E> V(-a).
Here the whole of the energy curve, as far as the value of x for which
Vex) = E, is available in the physical problem. This is the only value of x for
which the velocity is zero and, since the acceleration there is negative, the
13.11 Conservative Forces 391
V(x)
a x
(13.1.2)
for the spherical pendulum and showed that the pendulum executes elliptic
harmonic motion when the amplitude is small. For arbitrary amplitude the
problem is non-linear and has certain features in common with the central force
problems of Chapter 7.
cp¢
I
I
I
L__ .... __
I r
Figure 13.1.2 Illustration for Example 13.1.2: notation for the spherical
pendulum.
392 Non-Linear Problems [ 13.1
d2 r 11-
m dt2 =-T' (13.1.3)
where r is the horizontal component of 1. This has the form of a central force
problem, although T is not known a priori. The associated angular momentum is
constant, since
.!
dt
(r x i) =r x i =0 (13.1.4)
by (13.1.3). Hence
where IP is the angular displacement about the vertical and h is the constant
component of angular momentum per unit mass about the vertical. If the
horizontal projection of I sweeps out an area A in time t, it follows that
(13.1.5)
1 2 ·2 h2 _
'II 8 + 2/2 sin2 8 - gl cos 8 - E. (13.1.6)
The governing equation has now been reduced to a first order equation for 8
involving two constants of the motion h and E. It is mathematically similar to
equation (13.1.1), which governs one-dimensional motion under a conservative
force.
The solution of (13.1.6) can be expressed in terms of elliptic functions, from
which the details of the motion can be deduced. Some insight can be obtained
simply (rom a calculation of the apsidal angle, that is the angular displacement in
IP between the minimum and maximum values of 8. This is more readily
13.11 Conservative Forces 393
h2
(dO)2
2gZ3 sin2 0 dl{>
(E) (1 - cos2 0) - 2gZ3
= gZ + cos 0
h 2
(13.1.7)
Relation (13.1.8) follows from (13.1.7) with cos 0 = ±1. Relation (13.1.9)
follows from a comparison of the coefficients of cos 0, and implies that
(13.1.10)
from which it follows that Cl and C2 cannot both be negative. Hence Cl cannot
be negative, since Cl ;;;'C2' Furthermore C2 cannot be less than -Cl and we have
the following inequalities (see also Fig. 13.l.3):
(13.1.11)
1 c
For physically possible motion, 8 is confmed to the interval 8 1 .e;;; 8 .e;;; 8 2 , and
the motion is oscillatory between these limits. The apsidal angle q,a is then, from
(13.1.7),
q,a =( yl2
h 23
2g1)
f8
8J
2
sin 8{(C1 -
d8
cos 8Xcos 8 - C2 )(C3 + cos 8)}
112' (13.1.12)
to give
(13.1.13)
g )1/2
( (13.1.14)
w= Icos8 1 '
28 1 = 1
71:
1)08
-I
il ..
0'6
0'4
,0'01
0'4 0'6
Figure 13.1.4 Variation of the apsidal angle CPa with the minimum and maxi-
mum values 0 1 , 8 2 of 8 for the spherical pendulum.
(13.1.13) tends to become independent of I/J and gives a defmite value for
CPa, namely
_ 7T(C 3 + 1)112 7f
CPa - (1 + Cd(C3 + Cd l12 =(4 - 3 sin20d12 . (13.1.15)
This is the apsidal angle in the near circular orbit for which the departure
from 0 =8 1 is always small (see Fig. 13.1.5). Since the angular speed in this
orbit is approximately w, as given in (13.1.14), the period of the small
perturbation superimposed on the circular orbit is the time to traverse an
angle 2CPa with angular speed w, that is
7f
(13.1.16)
- -
----------
7
(
\
\
\
~
Figure 13.1. 5 Perturbed path of a conical pendulum. The full line represents
the orbit of the bob as a small departure from the dotted circular orbit.
those parts of the oscillation for which the pendulum is in the nearly
vertical position. There is a rapid change in if> of almost 1T in the vicinity of
the downward vertical, and little further change unless the swing of the
pendulum carries it near to the upward vertical. When this happens a small
increase in (J 2 can produce a substantial increase in if>a until the stage is
reached when the pendulum begins to rotate in a nearly vertical circle and
the limiting value 1T of if>a is approached.
Although many of the forces which occur naturally depend on position only,
such forces do not exhaust all the possibilities. A more general type of force is
one which depends also on the velocity, and perhaps on the time explicitly, so
r,
that it is of the form F(r, t). The discussion in this section is restricted to
forces of the form F(r, i), so that the time does not appear explicitly. For the
one-dimensional motion of a particle, the governing equation is then
say, where f is the force per unit mass. Equations in which the time does not
appear explicitly are called autonomous. They have the property that if x(t) is a
solution, so is x(t + to) and one of the constants of integration will always occur
in this fashion as a shift in the origin of time. There are special cases for which
(13.2.1) is integrable by simple analysis, but this is not always so. Numerical
integration of such an equation is straightforward, but gives only a particular
solution to a particular problem. We shall consider here what information of a
general character can be obtained about the properties of the solution of
(13.2.1).
It is often useful to write the equation in the alternative form
dv
v dx = f(X, v), (13.2.2)
where v = dx/dt. The plane in which (x, v) are used as cartesian coordinates is
called the phase plane. Solution curves in the phase plane are called integral
curves or trajectories. Note that whereas (13.2.1) is a second order equation,
(13.2.2) is of first order, so that the integral curves form a singly infinite family
depending on one parameter of integration. Of course when the solution of
(13.2.2) is known, we can identify v with dx/dt to get dx/dt = vor
I x dx
o vX
- ()=t+to .
Equation (13.2.2) does not always have a simple integral, but information of
a qualitative nature is readily obtainable from it. In particular, since dxjdt = v,
the integral curves are traversed in the direction of x increasing when v > 0 and x
decreasing for v < O. From equation (13.2.2) they have vertical tangents where
v =0 and horizontal tangents where f(x, v) =0, apart from exceptional cases
which are discussed below. More generally, (13.2.2) defmes the slope of the
integral curve at any point of the phase plane asf(x, v)jv. This means that if we
approximate the curve by its tangent, we can proceed a small distance to a
neighbouring point. By repeated application, this process can be made the basis
of a scheme of numerical integration, although it may then be desirable to return
to (13.2.1) rather than to use (13.2.2). One reason for this is that there is a
difficulty in the numerical integration of (13.2.2) when v = f(x, v) = 0, for then
the slope of the integral curve is not well defmed. A point of the phase plane
where this is so is called a singularity. In a physical problem it is a point where
both the velocity and acceleration (and hence the force) vanish and accordingly
it is a point of equilibrium. The neighbourhood of such a point requires special
investigation. Numerical integration will then complete the details of the solution
at non-singular points.
We shall consider only points of equilibrium, such as (xo, 0), in the neigh-
bourhood of which f(x, v) can be approximated by an expression of the form
(13.2.4)
*Phase plane analysis is often discussed with reference to the pair of first order equations
or, equivalently,
(I3.2.S)
To plot the integral curves in the phase plane one can take ¢ = t + to as the
parametric variable and write
x = a'eq<P/2 cos n¢, (13.2.8)
say, wh,~re a' = ae -q t o /2. Any particular integral curve is obtained by regarding
x and v as functions of ¢, and the different integral curves are obtained by
varying the single parameter a'. The resulting curves are spirals as shown in
°
Fig. 13.2.1(a). The equilibrium point in this case is called a focus.
Equations (I3.2.6) and (13.2.7) show that when q > the amplitudes of x
and v increase with t, and the direction of traverse of the spiral is away from the
focus. This is described as unstable equilibrium. Conversely the amplitudes of x
and v decrease when q < 0. The direction of traverse is towards the focus and
the situation is stable. This corresponds to the physical interpretation of stable
equilibrium in the sense that a particle whose motion is governed by equation
(13.2.4) would oscillate with decreasing amplitude if q < 0, and so there would
be a tendency towards the equilibrium position as is the case for damped
harmonic motion (see §6.2).
The integral curves are confocal ellipses and the equilibrium point is called a
centre. The force is conservative and the integral curves are the curves of
constant energy
400 Non-Linear Problems [13.2
(c) p + q2 /4 = 0, q =1= 0,
The integral curve for A = 0 is the straight line 2v = qx. All other curves pass
through the origin and are tangent there to the line 2v = qx. This is because the
origin is approached on all curves as qt -+ - 0 0 . Then the first term in equation
(13.2.13) for v is small compared with the second. If only the most significant
term is retained we have that 2v - qx as qt -+ - 0 0 .
The singular point is called a node and the integral curves are as shown in
Figure 13.2. 1(c). Equilibrium is stable for q < 0 and unstable for q > O.
(d) p+q2/4>0,p<0,
There are two integral curves which are straight lines, namely v = A2X (A = 0)
and v = AtX (B = 0). All other curves are tangent to one or other of these straight
lines at the origin. If q < 0, so that Al < 0, A2 < 0, I A2 I> 1Al I, the origin is
stable and is approached as t -+ 00. Hence
as t -+ 00, for all the curves except v = A2x(A =0) which is called the exceptional
direction.
13.2) Autonomous Forces 401
If q > 0, then Al > A2 > O. The origin is unstable and the vicinity of the
origin represents large negative values of t for which
(e)p=O.
When p = 0, equation (13.2.4) is equivalent to
dv/dx = q
x = Ae qt + B, v =Aqeqt (l3.2.18)
Example 13.2.1
As a simple example we consider the integral curves of the equation
(l3.2.20)
where a >0.
The force here is conservative and easily integrable, so that an explicit
formula is available for the integral curves. In Example 13.1.1 the qualitative
aspects of the motion have aheady been discussed from an analysis of the
potential energy. For comparison, the integral curves in the phase plane will now
be discussed on the basis of equation (13.2.20).
402 Non-Linear Problems [13.2
"
x x
(a) (b)
IJ
(c)
(e)
Figure 13.2.1 Classification of the singularities in the phase plane. The arrows
show stable equilibrium in (a), (c), (d), neutral equilibrium in (b) and (e) and
unstable equilibrium in (0.
unstable node q
unstable spiral
centre
stable spiral
stable node
The curves have horizontal tangents on x =±a and vertical tangents on v =o.
These lines separate the regions of the (x, v) plane where dv/dx> 0 from those
where dv/dx < O.
There are singular points at x = ±a, v = O. Near x =-a
vi v
~ /
<0 !!.!:...>o <0
~ / ~
I I Ef
,...--.;
I I 0 I I \-x
I x
/ ~ /
>0 ~<o
dx >0
/
I' (a) (b)
Example 13.2.2
The equation of motion for a pendulum oscillating in a vertical plane with
quadratic damping is of the form
(13.2.21)
which gives a centre. This is, however, a situation in which the linearised
approximation fails, from lack of sufficient accuracy, to give a true indication of
the singular point. Physical intuition suggests that the damping term gives
genuine stability with a decay in amplitUde, as for a spiral singularity. This is in
fact the case. It so happens that equation (13.2.21) is easily integrated, since it is
a linear differential equation in w 2 • The result is
w>O
(13.2.22)
13.2] Autonomous Forces 405
Figure 13.2.4 Illustration for Example 13.2.2: integral curves for pendulum
with quadratic damping.
where tan a = 2k and Cl ,C2 are constants of integration.* The integral curves, as
given by equations (13.2.22), are shown in Figure 13.2.4. The neighbourhood of
the spiral singularity at the origin represents damped oscillations as expected. The
existence of a spiral singularity at each even multiple of 7r implies that the
pendulum can make a number of complete revolutions before finally settling
down to damped oscillatory motion. The integral curves which pass through the
saddle point singularities at odd multiples of 7r represent oscillations which come
to rest in the unstable equilibrium position.
Example 13.2.3
For a pendulum with linear damping, the equation of motion takes the form
(13.2.23)
which again has singularities on the 8-axis when sin 8 =O. At 8 =7r we have
w(dw/d8) = n 2 (8 - 7r) - kw,
* On a particular integral curve these constants must be chosen so that the curve is
continuous at w =o.
406 Non-Linear Problems [ 13.3
Figure 13.2.5 Illustration for Example 13.2.3: integral curves for pendulum
with heavy linear damping.
so that (J ='IT is a saddle point. But for (J =0 the linearised approximation now
If n 2 - k2/4 > 0 there is a spiral singularity and if n 2 - k2/4 .;;;; 0 there is a nodal
singularitv. These cases correspond respectively to the damped harmonic oscil-
lations and the heavy damping aperiodic motion. The integral curves for the
oscillatory case are qualitatively similar to Figure 13.2.4, and Figure 13.2.5 shows
the main features when the motion is aperiodic.
The discussion of damped oscillations in § 6.2 shows that, although the solution
(6.2.4) of equation (6.2.1) tends to a simple harmonic oscillation in the limit of
no damping (k = 0), this limiting solution is not a satisfactory approximation
when the resisting force is small. The reason is that the amplitude of the
oscillation, ao e- kt /2, will eventually differ substantially from its initial value ao
for sufficiently large values of the time, no matter how small k might be
provided that it is not zero. This is a particular example of a situation in which a
small effect is nevertheless cumulative and so can produce substantial changes
over sufficiently long periods of time. We say that the approximation is not
uniformly valid as t ~ 00.
For damped oscillations governed by equation (6.2.1) we have the exact
solution (6.2.4), but there are many examples of oscillations which are governed
by equations of the type
(13.~.1)
13.3] The Method of Averaging 407
where the small positive constant e is included to indicate that the term ef(x, x)
is small compared with n 2 x, but is sufficiently complicated to prevent the
integration of the equation in simple terms. An approximate method of solution
is therefore to be desired, but it should be one which gives due consideration to
the form of the solution for large values of t. One such technique is the method
of averaging.
We start from the observation that, since the solution of(13.3.1), when e = 0,
is of the form x = a cos rp with a and ~ (= n) constant, there is something to be
said for retaining the same form of solution for the full equation provided
allowance is made for possible variations in a and ~, which will presumably be
slow when e is small. To take again the example of damped harmonic oscillations
as illustrated by (6.2.4), the solution does approximate to a simple harmonic
oscillation in any range of t for which kt is small. There is a change in amplitude,
which can be substantial, but it will take place slowly if k is small and this is the
property ofthe solution which is exploited in the method of averaging.
It is convenient to rewrite equation (13.3.1) as the pair of first order
differential equations
-a sin rp - a(~ - n )cos cf> = n~ f(a cos rp, -na sin rp). (13.3.7)
If equation (13.3.6) is multiplied by cos rp, equation (13.3.7) by sin rp and the
resulting equations subtracted, ~ is eliminated and the following equation for ti is
obtained:
a
Again, by elimination of from (13.3.6) and (13.3.7), we get
408 Non-Linear Problems [13.3
•
a{4> - n) =- -ne I{a cos 4>, -na sin 4»cos 4>. (13.3.9)
A.t this stage no approximations have been made, and equations (13.3.8) and
(13.3.9) are equivalent to the original equations (13.3.2) through the trans-
formation (13.3.3). Equation (13.3.8) verifies the previous conjecture that the
rate of change of a is small when e is small. In other words, although it is possible
for a to change substantially, the change takes place on a slow time scale com-
pared, say, with the changes in the phase 4> which take place on a time scale of order
n- 1 • Similarly equation (13.3.9) shows that~ - n is also small, of the order of e.
However, as they stand, equations (13.3.8) and (13.3.9) are not more easily
solved than the Original equations (13.3.2). They are simplified on the following
basis. A substantial change in a is produced only if d is on average non-zero. A
contribution to d which is both small and zero on average will not produce a
substantial change in a because the small integrated contributions will con-
tinually cancel out. As a simple illustration, consider the equation
a = -ea{1 + cos t) (13.3.10)
It is clear from the solution (13.3.11) that the substantial changes in a that can
arise, do so from the factor e-Et • This, in turn, comes from that part of d which is
non-zero on average, namely -ea. But the term -ea cos tin d contributes the
factor e -e sin t, which never differs substantially from unity for all t when e
is small. We say that
(13.3.12)
1 eo 211"
a=-2'
•
J211"
0
-ea{l +cost)dt=--
2.~
r (1 +cost)dt=-ea,
if the variation in a over a period is negligible. Hence d =-ea and the integral is
given by (13.3.12).
This is the basis of the method of averaging. It brings out clearly the
difficulty encountered in approximating to the solution of equations (13.3.2)
and, hence, the way to overcome it. To justify the procedure formally is,
however, difficult and will not be attempted here. We proceed by analogy and
approximate equations (13.3.8) and (13.3.9) by taking the average value of the
13.3 ] The Method of Averaging 409
right-hand sides. Since they are periodic in <P with period 21T, it is sufficient to
take the average over this interval of <p. The result is
E
Ii = - -2
1211" t(a cos <p, -na sin <p)sin <p d<p, (13.3.13)
n1T 0
• E
<p - n = - -2-
1211" t(a cos <p, -na sin <p)cos <p M. (13.3.14)
n1Ta 0
t= -v = na sin <p
which gives the equation for linear damping. Since the average value of sin 2 <p is
! and the average value of sin <p cos <p is 0, equations (13.3.13) and (13.3.14)
become
<p - f n dt.
In the present case this is just <p - nt since n has been assumed constant. The
following example discusses the technique when n has a prescribed variation.
Example 13.3.1
The method of averaging can be readily applied to the equation
(13.3.15)
410 Non-Linear Problems [13.3
or, equivalently,
(13.3.16)
where n is not constant, but is a given function of the time which varies slowly.
The equation occurs in the theory of wave propagation and was first studied
because of its application in quantum mechanics.
As in the general theory we put
and now, in the evaluation of Ii, it is necessary to take into account the variation
ofn with t. We have
.
a = --SIn
na. 2 A.
0/, (13.3.22)
n
. Ii
r/> - n= - - sin rt> cos r/>. (13.3.23)
n
Ii
-=--
ri
whence- a =(no- )112 , (13.3.24)
a 2n ao n
Example 13.3.2
We consider next the plane oscillations of a simple pendulum, of length I and
angular displacement 8, when there is a small damping term proportional to the
velocity and when the length of the pendulum is allowed to vary in a slow
prescribed manner.
The transverse acceleration is Iii + 2ie. In the same direction the components
of the resistance and of the gravitational force are, respectively, -kid and
-g sin 8 per unit mass. The transverse component of the equation of motion is
therefore
Let n 2 =gIl and let the amplitude of the oscillation be small, so that sin 8 can
be replaced by the first two terms of its expansion in powers of 8, namely
8 - 83 /6. Then equation (13.3.27) can be written as the pair of equations
0= w, (13.3.28)
(13.3.30)
2 •
~ - n = - na cos4 cp + 3 ~ sin cp cos cp - k sin cp cos cp. (13.3.31)
6 n
It is assumed that all the terms on the right-hand sides of equations (13.3.30)
and (13.3.31) are small and can be averaged. The only terms which are non zero
on average are those proportional to sin 2 cp and cos4 cp whose average values are,
respectively,! and i. The averaged equations are therefore
a 3n I
-=-- --k whence a =ao ( no
n )3/2 e -kt/2 , (13.3.32)
a 2n 2 '
r
and
2
whence cp = n(1 - a ) dt + CPo· (13.3.33)
o 16
412 Non-Linear Problems [ 13.3
(13.3.34)
a )-1 21T (
T=-;
2
21T ( 2
a
1-16 =-;- 1+ 16 )
Example 13.3.3
In the previous examples, the method of averaging has been demonstrated when
the governing equation is a small modification of the equation x + n 2 x =0 which
describes simple harmonic motion. The essence of the technique is to retain the
form of solution x = a cos c/> of the simple harmonic equation, and to allow for a
small modification of the equation by appropriate modifications in a and cp. It is
the calculation of these modifications which is simplified by the method of
averaging.
The aim of the present example is to demonstrate that the method of
averaging is of more general application, and that the essential property of the
reference equation is that it should have a periodic solution. The example is
based on the equation
x + V'(x) = 0, (13.3.35)
ti 2 + Vex) = a, (13.3.36)
say, where a is the constant energy. We know that this equation has a periodic
solution when a - Vex) is of the appropriate form (see §6.5). Let us assume
that the solution is periodic, with period 21T/n in t given by
21T 1
-; = 2112
J±(a _ dxV(X))1I2 ' (13.3.37)
13.3] The Method of Averaging 413
where the integral is to be taken over one complete oscillation. The sign of
(a - V(X»1/2 is positive over that part of the oscillation where x increases with t
and negative where x decreases. Equation (13.3.37) defmes n as a function of a.
In equation (13.3.36) a appears as a constant of integration, but in the
following analysis a, and therefore n, will be regarded as a variable. Now in
the application of the method of averaging it is an essential requirement that the
periodic functions which appear should have a ftxed period, as we shall see later.
We therefore regard the solution x(a,4» of (13.3.36) as a function of the
parameter a and a variable 4> which is normalised so that 4> changes by 21T in one
complete oscillation. It follows that x(a, 4» satisftes the equation
3X)2 (13.3.38)
!-n 2 ( 34> + V(x)=a
and x(a, 4» is the solution of (13.3.38), which may also be written in the form
!v 2 + V(x)=a. (13.3.42)
Equations (13.3.43) and (13.3.44) are now in a form suitable for averaging.*
Since the right-hand sides are small and periodic in r/> with period 21T, the
approximate averaged equations are
21T f21T
11= en ax
0 [(x, v) ar/> dr/>,
)
(13.3.45)
. en ax
fo
21T
r/> -n = - 21T [(x, v) aa dr/>,
with x and v given in terms of a and r/> by (13.3.41). The right-hand sides of
(13.3.45) are functions only of a. The first equation becomes uncoupled from
the second and can be solved to give a as a function of t. Then the second
equation gives ~ as a function of a and hence as a function of t, so that r/> can
also be evaluated as a function of t. Note that n varies slowly since it is a
function of a given by (13.3.37). In principle the first of equations (13.3.45) can
be solved without a knowledge of x and v as functions of a and r/>. With the help
of (13.3.42), it can be written
(13.3.46)
The integral in (13.3.46) is taken over one complete oscillation and the sign of
21 /2(a - V(x)) 1 /2 is determined by the sign of v or ax/ar/>, so that it is positive
when x is increasing and negative when x is decreasing. Note that a knowledge of
a as a function of t is sufficient to give the variation of the amplitude and the
approximate period 21T/n of the oscillation as functions of t.
The analysis can be taken a stage further when the term ef(x, v) arises from
linear damping, so that [= -v. If we define
* It is at this stage that the requirement that the period of oscillation be fixed is
necessary. In order to apply the method of averaging to (13.3.43) and (13.3.44), the right-
hand sides of the equations should be periodic functions of rp. If x is periodic with period P,
say, whereP is a function of a, we have x(a, rp) =x(a, rp + P) and ox(a, rp)/orp = ox(a, rp + P)/orp,
so that x and ax/arp are periodic. However
-
a x(a, rp)
a
=-x(a, dP a
rp+P) + - - x(a, rp+P)
aa aa. da arp
so that ax/aa is not periodic unless dP/da = 0 and the period is fixed. To take a simple
example, if x = sin arp, then ax/aa = rp cos a<l> and the amplitude increases linearly with rp.
13.4] Forced Non-Linear Oscillations 415
where the integral is taken over one complete period, relations (13.3.37) and
(13.3.46) show that for the case of linear damping
. €A ~d
a = - dA/da or dt log A = -e.
(13.4.1)
or, equivalently,
with n > 0, v, k;;;' 0, f;;;' 0, p > °all constant parameters. A term representing
416 Non-Linear Problems [ 13.4
linear resistance is also included so that the interplay between resistive and
non-linear forces, and their modification of the linear solution near resonance,
can be assessed.
The averaging technique, as described in the previous section, suggests that we
make the preliminary transformation x = a cos if>, v = -na sin if> and neglect those
small terms which are zero on average in the resulting equations for a and if>.
However, we shall concentrate here on the solutions of (13.4.2) representing
forced oscillations. When the linearised equation is used, it is known that the
frequency of the forced oscillations is determined by the frequency p of the
forcing term, and that the free oscillations ultimately decay because of resis-
tance. It is to be expected that the situation is qualitatively similar when there
are small non-linear effects, so that we should look for solutions where ~ approxi-
mates to p rather than to n. This suggests that the transformation
pd = (n 2 - p2)a cos if> sin if> + va 3 cos 3 if> sin if> - kpa sin 2 if> - f cos pt sin if>,
(13.4.5)
p(~ - p) = (n 2 - p2)COS 2 if> + va 2 cos 4 if> - kp sin if> cos if> - fa- l cos pt cos if>.
(13.4.6)
In order to Simplify equations (1.3.4.5) and (13.4.6), it is assumed that all the
terms on the right-hand sides are small. This means that the non-linear terms, the
resistive terms and the forcing terms should all be small. In addition the forcing
frequency p should approximate to the natural frequency n in order to ensure
that the first terms on the right-hand sides are small. This concentrates attention
on the oscillations near resonance, where a linear approximation is invalid. With
these assumptions, (13.4.5) shows that d is small and (13.4.6) verifies the earlier
13.4] Forced Non-Linear Oscillations 417
t/J =pt + X
(13.4.8)
Unlike the equations in the examples of the previous section, these equations are
coupled and cannot be solved independently. They are non-linear and do not
have simple solutions. Nevertheless they are much easier to handle than the
original equations (13.4.5), (13.4.6) and, most significantly, they are auton-
omous since t no longer appears explicitly.
We can, for example, examine the equilibrium solutions, for which,ti = X=O.
Physically these solutions represent oscillations for which the amplitude and
phase are constant on average. They are of particular interest because the
solutions corresponding to stable equilibrium are the non-linear equivalent of the
forced oscillations which remain in linear theory after the transient oscillations
have decayed.
We consider first the equilibrium solutions when there is no resistance. With
k = 0, the solutions of (13.4.7), (13.4.8) for,g = X= 0 are sin X = 0 and
(13.4.9)
provided we choose X=0 if P < n and X = 'IT if P > n (we do not consider
negative values of the amplitude a; such a change of sign is rather taken into
account by a change of the phase Xby 'IT). At resonance, however, the non-linear
contribution is crucial, for without it the amplitude is singular. Its introduction
has several modifying features. Not only is the amplitude fmite when p =n.
418 Non-Linear Problems [ 13.4
v= 0
a a
j/
'/
I
1
I/; X =n:
/ /
/ I
/ (h
p n
~ p
v<o v::> 0
where a =(41/31 v I)l!r, but it is not even a maximum there. The fact that
(13.4.9) is a cubic for a means that there are three possible values for a, at least
for some values of the parameters, so that equilibrium solutions at a given
frequency may not be unique. These features are clearly shown in Figure 13.4.1.
It appears that there are three values of a for a given value of p if P < n
°
and v < or if p > n and v > 0.
Of equal importance is the question of the stability of the equilibrium
solutions, for if they are unstable they will not in practice be observed. To
answer this question, we consider a perturbation of (a, X) from an equilibrium
solution (ao, Xo) to a neighbouring solution (ao + a I , Xo + Xl ), where ao ,Xo are
constant and ai, Xl are relatively small. With k = 0, equations (13.4.7), (13.4.8)
13.4] Forced Non-Linear Oscillations 419
linearise to
2pQI =-(-lrfxl' (13.4.10)
(13.4.11)
where A(a) is given by (13.4.9) and the fact that Xo is either 0 or 7T has been
used. One can now verify from § 13.2 that the singular point al = XI = 0 is a
centre or a saddle according as (-lrA'(ao) is positive or negative. Alternatively,
it follows directly from (13.4.10) and (13.4.11) that
(13.4.12)
with a similar equation for Xl' The solution of (13.4.12) is oscillatory, implying
r
neutral stability (centre), if (-1 A'(ao) is positive. If it is negative a will grow
exponentially in time leading to instability (saddle).
From (13.4.9), it is seen that A '(a) is just 2pdp/da and so has the same sign as
the slope da/dp of the response curve. It follows that there is instability if
r =O(X =0) and a decreases with p, or if r = I(X =7T) and a increases withp. The
unstable oscillations are shown dotted in Figure 13.4.1, the net result being that
there are either one or two neutrally stable oscillations for any given value of p.
Equation (13.4.9) even has a solution for the free oscillations when f= 0,
namely
(13.4.13)
The response curves for non zero values of f are all asymptotic to (13.4.13).
Although the method of solution described here is not strictly appropriate for
free oscillations, it is a simple matter to check the validity of (13.4.13) in-
dependently, since equation (13.4.1) with k = f= 0 is conservative and so can be
integrated.
It remains to discuss the further modifications of the small resistive term, so
far omitted. When d = X= 0, equations (13.4.7), (13.4.8) imply that
(13.4J4)
there is no equilibrium solution (other than a =0) without a forcing term, since
the resistance will ensure that any initial oscillation will decay to zero.
Finally the resistance changes the neutrally stable oscillations into stable
oscillations (the singular point becoming a stable spiral instead of a centre). The
effect of this is to ensure that any small perturbation from a stable equilibrium
solution will ultimately decay and give a defmite tendency to the equilibrium
oscillation.
The existence of cut-off frequencies at P1 and P2, as shown in Figure 13.4.2,
gives rise to jump phenomena, which are easily observed in practice and peculiar
to forced non-linear oscillations. Let the particle oscillate with a frequency
P > P1 and an amplitude corresponding to a point on the lower branch of the
response curve labelled 12. Suppose now that the frequency is decreased con-
tinuously so that the amplitude of the oscillation follows the response curve
towards Pl. At Pl no further continuous increase in amplitude is possible. The
oscillation becomes unsteady and finally settles down to an amplitude appropriate
to the higher stable branch of the response curve. A similar, and more violent,
change takes place if the oscillation initially corresponds to a point on the upper
branch of the response curve and the frequency is continuously increased towards
P2·
The preceding discussion has concentrated on the equilibriu~ oscillations in the
vicinity of the principal resonant frequency. For non-linear oscillations there is,
however, the possibility of resonant response at other frequencies. With equation
(13.4.1), for example, there is an important subharmonic resonance, at a
frequency p/3, when P is in the neighbourhood of 3n. Such solutions can be
discussed by means of an analysis similar to that described here, but in the case
of equation (13.4.1) the subharmonic response only emerges when the approx-
imation is taken to second order.
Exercises 421
EXERCISES
d
m - x = - -Ax - flmg sgn -
2
(dX)
dt 2 I dt '
Ax f = d~ .
Let ~=-,
flmgl dr
1 + 4k2
1 + exp{ 2(2r - 1)k7T} < - - 2 - w 2 < 1 + exp{ 2(2r + l)k7T}.
2n
a resistance equal to -ki I i I per unit mass and a force [(x) per unit
mass. Show that, while i remains positive,
Show that if [(x) is periodic, so is A (x) and with the same period.
Let [(x) be the periodic function defined by
x=nX,
and using the fact that A (x) is periodic with period X to calculate C.
Hence show that
{
~ (1 _ 2exp{ -2k(x - nx)}) ,
nX~x ~(n + t)X,
k I + exp(-kx)
A(x) =
_~ (1 _
k
2exp{ 2k(x - nX - t X)}) , (n + t )X
l+exp(-kx)
~ x ~ (n + 1)X.
Deduce that the velocity of the particle will eventually become zero.
What are the points of equilibrium, and which of them are stable?
(Ans.: x = nX, unstable; x =(n +! )X, stable.)
13.11 Verify, by direct substitution, that the solution of the equation
if kl = 0 and
where
This differential equation is called the Van der Pol equation. The
stable oscillation x = 2 cos nt is called a self-excited oscillation or a limit
Exercises 425
-m(~+ :5) r,
acting on a particle of mass m, the differential equation for trajectories of
the particle can be written in the form
d2 u /l vu 2
- +u- - =- u = r- 1 ,
d0 2 h2 h2 '
where n5 = gil and the energy constant is written as 2n5a 2 • This equation
represents oscillatory solutions with period 21rln, where
The integral is taken over a complete oscillation and the sign of the
denominator is determined by the sign of e.
Investigate the equations
00
o= O(a, ct», w=n-
act> '
~= n,
426 Non-Linear Problems
A Io
=
Bm
(a 2 - sin2 !0)1/2 dO = 2{E - (1 - a)K},
The integrals K and E are the complete elliptic integrals of the first
and second kinds.
Show that, approximately for a ~ 1,
om =0 0 e-e,t/2 (1 __192
5_ 02 (1 _ e-e,t»).
0
where the terms on the right-hand side are assumed to be small compared
with the individual terms on the left-hand side. Verify, by direct sub-
stitution, that the expression
sin a =kpa/f,
a(p2 _n2)=~va3 -fcosa,
e(9p2 _ n 2 ) =!va 2 .
Exercises 427
Verify that the leading terms agree with the solution obtained from the
method of averaging.
13.18 The point of support of a simple pendulum, of length I, has a small
vertical displacement €l sin pt (0 < € ~ 1). The small angular displace-
ment of the pendulum is (J. Show that the approximate equation of
motion is
x= ±11'/2, P = 2n(l - ha 2 ± €)
Deduce that the linearisation is not valid when 2(1 - €)n < p < 2(1 + e)n.
13.19 For the equations
dx dv
-=v -+n 2 x=€d(x,v)+€2 cbspt
dt ' dt
da dX
2p - = -k(a) - €2 sin X, 2pa - = -/.l(a) - €2 cos X,
dt dt
where
€1
k(a) = -
11'
f 1T
-1T
I(a cos cp, -pa sin cp)sin cp dcp,
428 Non-Linear Problems
€l
f..l(a) = -
7T
f-71
71
f(a cos 1/>, -pa sin I/»cos I/> dl/> - (n 2 - p2)a.
x = x(a, 1/»,
Use the method of averaging to show that there are resonant oscil-
lations when rl/> = pt + X, where X is a slowly varying function and r is an
integer. Hence show that the approximate equations for a and X are
X=O and Ar da
d
nO -
l'
lfAr) > °
d
or X=7T and Ar -
da
l'
nO + lfAr) < 0.
Index
When using approximations for damped oscillations, a small resisting force causes cumulative effects over long durations, rendering the approximation unsuitable as it diverges from actual behavior. The method of averaging resolves this by considering time-averaged effects, allowing bounded behavior and valid approximations for large time values .
Resonance in forced non-linear oscillations occurs when the forcing frequency closely matches the system's natural frequency, magnifying the response amplitude due to non-linear effects . This situation is complex because the response is not linear, and equations describing non-linear forced oscillations are autonomous, indicating that time doesn't explicitly appear, which allows for analyzing equilibrium solutions where average amplitude and phase are constant . Response curves are significant as they depict the relationship between the frequency of the external force and the amplitude of the oscillation . These curves can show multiple equilibrium points with differing stabilities, which are not possible in linear systems . The non-linear response is crucial at resonance, preventing the theoretically infinite amplitude seen in linear systems . The method of averaging, which simplifies the behavior over time by ignoring rapid oscillations, helps in understanding this complex interplay of forces near resonance .
The problem of free rolling motion on an inclined plane is described in terms of the sphere's path by considering the center of the sphere following a parabolic path. For a sphere rolling without slipping, when gravitational and normal forces counterbalance, the equations of motion simplify to show that the path traced by the mass center forms a parabola, characterized by a consistent decrease in velocity until it reaches zero . This is due to the constant vector acting in the motion equations, which leads to a uniformly decreasing velocity and a parabolic trajectory .
The stability of equilibrium solutions in forced, non-linear oscillations is contingent upon the sign of the slope of the response curve. Neutral stability (center) arises if (-1)^r A'(a₀) is positive, while instability (saddle) results if negative. These conditions delineate stable oscillation points where amplitude and phase remain constant, drawn from eigenvalue analysis of perturbations .
The derivative of a vector function with respect to a scalar variable like time can be related to its components by expressing the vector function in terms of its components along unit vectors i, j, k. If a vector r(t) is expressed as r(t) = x(t)i + y(t)j + z(t)k, then its derivative with respect to time, dr/dt, is obtained by differentiating each component function separately, resulting in dr/dt = dx/dt i + dy/dt j + dz/dt k. This implies that the derivative of the vector is the vector sum of the derivatives of its components multiplied by their respective unit vectors . This approach is analogous to the differentiation of a scalar function and retains the properties of the derivative, such as providing the instantaneous rate of change of the vector with respect to time .
If a vector function v(t) has an integral r satisfying dr/dt = v(t), then the indefinite integral of v is Jv(t) dt = r + c, where c is a constant vector of integration. This implies that a vector r can be constructed which differentiates to the given vector function v(t), signifying fundamental connections between differentiation and integration in vector calculus .
Linear damping adjusts the equation of motion of a pendulum to form w(dw/dθ) = -n²θ - kw, where the damping term kw influences the system's tendency towards equilibrium through energy dissipation, leading to spirals or nodal points in dynamic systems depending on specific parameters .
For the angular velocity of a reference frame with respect to another to be additive in the vector sense, certain conditions must be satisfied. The frames must be in relative rotatory motion about a common fixed origin, allowing angular velocities to be composed such that the resultant is also a vector quantity. This holds true in a rigid frame of reference, where angular velocities combine vectorially when considering relative rotations among multiple frames of reference . Additionally, the concept of adding angular velocities as vectors is valid if we treat them as small angular displacements where higher-order terms can be neglected . This ensures that the angular velocity vector _{ik} can be expressed as the sum of vectors _{ij} and _{jk}, which implies the angular velocities combine in such a way as to satisfy the vector addition rule in a rigid framework .
The combination of two springs alters the system's harmonic motion, depending on if they are in parallel or series. Parallel configuration leads to a combined stiffness, influencing the time period to be T = 2π{m/(k₁ + k₂)}^1/2. In series, the springs' combined compliance affects the time period differently, highlighting the elastic properties' role in modifying the natural frequency of oscillations .
The derivative of a position vector \( r(t) \) with respect to time \( t \) represents the velocity vector of a point \( P \) whose position varies with \( t \). Geometrically, this derivative is the tangent vector to the curve traced by \( P \) in space. As the time interval approaches zero, the direction of the unit vector along \( PP' \) (where \( P' \) is a point infinitesimally close to \( P \)) aligns with the tangent to the curve at \( P \). Thus, the derivative provides the direction and rate of change of position along the path, known as the velocity vector, and is tangent to the path at each point .