Vous êtes sur la page 1sur 63

PHYS 143 - Mechanics and Thermal Physics

M. Egblewogbe
August 12, 2010
Texts for Background Reading.
1. Hugh D Young and Roger A Freedman. Sears and Zemanskys University Physics: with Modern
Physics. 11
th
Edition. Pearson Education/ Addison Wesley (2004).
2. J S Gyakye Jackson. Mathematical Methods for Mathematics, Science and Engineering Students.
Vol 1. Ghana Universities Press (2002).
Contents
1 Vectors 4
1.1 Properties of Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.1.1 Geometrical Representation of Vectors . . . . . . . . . . . . . . . . . . . . . . 5
1.1.2 Multiplication of Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.1.3 The Dot Product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.1.4 The Cross Product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.1.5 The three dimensional Cartesian co-ordinate system. . . . . . . . . . . . . . . . 6
1.1.6 Components of a vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.1.7 Magnitude of a vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.1.8 Components of a vector and the dot product . . . . . . . . . . . . . . . . . . . . 8
1.2 Geometrical methods of vector addition . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.1 The sine rule and the cosine rule . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.2 Vectors in two dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2 Linear Momentum 10
2.1 Conservation of Linear Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2 Collisions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3 Motion 12
3.1 Newtons Laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.2 Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.2.1 Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.2.2 Motion in one dimension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2.3 Parametric Equations of motion . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2.4 Motion in two dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2.5 Acceleration due to gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2.6 Projectile Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
1
4 Force 25
4.1 Addition of forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.2 Impulse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.3 Tension and the motion of connected masses . . . . . . . . . . . . . . . . . . . . . . . . 26
4.4 Friction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
4.5 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
5 Circular Motion 35
5.1 The basic equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
5.2 Uniform Circular Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
5.3 The Conical Pendulum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
5.4 Non-uniformcircular motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
5.5 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
6 Work and Energy 40
6.1 Work done by a varying force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
6.2 Work done by a constant force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
6.3 Work and Kinetic Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
6.4 Work and Potential Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
6.5 The Principle of Conservation of Energy . . . . . . . . . . . . . . . . . . . . . . . . . . 42
6.6 Conservative and non-conservative forces . . . . . . . . . . . . . . . . . . . . . . . . . 42
6.7 Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
7 Some aspects of Newtonian Gravitation 44
7.1 Keplers Laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
7.2 The Law of Universal Gravitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
7.3 Gravitational Potential Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
7.4 Escape Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
7.5 Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
8 Thermal Physics 49
8.1 Thermal Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
8.1.1 Thermometric properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
8.1.2 Temperature scales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
8.2 Thermodynamic Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
8.3 Quantity of heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
8.4 Thermal Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
8.4.1 The anomalous expansion of water . . . . . . . . . . . . . . . . . . . . . . . . . 52
8.5 Calorimetry and Phase changes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
8.6 Work done by a gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
8.7 The Ideal Gas: Macroscopic Properties of a system. . . . . . . . . . . . . . . . . . . . . 54
8.7.1 Standard temperature and pressure (STP) . . . . . . . . . . . . . . . . . . . . . 55
9 The First law of Thermodynamics 55
9.1 The Microscopic description and the kinetic theory of gases. . . . . . . . . . . . . . . . 57
9.1.1 Mean free path. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
9.1.2 Equipartition of energy. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
9.1.3 The Dulong and Petit law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
9.1.4 The Maxwellian distribution of velocities. . . . . . . . . . . . . . . . . . . . . . 60
9.2 The Molecular properties of Matter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
2
9.3 Transfer of heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
10 Appendix A 62
3
1 Vectors
Background reading: Chapter 1; you may stop on page 28
A vector quantity has both magnitude and direction. The magnitude is a real number that indicates
how large or small the vector is. The direction species the spatial orientation of the vector (i.e., the
direction in which the vector is pointing). Vector quantities, for example, Force, Velocity, Acceleration,
and Displacement are fully described by specifying their magnitude and direction.
A physical quantity that can be described completely by its magnitude is called a scalar. Examples
are: Mass, Energy.
A vector, for example,

A is written as

A = |

A| a, where |

A| represents the magnitude of the vector


and a represents the direction of the vector

A. a is a special vector, known as the unit vector. Its duty is
to specify the direction of the vector. Note the following about the unit vector:
1. | a |= 1. This means that the magnitude of the unit vector is 1 (or unity, as you may nd in some
texts).
2. a =

A
|

A|
i.e., the unit vector can be found by dividing a vector by its magnitude.
Note: The magnitude of a vector

A, which we have previously written as |

A|, will be represented by


|A| or just A, in the rest of this text.
1.1 Properties of Vectors
Consider two vectors

A and

B. The following rules can be dened for addition (and subtraction):
1. Addition:

A+

B =

C
2. Subtraction:

A + (

B) =

A

B
3. Distributive Law: ( +)

A =

A +

A; and are scalars.


4. Distributive Law: (

A+

B) =

A +

B; where is a scalar.
5. Commutative Law:

A +

B =

B +

A
6. Associative Law: (

A +

B) +

C =

A + (

B +

C)
When two vectors are added, the resultant vector is called the resultant vector or simply, the resul-
tant.
Multiplying a vector by a scalar, e.g., > 0 gives another vector, but with a different magnitude,
unless = 1. In the case where < 0, the direction of the vector is reversed as well.
Dividing a scalar by a vector is not allowed. Dividing a vector by another vector is not allowed.
However, dividing a vector by a scalar is possible, as implied by the following:
Given =
1

, we have that

A =
1

A. The right hand side is just division by .


Vectors

A and

B are said to be antiparallel is they are in opposite directions, regardless of their
magnitudes. If they are in the same direction, then they are parallel. If they are in the same direction
and have the same direction, then they are equal, regardless of their location in space.
4
A
Figure 1: A vector is represented by a line segment with an arrowhead
1.1.1 Geometrical Representation of Vectors
A vector is represented geometrically by a line segment with an arrowhead at one end (Figure 1). The
length of the line segment species the magnitude of the vector, and the direction is indicated by the
arrowhead.
When the beginning and end points of a vector are labeled with letters, for example P at the tail and
Q at the head, the vector is sometimes written as

PQ. In this text, we will use single letters to indicate a


vector, for example vector

P.
1.1.2 Multiplication of Vectors
Vectors obey two laws of multiplication, namely,
1. The dot product:

A.

B = ABcos
2. The cross product:

A

B = ABsin n
1.1.3 The Dot Product
For two vectors

A and

B, the dot product yields a scalar. The angle is the angle between the vectors as
shown in Figure 2. Clearly, when = 90
o
,

A.

B = 0. This is a condition for



A and

B to be perpendicular.
A
B
Figure 2: Vector Dot product
Here are a couple of laws for the dot product:
1. Commutative Law

A.

B =

B.

A
2. k

A.

B = (k

A).(

B) =

A.(k

B) = (

A.

B)k
1.1.4 The Cross Product
For two vectors

A and

B, the cross product yields a vector:

A

B = ABsin n (1)
5
is the angle between

A and

B, and the n is in a direction perpendicular to both

A and

B. Geometrically,
the cross product gives the area of the parallelogram of which

A and

B are sides.
It is often convenient to refer vectors to a common point, known as the origin of a co-ordinate axis.
Let us consider the three dimensional Cartesian co-ordinate system.
1.1.5 The three dimensional Cartesian co-ordinate system.
There are three mutually perpendicular directions each extending to innity in the forward and backward
sense. These are usually labeled the x,y,z directions, or the x,y,z axis. The axis intersect at a point, at
which x = y = z = 0. This is the origin (Figure 3).
i
k
j
x
y
z
Figure 3: The Cartesian co-ordinate system
By convention, the x direction is labelled by the unit vector

i, the y direction by the unit vector

j,
and the z direction by the unit vector

k. These unit vectors are mutually perpendicular and are usually
reserved for the directions x, y, z. Being unit vectors, they have the property
|

i| = 1, |

j| = 1, |

k| = 1. (2)
In addition, since they are mutually perpendicular, they have the properties

i.

i = 1;

j.

j = 1;

k.

k = 1 (3)

i.

j = 0;

i.

k = 0;

j.

k = 0. (4)
1.1.6 Components of a vector
A vector

R connecting the origin to a point P(A
x
, A
y
, A
z
) can be written as

R = A
x

i + A
y

j + A
z

k.
The co-efcients A
x
, A
y
, and A
z
are called the components of the vector

R. A
x
, A
y
, A
z
often represent
points in 3-D space, but in some cases they are functions. When the components are points, the vector

R
is a position vector.
A
x

i, A
y

j, A
z

k are vectors as well and obey the laws earlier specied. Consider the addition of

M =
M
x

i + M
y

j +M
z

k and

N = N
x

i +N
y

j +N
z

k. By applying the distributive law, we get

M +

N = (M
x
+N
x
)

i + (M
y
+N
y
)

j + (M
z
+ N
z
)

k.
6
i
k
j
P
1
2
P
A
r
r
2
1
Figure 4: Vector A
Nowsuppose we have a vector

Athat is not directly connected to the orgin. Let the tail of the vector be
located at point P
1
with co-ordinates (x
1
, y
1
, z
1
), and let the head be located at P
2
with co-ordinates (x
2
, y
2
, z
2
).
Now let the position vector r
1
link the origin to P
1
, and let r
2
link the origin to P
2
. Then we have
r
1
+

A = r
2

A = r
2
r
1
.
Since r
1
= x
1

i +y
1

j +z
1

k and r
2
= x
2

i +y
2

j +z
2

k, we have

A = x
2

i +y
2

j +z
2

k (x
1

i +y
1

j +z
1

k)

A = (x
2
x
1
)

i + (y
2
y
1
)

j + (z
2
z
1
)

k. (5)
We have expressed the vector

A in terms of the co-ordinates of the start and the end point.
1.1.7 Magnitude of a vector
To nd the magnitude of a vector

A = A a, we take the dot product

A.

A , which gives, using (5)

A.

A = ((x
2
x
1
)

i + (y
2
y
1
)

j + (z
2
z
1
)

k).((x
2
x
1
)

i + (y
2
y
1
)

j + (z
2
z
1
)

k)
A
2
= (x
2
x
1
)
2
+ (y
2
y
1
)
2
+ (z
2
z
1
)
2
This last step is possible by using the properties in equations 3 and 4. Finally, we have the magnitude of
the vector in terms of 3-D cartesian co-ordinates:
A =
_
(x
2
x
1
)
2
+ (y
2
y
1
)
2
+ (z
2
z
1
)
2
(6)
Clearly, given a vector

R = X

i +Y

j +Z

k, we have the magnitude expressed in the general form


R =
_
X
2
+Y
2
+Z
2
.
The magnitude of a vector is the same as its length.
7
1.1.8 Components of a vector and the dot product
The dot product can be used to nd the component of a vector in a given direction. Suppose we have a
vector

M; to nd the component in a given direction, say n, we take the dot product n.

M. As a simple
example, let us nd the

k component of

M = M
x

i +M
y

j +M
z

k.

k.

M =

k.(M
x

i +M
y

j +M
z

k)

k.

M = M
z
,
where we have used the properties in equations 3 and 4. The right hand side of this equation gives the
expected component in the

k direction.
1.2 Geometrical methods of vector addition
A
B
Figure 5: Vector A and Vector B
Consider the vectors

A and

B as shown in Figure 5. The vectors can be translated so that the head
of one touches the tail of the other. Then

C is the resultant vector; i.e.,

C =

A +

B (Figure 6).
Alternatively, the vectors could be translated so that they share a common origin O, and by completing
the parallelogram, it is clear once again that the diagonal connecting O and P is just

C, the resultant.
1.2.1 The sine rule and the cosine rule
With reference to Figure 6, we recall the sine rule:
A
sin
=
B
sin
=
C
sin
(7)
B
A C
Figure 6: Head-to-tail addition
8
B
A
C
B
O
A
P
Figure 7: Tail-to-tail addition
We also recall the cosine rule:
C
2
= A
2
+B
2
2ABcos (8)
1.2.2 Vectors in two dimensions
From the ideas in section 1.1.5 on page 6, we can write a vector

A in two dimensions (say x and y) by
considering the magnitude of the vector in the third dimension (z) to be zero. We have

A = A
x

i +A
y

j + 0

k.
From Figure 8, we nd that
P A , A
x y
( )
A
x
A y
i
j
A
Figure 8: Vector

A in two dimensions
cos =
A
x
A
and sin =
A
y
A
.
This gives us another way to write the vector

A in terms of the angle :

A = Acos

i +Asin

j. (9)
The magnitude of the vector expresses Pythagoras theorem:
A =
_
A
2
x
+A
2
y
(10)
Also, we can nd the angle :
tan =
A
y
A
x
. (11)
It is important to note that by convention, the angle is measured counterclockwise fromthe +

i direction.
9
2 Linear Momentum
Background reading: Pages 289 - 306
We begin this section with a reminder of certain concepts concerning the motion of a particle. We
dene a particle in order to simplify computation, by considering that all the mass of an object acts
through a single point (or to consider the object itself as a single point) thereby ignoring the internal
structure, shape, and size of the object.
1. When a particle is at rest within an inertial frame of reference, its position can be described by a
vector r with respect to an origin within the same frame of reference.
2. When the position of the particle changes with time, the particle has a velocity v that is a measure
of the rate of change of the position with time.
3. When the velocity of the particle changes with time, the particle has an acceleration a that repre-
sents the rate at which the velocity changes with time.
Suppose that a sedan and a cargo truck are both travelling on a highway at the same velocity. Under
normal circumstances, we would expect that the force required to bring the truck to rest would be larger
than the force required to bring the sedan to rest. This is due to the difference in mass, and draws our
attention to a very important quantity in physics: the momentum.
An object of mass m moving with velocity v has a linear momentum p according to the equation
p = mv. (12)
2.1 Conservation of Linear Momentum
One important concept in Physics is the Principle of Conservation of Momentum. This states that the
total momentum of an isolated system of interacting particles is always conserved. An isolated system is
one that is neither losing energy nor gaining energy from outside the boundaries of the system.
Now suppose that we have a system of n particles. The total momentum will simply be a vector sum
of the momentum of each particle:

P = p
1
+ p
2
+p
3
+ + p
n

P =
n

s=1
p
s
.
These particles have a certain total initial momentum

P
o
,

P
o
=
n

s=1
p
s
After some time, the momentumof individual particles might change, butthe total momentumwill remain
the same:

P
f
=
n

s=1
p
s
The principle of conservation of momentum states that

P
f
=

P
o
.
10
Consider an initial momentum

P
o
= P
ox

i + P
oy

j + +P
oz

k and a nal momentum



P
f
= P
fx

i +
P
fy

j +P
fz

k. By applying the principle of conservation of momentum, we have,

P
f
=

P
o

P
f


P
o
=

0
(P
fx

i +P
fy

j +P
fz

k) (P
ox

i +P
oy

j +P
oz

k) = 0

i + 0

j + 0

k
(P
fx
P
ox
)

i + (P
fy
P
oy
)

j + (P
fz
P
oz
)

k = 0

i + 0

j + 0

k
P
fx
= P
ox
and P
fy
= P
oy
and P
fz
= P
oz
.
In words, momentum is conserved in each direction.
2.2 Collisions
If objects interact by colliding, part of the kinetic energy involved could be dissipated in various forms. A
collision in which kinetic energy is lost is said to be inelastic. In an elastic collision, no kinetic energy is
lost; this means that kinetic energy is conserved in an elastic collision. If all the energy is lost or retained
in a collision, then we have a totally inelastic or totally elastic collision, as the case may be.
The conservation of kinetic energy gives another equation that is useful in solving problems involving
collisions. This will be discussed in greater detail later.
11
3 Motion
Background Reading: Chapters 2 and 3
3.1 Newtons Laws
We will begin this section by stating three important laws of mechanics proposed by Sir Isaac Newton
(1642 1727).
A particle continues to move at constant velocity in a straight line unless acted on by a force.
By choosing an appropriate inertial frame of reference, a particle moving in a straight line with con-
stant velocity can be seen as a particle at rest; the law still holds in this case.
Newtons second Law of motion is:
The rate of change of momentum of a particle is proportional to the force acting on it.
The second law indicates that a change in the momentumof a particle is due to an applied force. Math-
ematically, Newtons second law is written as a differential equation, the solution of which, theoretically,
determines the trajectory of the particle for all time. Representing force as

F, we have

F =
d p
dt

F =
d
dt
(mv)

F =
dm
dt
v +m
dv
dt
We shall restrict ourselves to cases in which the mass is constant, while keeping in mind that there are
many situations in which motion occurs with varying mass: a space rocket, for example, burns a lot of
fuel and its mass reduces signicantly during ight. In the case where mass is constant,
dm
dt
= 0. Also,
we recall that acceleration a is just the rate of change of velocity:
a =
dv
dt
. (13)
We therefore write Newtons second law as

F = ma (14)
The force

F may be constant or may vary as a function of time, position, or velocity.
Newtons third law of motion is:
An applied force generates an equal reaction force acting in a direction opposite to the direction of
the applied force.
We will return to Newtons laws in more detail, but at this time, we will use the second law to derive
the equations of motion for a particle of mass m acted upon by a constant force.
Before we start, some denitions: The study of the motion of an object without considering the nature
or types of forces causing the motion is called kinematics; when the forces are taken into consideration,
it is called dynamics.
12
3.2 Kinematics
3.2.1 Equations of Motion
We start off with Newtons Second Law, for a constant force

F acting on a particle.

F = m
dv
dt
ma = m
dv
dt
a =
dv
dt
We are interested in the velocity and position of the particle as a function of time. Therefore, we integrate
this equation as follows:
dv
dt
= a
_
dv
dt
dt =
_
adt
v = at +c
1
c
1
is the constant of integration. At time t = 0, the constant c
1
= v
o
, the initial velocity. So we have the
velocity at any time t:
v = v
o
+at (15)
If we write the velocity of the particle as a function of position and time, i.e., v = v[r(t)], we can
write the acceleration as:
dv
dt
=
dr
dt
.
dv
dr
dv
dt
= v.
dv
dr
a = v.
dv
dr
(16)
Note that because of the dot product on the right hand side of equation 16, this equation gives only
the magnitude of the acceleration. We take this further:
d(v.v)
dr
= 2v.
dv
dr
v.
dv
dr
=
1
2
d(v.v)
dr
a =
1
2
dv
2
dr
(17)
Equation 17 can now be integrated to give:
_
1
2
dv
2
dr
dr =
_
adr
v
2
2
= ar +c
2
(18)
13
At the initial position, r = r
o
and v = v
o
, so we have
v
2
o
2
= ar
o
+c
2
c
2
=
v
2
o
2
ar
o
Putting this into Equation 18 and re-arranging, we get
v
2
= v
2
o
+ 2a(r r
o
). (19)
Having found expressions for the velocity, we move on to nd an expression for the position. The instan-
taneous velocity v is the rst derivative of the position, i.e,
v =
dr
dt
. (20)
Therefore, using equation 15,
_
dr
dt
dt =
_
(v
o
+at)dt
r = v
o
t +
1
2
at
2
+c
3
c
3
is the constant of integration. When t = 0, the particle is at the initial position r
o
. Therefore c
3
= r
o
and we have
r = r
o
+v
o
t +
1
2
at
2
. (21)
Throughout this section, we have expressed the acceleration and velocity as derivatives with respect
to time; i.e., these are instantaneous values of acceleration and velocity. It is often desirable to use the
average values. In this case, the acceleration is expressed as
a
av
=
v v
o
t t
o
We can set the initial time t
o
= 0 so that
a
av
=
v v
o
t
. (22)
Similarly, the average velocity is
v
av
=
r r
o
t
. (23)
Now we look at an interesting equation. When motion occurs under constant acceleration, the average
acceleration is equal to the instantaneous acceleration. In this case, starting with equation 19,
v
2
v
2
o
= 2a(r r
o
)
(v +v
o
)(v v
o
) = 2a(r r
o
)
Dividing each side by t,
(v +v
o
)(v v
o
)
t
=
2a(r r
o
)
t
(v +v
o
)a = 2av
av
v
av
=
v +v
o
2
, (24)
which is a useful equation, expressing the average velocity as the arithmetic mean of the nal and initial
velocity.
14
3.2.2 Motion in one dimension
When a particle is moving in one dimension, the kinematic equations take on a simpler form. We do
not really need to consider the vector nature of the equations, except to note carefully when a particle is
moving forward or backwards and assign a positive or negative sign as required.
3.2.3 Parametric Equations of motion
Displacement, velocity, and acceleration can be written in terms of time t alone. Expressed thus, the
resultant equations are parametric in terms of time t. a When displacement is expressed in terms of time,
say r(t), taking the rst derivative with respect to time gives the velocity, and the second derivative gives
the acceleration.
v(t) =
d
dt
r(t)
a(t) =
d
dt
v(t)
3.2.4 Motion in two dimensions
We now turn our attention to the motion of an object thrown into the air at an angle to the horizontal.
Such motion is sometimes called projectile motion. We begin by making the following assumptions:
1. The object is moving in a vertical plane.
2. Frictional effects are negligible.
3. The object moves close to the surface of the Earth thoughout its motion.
The last item in the list is necessary because it is the condition for constant acceleration due to grav-
ity. Before we continue with the discussion on projectile motion, we digress and briey mention the
acceleration due to gravity.
3.2.5 Acceleration due to gravity
From Newtons second law it is clear that any time we drop an object, it experiences a force because it
moves from rest towards the ground. The force that causes objects to fall towards the Earth is called the
force of gravity. Indeed, the force of gravity exists between any objects that have mass. For any two
objects of mass m
1
and m
2
seperated by a distance r bwteen their centres of mass, there is a mutual force
of attraction between them, given by
F = G
m
1
m
2
r
2
. (25)
G is the Universal Gravitational Constant.
Suppose that we have a small object of mass m
1
at a distance h from the surface of the Earth (mass
m
e
). Let the radius of the Earth be r
e
. The force acting on the mass m
1
is
F = G
m
1
m
e
(r
e
+ h)
2
(26)
15
If h is small (i.e., the object is close to the Earths surface) then (r
e
+h)
2
r
2
e
. Therefore, using Newtons
second law (equation 14), the mass m
1
experiences a force
F = G
m
1
m
e
r
2
e
m
1
a = G
m
1
m
e
r
2
e
a = G
m
e
r
2
e
(27)
Equation 27 gives the denition of the acceleration due to gravity, which is given the special symbol, g.
It has been found that the acceleration due to gravity has a fairly constant value:
g = 9.8ms
2
(28)
Note that the condition was that the mass m
1
was not far from the Earths surface. We shall encounter
more on Newtons Law of Gravity later. But now, back to projectile motion.
3.2.6 Projectile Motion
v
o
i
j
Figure 9: Path of a projectile
Figure 9 shows the trajectory of a projectile. This is a particle projected at an angle to the horizontal
with an initial velocity v
o
, and acted upon by the force of gravity, which pulls the particle downward with
an acceleration g.
The horizontal direction is labeled

i and the vertical direction



j.
Our rst task is to express the displacement r, the velocity v and the acceleration a in terms of the
unit vectors

i and

j.
r = x

i +y

j (29)
v = v
x

i +v
y

j (30)
v
o
= v
o
cos

i +v
o
sin

j (31)
a = a
x

i +a
y

j (32)
In the case where there is no acceleration in the horizontal direction, a
x
= 0. Also, the vertical accel-
eration is simply the acceleration due to gravity acting downwards: a
y
= g. Therefore, equation 32
becomes
a = g

j (33)
16
The displacement at any time t is
r = r
o
+v
o
t +
1
2
at
2
x

i +y

j = x
o

i +y
o

j + (v
o
cos

i +v
o
sin

j)t +
1
2
(a
x

i +a
y

j)t
2
x

i +y

j = x
o

i +y
o

j + (v
o
cos

i +v
o
sin

j)t
1
2
(g

j)t
2
(34)
Taking dot products of Equation 34 with

i and

j respectively will give us the components in the horizontal


and vertical directions, as follows:
x = x
o
+v
o
cos t Horizontal component (35)
y = y
o
+v
o
sin t
1
2
gt
2
Vertical component (36)
The velocity is
v = v
o
+at
v
x

i +v
y

j = v
o
cos

i +v
o
sin

j (g

j)t (37)
Selecting the

i and

j directions again, we have
v
x
= v
o
cos Horizontal component (38)
v
y
= v
o
sin gt Vertical component (39)
Equations 35 39 give the following equations for the position and velocity of a projectile:
x = x
o
+v
o
cos t (40)
y = y
o
+v
o
sin t
1
2
gt
2
(41)
v
x
= v
o
cos (42)
v
y
= v
o
sin gt (43)
These are two sets of equations that specify the position and velocity of the particle in the horizontal
and vertical directions. The common variable in these equations is the time, t. Therefore, it is usually
convenient, in solving problems, to eliminate t from the equations of motion. Let us now examine a few
characteristics of projectile motion.
How long does it take to reach the greatest height? We recognise that at the greatest height,
the projectile stalls momentarily in the vertical direction before it begins its descent; i.e., the vertical
component of the velocity goes to zero: v
y
= 0. Applying this condition to Equation 43 gives
0 = v
o
sin gt
gt = v
o
sin
t =
v
o
sin
g
(44)
Equation 44 gives the time to reach the greatest height.
17
What is the greatest height reached by the projectile? We can nd this by putting the time t in
Equation 44 into Equation 41:
y = y
o
+v
o
sin
_
v
o
sin
g
_

1
2
g
_
v
o
sin
g
_
2
y = y
o
+
v
2
o
sin
2

g

v
2
o
sin
2

2g
y = y
o
+
v
2
o
sin
2

2g
y y
o
=
v
2
o
sin
2

2g
(45)
Equation 45 expresses the greatest height reached by the projectile.
What is the total time of ight? We will consider the case where the projectile returns to the original
vertical displacement; i.e., y y
o
= 0. From equation 41,
y y
o
= v
o
sint
1
2
gt
2
0 = v
o
sint
1
2
gt
2
1
2
gt = v
o
sin
t =
2v
o
sin
g
. (46)
Equation 46 gives the total time of ight, when the projectile returns to its original vertical displacement.
What is the greatest horizontal distance traveled? We insert the total time of ight (Equation 46)
into Equation 40:
x = x
o
+v
o
cos t
x x
o
= v
o
cos
_
2v
o
sin
g
_
x x
o
=
2v
2
o
cos sin
g
(47)
x x
o
=
v
2
o
sin 2
g
(48)
The greatest horizontal distance is also called the range, R.
So far, we have expressed the position and velocity in terms of the time t. Now, we will write an
equation describing the motion in terms of position variables alone. We start by making t the subject in
Equation 40
x = x
o
+v
o
cos t
t =
x x
o
v
o
cos
(49)
Putting this into Equation 41
y y
o
= v
o
sin
_
x x
o
v
o
cos
_

1
2
g
_
x x
o
v
o
cos
_
2
y y
o
= (x x
o
) tan
g(x x
o
)
2
2v
2
o
cos
2

(50)
18
Equation 50 expresses the motion of the projectile in terms of spatial coordinates.
3.3 Examples
1. A platform is descending with a constant velocity of 4.0ms
1
. aA ball is dropped from rest on to
the platform froma point 6.0mabove it. Find the time that elapses before the ball hits the platform.
4 m/ s
6 m
Figure 10: Descending Platform
Solution
We need two equations expressing position: one equation for the ball and one equation for the
platform. Setting the ball at the origin we have,
r
b
= r
o
+v
o
t +
1
2
at
2
r
b
=
1
2
at
2
(51)
Equation 51 is possible because the ball starts from rest (v
o
= 0) at the origin (r
o
= 0). For the
platform,
r
p
= r

o
+v

o
t +
1
2
a

t
2
r
p
= 6.0 v

o
t (52)
Equation 52 is possible because the platform moves at constant velocity (a

= 0), and starts at a


distance 6.0m below the origin. When the ball hits the platform, the position co-ordinates must be
19
Figure 11: Vector diagram for ships out of port
equal; this gives us a quadratic equation to solve for the elapsed time, as follows:
r
b
= r
p
1
2
at
2
= 6.0 v

o
t
1
2
(9.8)t
2
= 6.0 4.0t
9.8
2
t
2
+ 4.0t + 6.0 = 0
4.9t
2
+ 4.0t + 6.0 = 0
4.9t
2
4.0t 6.0 = 0
t =
4.0
_
4.0
2
4(4.9)(6.0)
(2)(4.9)
t =
4.0 12
9.8
t = 1.6s or
t = 0.82s
We reject the negative root and report the answer as t = 1.6s.
2. Two ships started simultaneously from the port, one travelling to the north and the other to the east.
Two hours later, the distance between them was 60.0km. Find the speed of each ship, knowing that
the speed of one of the ships was 6kmh
1
greater than the speed of the other.
Solution
Let subscripts e and n denote the ship travelling east and the ship travelling north, respectively.
After two hours, the ships would have traveled distances
r
e
= 2v
e
(53)
r
n
= 2v
n
(54)
From Figure 11 we have, using Pythagoras theorem:
60
2
= (2v
e
)
2
+ (2v
n
)
2
3600
4
= v
2
e
+v
2
n
900 = v
2
e
+v
2
n
(55)
20
But one of the ships, say the east-bound one, has a velocity 6kmh
1
greater than the other, i.e,
v
e
= 6 +v
n
(56)
Put equation 56 into equation 55 to get
(6 +v
n
)
2
+v
2
n
= 900
36 +v
2
n
+ 12v
n
+v
2
n
= 900
2v
2
n
+ 12v
n
= 864
v
2
n
+ 6v
n
= 432 (57)
Solving the quadratic equation,
v
n
=
6
_
6
2
+ 4(432)
2
v
n
=
6 42
2
This gives v
n
= 18kmh
1
, or v
n
= 24kmh
1
. We reject the negative root, and so
v
n
= 18kmh
1
and v
e
= 6 + 18 kmh
1
v
e
= 24kmh
1
3. The driver of a train travelling at 60kmh
1
sees on the same track 200m in front of him a slow
train travelling in the same direction at 20kmh
1
. What is the least retardation that must be applied
to the faster train so as to avoid a collision?
Solution
We will rst write two separate equations; one for the slow train and the other for the fast train.
Note the following:
The slow train has a constant velocity and is not accelerating.
In order to avoid a collision, the fast train must undergo a retardation a, but not necessarily to
stop the train - it is only required to bring the fast train to a velocity just less than the velocity
of the slow train.
The units must be consistent - we will convert kmh
1
into ms
1
.
For the slow train,
r = r
o
+v
o
t +
1
2
at
2
r = 200 + 5.56t (58)
For the fast train,
r

= r

o
+v
o

t
1
2
at
2
r

= 16.7t
1
2
at
2
(59)
21
At the time a collision occurs, both trains will be at the same position: r = r

. This gives us
200 + 5.56t = 16.7t
1
2
at
2
200 11.1t =
1
2
at
2
(60)
In Equation 60, there are two unknown variables: a and t. We need another equation with these
variables in order to solve the problem.
It is required to bring the fast trains velocity down to just under the velocity of the slow train, so
v

= v

o
at
5.56 = 16.7 at
11.1 = at (61)
Now putting Equation 61 into Equation 60 gives
200 11.1t =
1
2
(11.1)t
400 22.2t = 11.1t
400 = 11.1t
t = 36s (62)
Equations 62 and 61 give us the required retardation a :
a =
11.1
t
a =
11.1
36
a = 0.31ms
2
4. A projectile is red from a point on a cliff to hit a mark 200m horizontally from the point and
200m vertically below it. The velocity of the projection is that due to falling freely under gravity
100mfrom rest. Show that the two possible directions of projection are at right angles and that the
times of ight are approximately 4.8 and 11.7 seconds. Take g = 10.0ms
2
.
Solution
The velocity of the projection is given by
v
2
= v
2
o
+ 2gh
v =
_
2gh
v =

2 10 100
v =

2000 ms
1
22
We will now use equation 50:
y y
o
= (x x
o
) tan
g(x x
o
)
2
2v
2
o
cos
2

200 = 200 tan


(10)(200)
2
2(

2000)
2
sec
2

1 = tan
1
2
sec
2

0 = tan
1
2
(1 + tan
2
) + 1
0 = 2 tan 1 tan
2
+ 2
0 = tan
2
+ 2 tan + 1
tan
2
2 tan 1 = 0
tan =
2
_
4 + 4(1)
2
tan =
2 2.83
2
tan = 2.415; or
tan = 0.415
= 67.5
o
or
= 22.5
o
The two angles of projection clearly add up to 90
o
. One of the angles is negative because it dips
clockwise below the +

i direction. (See Figure 12) In order to nd the time, we can use equation
67. 5
22. 5
o
o
i
j
(200, -200)
Figure 12: Two angles of projection
23
49:
t =
x x
o
v
o
cos
t =
200

2000(cos 67.5
o
)
t = 11.7s
and t =
200

2000(cos 22.5
o
)
t = 4.8s
5. A particle moves in a straight line with acceleration 4t
2
ms
2
, where t is in seconds, away from
a point O. Initially the particle is passing through O with velocity 2ms
1
. Find the velocity and
distance moved after 3 seconds.
Solution
We rst write an expression for the acceleration, and then integrate to nd the velocity and the
distance, while applying the initial conditions at each stage.
a = 4t
2
dv
dt
= 4t
2
v =
_
4t
2
dt
v =
4t
3
3
+c
1
c
1
is a constant
But when t = 0, v = 2ms
1
, so
2 = 0 +c
1
2 = c
1
therefore, v =
4t
3
3
+ 2
At t = 3s v =
4 3
3
3
+ 2
v = 38 ms
1
The distance r will be found by further integration.
dr
dt
=
4t
3
3
+ 2
r =
_
(
4t
3
3
+ 2)dt
r =
4t
4
12
+ 2t +c
2
where c
2
is a constant
However, when t = 0s, r = 0m, and therefore c
2
= 0. We have, at t = 3s,
r =
t
4
3
+ 2t
r =
3
4
3
+ 2 3
r = 33 m
24
4 Force
Background Reading: Chapters 4 and 5 of Ref. 1
We now return to Newtons Laws. We dene a force as the element capable of initiating motion as
expressed in Newtons First Law.
4.1 Addition of forces
When a number of forces act on a particle at the same time, the total force

F
T
is found by taking the
vector sum of all the forces:

F
T
=

F
1
+

F
2
+

F
3
+ +

F
n

F
T
=
n

s=1

F
s
(63)
Newtons second law is then more appropriately written as
n

s=1

F
s
=
d p
dt
. (64)
When a particle is subject to forces but there is no change in momentum (i.e., momentum is constant),
Equation 64 becomes
n

s=1

F
s
=
d p
dt
= 0
This means that the vector sum of all the forces acting on the particle is zero, which is the condition for
equilibrium. Note that equilibrium occurs when p is constant or zero; in both cases
d p
dt
is zero.
In order to solve problems involving forces, the methods of vector addition outlined earlier come in
useful. Choosing between the use of geometrical methods or vector (analytical) methods depends on the
complexity of the problem. However, it is important, before attempting a problem, to get a good visual
representation of the situation described. It is helpful to draw a sketch of the problem, indicating all the
forces acting, and noting the point through which the forces act. The point through which all the forces
act should be set at the origin of the coordinate system. (Note, however, that in some cases, forces act
at different points on an object or objects. In that case, Newtons law should be applied to each point
seperately).
4.2 Impulse
During an impact or a collision, very large forces act over a short time period - for example, when a golf
ball is struck by a club. Such forces are called an impulsive forces. Such forces are functions of time.
An impulse

I is dened as

I =
_
t
t0

F(t)dt. (65)
However, it might not be possible to express the force explicitly as a function of time. We take care of
25
T
F
g
Figure 13: Mass hanging from a support
this situation by writing the force in terms of the momentum. Recall that
d p
dt
=

F(t)
_
p
f
pi
d p =
_
t
t0

F(t)dt
p
f
p
i
=
_
t
t0

F(t)dt (66)
Equation 66 relates the impulse to the change in momentum. Therefore, by nding the change in momen-
tum of an object, one can determine the impulse that causes the change in momentum. The right hand
side of Equation 66 can be expressed in terms of an average constant force

F
av
acting over a time interval
t as follows:
p
f
p
i
=

F
av
t (67)
4.3 Tension and the motion of connected masses
A string or rope is said to be under tension when applied forces cause it to become taut. Figure 13 shows a
mass hung from a rigid support by a light, inextensible string. The force in the string is called the tension,
or more appropriately, the tension force.
In this course, we will consider strings or ropes that are light (=massless) and inextensible (=we can
ignore elastic properties). For such cases, the tension can be thought of as a transmitted force, and is
the same throughout the string or rope. (When the mass is considered the tension force is generally not
the same throughout the rope).
Let us consider two cases of masses that are connected by strings. The rst one is called Atwoods
Machine, after its originator (we will ignore the physical effect of the pulleys in what follows). Referring
to Figure 14, we apply Newtons second law to each mass separately, assuming that the mass m
1
moves
upwards and the mass m
2
moves downwards:
T
1

j m
1
g

j = m
1
a
1

j (68)
T
2

j m
2
g

j = m
2
a
2

j (69)
We note two important items:
1. Since the masses move together, they have the same acceleration, so that a
1
= a
2
= a
2. The tension is the same throughout the string, so T
1
= T
2
= T.
26
m
2
m
1
T
1
T
2
Figure 14: Atwoods Machine
With these two details, equations 68 and 69 reduce to
T m
1
g = m
1
a (70)
T m
2
g = m
2
a (71)
We solve these simultaneously to get
a =
(m
2
m
1
)g
m
1
+m
2
T =
2m
1
m
2
g
m
1
+m
2
Next, we consider the system in Figure 15, where m
1
is on a smooth surface. Applying Newtons
second law to each mass separately, we get the following equations:
m
1
a
1

i = T
1

i (72)
m
2
a
2

j = T
2

j m
2
g

j (73)
Applying the conditions
1. a
1
= a
2
= a
2. T
1
= T
2
= T.
We have from Equation 72:
T = m
1
a, (74)
27
m
2
m
1
T
T
Figure 15: Two masses connected by a string
which we substitute into Equation 73 to get
m
2
a

j = m
1
a

j m
2
g

j
m
2
a = m
1
a m
2
g
m
2
a m
1
a = m
2
g
a =
m
2
g
m
2
+m
1
(75)
We use Equation 74 to get the tension T:
T = m
1
a
T =
m
1
m
2
g
m
2
+m
1
4.4 Friction
Whenever surfaces are in contact, electrostatic interactions on a microscopic scale occur between the
surfaces, which gives rise to a force that opposes the movement of the surfaces against each other. This
force is called friction, and in general, it opposes motion. Friction is present when a solid object moves
through a uid; when a uid moves through a pipe; when a solid object slides or rolls on a surface, etc.
Interestingly, friction is actually responsible for many types of motion. For example, a motor-car
would not move at all if it were not for friction between the tyres and the road. Also, walking would be
impossible - imagine trying to walk on a very, very slippery oor.
Friction is a dissipative force: in an isolated system of interacting particles where friction is present,
kinetic energy is lost as heat and cannot be recovered.
In this section, we will investigate how friction affects motion.
We identify two kinds of friction:
Static Friction
Kinetic Friction.
Kinetic friction f
k
opposes the motion of an object that is already moving, i.e., it always acts opposite
to the direction of motion. Static friction f
s
tries to keep an object in its original position, and opposes
the inception of motion.
Consider a block resting on a rough surface in Figure 16. Two forces are identied in the diagram:
the weight W of the block, and the normal reaction N. The normal reaction is normal to the surface, i.e.,
is at 90
o
to the surface.
28
Figure 16: Block resting on a rough surface
Experiment has shown that the force of friction is related to the normal reaction through a constant
called the co-efcient of friction, . There is a co-efcient of static friction and a co-efcient of kinetic
friction, dened as follows:
f
s

s
N (76)
f
k
=
k
N (77)
In equation 76, the inequality suggests that the static friction is not constant. Static friction varies with
the applied force, increasing to match the applied force until the applied force is able to initiate motion.
If we consider a block of wood resting on a rough surface, experience teaches us that a larger force is
required to start the block moving than to keep it in motion. This indicates that

s

k
For an object on a horizontal surface (Figure 16), the normal reaction N is equal to the weight W of
the object, but this is not the case for an object on an inclined plane.
W
N
a
b
P
f
k
Figure 17: Object sliding down an incline
Figure 17 shows a small object sliding down an incline of
o
to the horizontal, under the inuence of
a force

P. Let a and

b represent the direction perpendicular to the incline and the direction up the incline,
respectively. The motion of the object is impeded by the force of friction f
k
, which acts in the direction

b while the block slides in the direction

b. We state the following: the applied force



P makes an angle
with

b, and the normal reaction



N makes an angle with the upward vertical

j. From Newtons second
29
law we have
n

s=1

F
s
=
d p
dt

P +

N +

f
k
+

W = ma

P +N a +f
k

b mg

j = ma

P +N a +
k
N

b mg

j = ma

b (78)
In the nal step of writing equation 78 we used the fact that f
k
=
k
N.
Activity From equation 78, show that when the object in Figure 17 is sliding down the incline at
constant velocity, the magnitude of the force

P can be written as
P =
mg(sin
k
cos )
cos +
k
sin
(79)
4.5 Examples
1. Find the tension in the cable supporting the 600N block in Figure 18. The system is in equilibrium.
600 N
30 30
o o
Figure 18: Mass hanging from vertical support
Solution We draw a free-body diagram as follows (Figure 19), with the point through which all the
forces act set at the origin.
30
30
o o
T
1
T
2
i
j
W
Figure 19: Free body diagram for Figure 18
30
We start by quoting Newtons second law, and then we substitute the seperate vectors as follows:
n

s=1

F =

0

T
1
+

T
2
+

W =

0
T
1
cos 30
o

i +T
1
sin 30
o

j +T
2
cos 150
o

i +T
2
sin 150
o

j W

j =

0
We can now separate the equation into the

i and

j directions.
T
1
cos 30
o
+T
2
cos 150
o
= 0 (80)
T
1
sin 30
o
+T
2
sin 150
o
W = 0 (81)
From equation 80 we have
T
1
=
T
2
cos 150
o
cos 30
o
T
1
=
T
2
(0.866)
0.866
T
1
= T
2
Inserting this result into equation 81 gives
T
2
sin30
o
+T
2
sin150
o
W = 0
T
2
(0.5 + 0.5) = 600
T
2
= 600N
Therefore, T
1
= 600N as well.
2. A 3.00kg steel ball strikes a wall with a speed of 10.0m/s at an angle of 60.0
o
with the surface. It
bounces off with the same speed and angle. If the ball is in contact with the wall for 0.200 s, what
is the average force exerted on the wall by the ball?
Solution
We can solve this problem by using the relation between the momentum, the average force, and the
time, expressed in Equation 67.
p
f
p
o
=

F
av
t
mv
f
mv
o
=

F
av
t
This problem involves an elastic collision, since we are told that the ball bounces off the wall with
the same speed and angle.
From Figure 20, we have that
mv
f
mv
o
=

F
av
t
mv
f
(cos 150

i + sin 150

j) mv
o
(cos 30
o

i + sin30
o

j) =

F
av
t
mv
f
(

3
2

i +
1
2

j) mv
o
(

3
2

i +
1
2

j) =

F
av
t
31
60
o
60
o
Figure 20: Ball bounces off the wall
Since v
f
= v
o
, the equation reduces to
mv
o
(

3
2

i +
1
2

j) mv
o
(

3
2

i +
1
2

j) =

F
av
t
mv
o
(

3
2

i) mv
o
(

3
2

i) +mv
o
1
2

j mv
o
1
2

j =

F
av
t
mv
o

i =

F
av
t

F
av
=
30

3
0.2

F
av
= 260

iN
3. A 6.00kg block of wood rests on a wooden plank that is 10.0m long. The co-efcient of static
friction
s
between the plank and the block is 0.364.
(a) To what height should one end of the plank be raised to just start the block sliding downwards
at constant velocity?
(b) In order to stop the block from sliding down any further, a force P is applied to the block,
directed perpendicularly downwards to the plank. What is P, given that the co-efcient of
kinetic friction
k
= 0.250?
Solution
Since the block is supposed to just start sliding downwards under constant velocity, the total accel-
eration is zero. Newtons law gives

F =

0
N a W

j +f
s

b =

0 (82)
We are supposed to nd h (see Figure 21), so it would help if we found cos , sin , or tan .
The total force acting in direction a is:
N W cos + 0 = 0
N = W cos (83)
32
Total force in direction

b is
0 W cos(90 ) +f
s
= 0
W sin = f
s
W sin =
s
N (84)
W
N
a
b
f
s
10. 00 m
h
Figure 21: Block about sliding down an incline
We can now solve Equations 83 and 84 simultaneously:

s
N
N
=
W sin
W cos

s
= tan
= tan
1
(0.364)
= 20.0
o
The angle at which sliding motion occurs (if no external force is applied) is called the critical angle.
Therefore, the critical angle for this situation is 20
o
.
We now nd the height:
sin 20 =
h
10
h = 3.42m
When an additional force P a is now applied to stop the object sliding any further, we have the
total force
P a +N a W

j +f
k

b =

0 (85)
In the direction a we have:
P +N W cos + 0 = 0
P +N = W cos (86)
In the direction

b:
W sin +f
k
= 0
W sin =
k
N
N =
W sin

k
(87)
33
We can solve Equations 86 and 87 simultaneously to get
P +
W sin

k
= W cos
P = W cos
W sin

k
P =
W

k
[sin
k
cos ]
P =
6 9.8
0.250
[sin 20 0.25 cos 20]
P = 25.2N
34
5 Circular Motion
Background reading: Chapter 3 Section 3.4, Chapter 5 Section 5.4, Chapter 9 Section 9.1 of Ref.1;
Chapter 4 Section 4.29 of Ref.2
5.1 The basic equations
Newtons rst law suggests that an object moving in a circular path must have a force acting on it. This
force is called the centripetal force,

F
c
, which constrains a particle of mass m to move in a circular path
of radius r by attracting it continuously toward the centre of the circle. We shall begin this section by
deriving equations of motion for an object moving in a circular path, and we will use the convention of
placing a dot over a variable to denote differentiation with respect to time.
r
v
Figure 22: Particle moving in a circular path
Consider a particle moving along a curve as in Fig. 5.1. We consider that the particle moves through
a small angle d in a time dt. The position of the particle has unit vector r given by
r = cos

i + sin

j (88)
The unit tangent to the position unit vector is given by

:

= sin

i + cos

j (89)
We can nd out how these unit vectors change with time by taking the rst derivative of each (note
that is a function of time as well). First, the unit vector r:
r = cos

i + sin

r =

sin

i +

cos

j (90)

r =

(sin

i + cos

j) (91)
Next, the unit vector

:

= sin

i + cos

cos

i

sin

(cos

i + sin

j) (92)
from which we see that

r =

(93)

r (94)
35
The position vector of the particle is
r = r cos

i +r sin

j
r = r r (95)
To nd the velocity of the particle, we take the derivative of the position with respect to time,
r = r r

r = r r +r

r = r r +r

(96)
The acceleration is the derivative of the velocity with respect to time:

r = r r + r

r + r

+r

+r

r = r r + r

+ r

+r

r = ( r r

2
) r + (2 r

+r

(97)
Therefore, in the direction r we have the radial component of acceleration
r r

2
(98)
and in the direction

the tangential component
2 r

+r

(99)
In the event that the particle is constrained to move along a path of xed radius, r = 0 and r = 0.
Then we have, from Equation 96, the tangential velocity
v = r

(100)
Also, from Equations 98 and 99, the radial (or centripetal) acceleration is
a
r
= r

2
(101)
and the tangential acceleration
a
t
= r

. (102)
The total acceleration is

r = a
r
r +a
t

(103)

r = r

+r

r
r =
_
(r

2
)
2
+ (r

)
2
= r
_

4
+

2
(104)
From Equation 101, we see that an object moving in a circular path has an acceleration towards the
center of the circle. This leads to the centripetal force by applying Newtons second law:
F
c
= mr

2
. (105)
Let us put together the relevant quantities:
= angular displacement
=

=
d
dt
= angular velocity
=

=
d
2

dt
2
= angular acceleration
In the foregoing, the angle must be dened in units of radians.
36
5.2 Uniform Circular Motion
When the particle moves along the circular path at constant velocity, the motion is described as uniform. It
means that the object is not accelerating in the direction of the velocity, even though there is a centripetal
acceleration towards the center of the circle.
If s is the distance traveled along the arc of a circle, then in terms of the radius r and the subtended
angle , we recall the relation
s = r (106)
The average angular velocity is given by
=

2

1
t
2
t
1
(107)
where all symbols have their usual meanings.
In terms of the angular velocity, the linear velocity is
v = r (108)
In terms of , the magnitude of the centripetal acceleration can be written as
a
c
=
2
r. (109)
5.3 The Conical Pendulum
r
l
h
Figure 23: A conical pendulum
When an object is whirled in a horizontal circle about a vertical axis, the physical situation is described
as a conical pendulum. In Figure 23, a small object fastened to a string is whirled around such that it
follows a circular path in the horizontal plane. We will consider the case where the angular velocity is
constant.
We analyse this motion by referring to Figure 24. From Newtons second Law, we can sum the forces:

F
g
+

T =

F
c
mg

j +T cos

i +T sin

j =
mv
2
r

i (110)
It is customary to write the equation in terms of the angle the string makes with the vertical (here, ),
37
so we have
mg

j +T sin

i +T cos

j =
mv
2
r

i (111)
T sin =
mv
2
r
(112)
and T cos = mg (113)
Equations 112 and 113 give
tan =
v
2
gr
(114)
F
T
F
c
g
Figure 24: Free-body diagram for conical pendulum
5.4 Non-uniform circular motion
In the case where the angular velocity of the object is not constant, the motion of the object is non-
uniform, and we have to use equation 103. Notice that in this case the total acceleration is the resultant of
components in two perpendicular directions r and

, and so our earlier discussion on vectors in a plane
hold; in particular, the angle between the resultant and any of the directions can be found.
A good example for non-uniform circular motion is provided by an object that is fastened to one end
of a string and whirled in a vertical circle. We can determine three forces acting on the object (Figure
25): mg, vertically downwards;
mv
2
r
, towards the centre of the circle; and T, the tension in the string.
In order to determine the nature of the motion in the vertical circle, we are better served by using
methods involving the conservation of energy, even though we could get a fair idea of how the net force
changes as the object moves through the circle by using the methods from Newtons second law.
At this time, though, we will use Newtons second law to establish the relation between the forces at
two extremes of the objects motion: when the object is at the highest point and when the object is at the
lowest point.
At the highest point we can write
T

j mg

j =
mv
2
r

j
T =
mv
2
r
mg (115)
38
Figure 25: Object moving in a vertical circle
At the lowest point we have
T

j mg

j =
mv
2
r

j
T =
mv
2
r
+mg (116)
5.5 Examples
1. A particle is moving in a horizontal circle with an initial angular angular velocity
o
=
2

rev/min.
What angular acceleration must be applied in order to bring the particle to rest after a time of 3s.?
Solution
We rst convert the units rev/min to rad/s, and then we apply the equation relating the average
angular acceleration to the change in angular velocity, with the requirement that the nal angular
velocity = 0.

o
=
2

rev
min

o
=
2

2rad
60s

o
=
1
15
rads
1
(117)
Then,
=

o
t
=
1
45
rads
2
39
6 Work and Energy
Background Reading: Chapters 6 and 7 of Ref 1.
We are by this time in our education familiar with the concept of energy: energy is what is required to
do work. But what is work, or when can we conclude that work has been done? It is perhaps easiest to
understand mechanical work, where we conclude that work has been done when a force moves an object
through a distance.
6.1 Work done by a varying force
Let us suppose that a force

F
1
, moves an object through an innitesimal displacement dr
1
. An innitesi-
mal amount of work dW will be done, given by the dot product
dW =

F
1
.dr
1
(118)
Now suppose that the force acting changes to

F
2
, moving the object through another innitesimal dis-
tance. Suppose again that this is followed by a series of changes in the force, each causing the object
to undergo an innitesimal displacement. Then, after a measurable distance, the total work done on the
object can be found by adding the innitesimal bits of work done by each force:
W =
n

s=1

F
s
.dr
s
(119)
However, since the innitesimal displacements are too small to add up realistically, the sum in equation
119 is evaluated as an integral:
W =
_
r1
r0

F.dr (120)
Equation 120 expresses the work done by a varying force in moving an object between points r
0
and r
1
.
6.2 Work done by a constant force
With reference to Equation 120, we nd that if the force is constant, we can write
W =
_
r1
r0

F.dr
W =
_
r1
r0
F cos dr
W = F cos
_
r1
r0
dr
W = F(r
1
r
0
) cos (121)
The angle is the angle between the direction of the force and the direction in which the object moves.
If the object moves in the same direction as the force, = 0 and the work done is
W = F(r
1
r
0
)
40
6.3 Work and Kinetic Energy
Starting from equation 120, and writing the acceleration a as v.
dv
dr
, we have
W =
_
r2
r1

F.dr
W =
_
r2
r1
ma.dr
W =
_
v2
v1
mvdv
W =
1
2
m(v
2
2
v
2
1
) (122)
Equation 122 indicates the change in kinetic energy of an object that is accelerated from velocity v
1
to v
2
. In general an object of mass m moving with velocity of magnitude v has a kinetic energy
K.E =
1
2
mv
2
(123)
Equation 122 is the work- kinetic energy equation, usually written as
W =
1
2
mv
2
2

1
2
mv
2
1
(124)
6.4 Work and Potential Energy
Consider what happens when a ripe mango falls from a tree. The mango accelerates towards the ground,
which means that work is done on it, which in turn means that energy is expended. Where did this energy
come from?
If we returned the mango to its original height and dropped it, it would once again accelerate down-
wards in the same manner. Furthermore, were the mango to drop from a greater height, its velocity
on reaching the ground should be greater. This suggests that the mango possesses energy by dint of its
position; better put, the energy depends on the distance between the Earth and the mango. By raising the
mango to a height, energy is stored within the Earth-mango system, and this energy is released when
the mango is dropped. Until the mango is dropped however, the energy is stored: it remains as potential
energy. In this particular case, we are dealing with the gravitational potential energy U(r).
If we set r
0
as our reference point (for example, the surface of the Earth), then the change in gravita-
tional potential energy for an object moving from r
0
to r
1
is
U(r
1
) U(r
0
) =
_
r1
r0

F.dr (125)
A reference point is an essential part of the denition of the potential energy because technically, we do
not speak about the potential energy of an object but rather of a system. Therefore the position of the
object within the system must be dened, and this is done by establishing a reference point.
Now suppose that an object moves from position r
1
to r
2
. The net change U(r) = U(r
2
) U(r
1
)
41
in potential energy is written in terms of the potential energy at each point, as follows:
U(r
2
) U(r
1
) = [U(r
2
) U(r
0
)] [(U(r
1
) U(r
0
)]
U(r
2
) U(r
1
) =
_

_
r2
r0

F.dr
_

_
r1
r0

F.dr
_
U(r
2
) U(r
1
) =
_
r2
r0

F.dr +
_
r1
r0

F.dr
U(r
2
) U(r
1
) =
_
r0
r2

F.dr +
_
r1
r0

F.dr
U(r
2
) U(r
1
) =
_
r1
r2

F.dr
U(r
2
) U(r
1
) =
_
r2
r1

F.dr (126)
Also, comparing the right-hand side of equation 125 to the right-hand side of equation 120, we con-
clude that the change in potential energy is just the negative of the work done in moving the object
between points r
0
and r
1
:
U(r) = W (127)
6.5 The Principle of Conservation of Energy
From the equations 127, 126 and 124, we nd
W = U(r
1
) U(r
2
)
1
2
mv
2
2

1
2
mv
2
1
= U(r
1
) U(r
2
)
1
2
mv
2
2
+U(r
2
) =
1
2
mv
2
1
+U(r
1
) (128)
Equation 128 expresses the conservation of mechanical energy: the total initial mechanical energy (when
the object is at point r
1
) is equal to the total nal mechanical energy (when the object is at point r
2
.)
The total mechanical energy of an object consists of the kinetic energy and the potential energy.
6.6 Conservative and non-conservative forces
The work done by a conservative force is path independent. This means that the work done depends only
on the initial and nal co-ordinates, so that when such a force moves an object through a distance but
ends up at the starting point, the total work done is zero. Examples of conservative forces are Hookes
force F = kx and Newtons force of gravity. Friction is a non-conservative force.
6.7 Example
A spring with force constant 40Nm
1
has a mass of 0.20kg attached to one end, the other end being
held xed. The mass is free to slide on a frictionless horizontal surface. If the mass is released when the
spring is stretched by 0.50m, what is the speed of the mass when the spring is compressed by 0.10m?
Solution
42
Figure 26: Spring stretched and compressed
With reference to Figure 26, we use the work-energy theorem,
W
12
= K.E
2
K.E
1
_
x2
x1

F.dr =
1
2
mv
2
f

1
2
mv
2
i

_
x2
x1
kxdx =
1
2
mv
2
f

1
2
mv
2
i
1
2
kx
2
1

1
2
kx
2
2
=
1
2
mv
2
f
1
2
kx
2
1

1
2
kx
2
2
=
1
2
mv
2
f
1
2
40(0.5
2
0.1
2
) =
1
2
0.2v
2
f
v
f
= 6.9ms
1
43
7 Some aspects of Newtonian Gravitation
Background Reading: Chapter 12 of Ref. 1.
Physics recognises four fundamental forces of nature:
The Strong Nuclear Force
The Weak Nuclear Force
The Electromagnetic force
The Force of Gravity
This section introduces the force of gravity, which is the weakest of the four. We begin this section by
discussing Keplers Laws, which were proposed on the basis of empirical evidence, even before Newton
proposed his Universal Law of Gravitation. Keplers laws are a natural result of Newtons law, and the
fact that gravity is a central force.
7.1 Keplers Laws
Johannes Kepler (1571-1630), German Astronomer and assistant to Danish Astronomer Tycho Brahe,
promulgated the following laws after 16 years of studying data on the motion of some planets about the
sun, collected over a period of 20 years.
The remarkable item that he discovered was that the motion of the planets about the sun could be
modeled after the simple geometrical gure, the ellipse, with the sun at one of the foci.
a
b
c
Peri hel i on
Aphel i on
Pl anet
Sun
Figure 27: A planet moving around the sun in an elliptical path
Keplers First Law: The motion of the planets around the sun follows an elliptical path.
Figure 27 shows a planet in an elliptical orbit around the sun. The distance a is the semi-major axis and
the distance b is the semi-minor axis. The point at which the planet is closest to the sun is called the
perihelion, and the point at which it is furtherest from the sun is called the aphelion. We must, at this
juncture, add that because gravitation is universal, Keplers laws hold for satellites of planets as well.
When a satellite is closest to the Earth, it is at perigee, when it is furtherest, it is at apogee.
Keplers Second Law: The line joining the sun to a planet sweeps out equal areas in equal time
intervals.
This is another way of saying that the angular momentum is conserved. We can prove this is a simple
way. With reference to Figure 28, the angular momentum

L is given as the vector cross product

L = r p

L = r mv (129)
44
v
m
r
Figure 28: Area swept out by radius vector
After a time dt, the (triangular) area swept by the radius vector is given by the element d

A represented
by the cross product
d

A =
1
2
r vdt
Equations 129 and 130 give
L = 2m
dA
dt
L
2m
=
dA
dt
(130)
The conservation of angular momentum implies that the total angular momentum L is always constant,
and since the mass m is a constant, the left-hand-side of Equation 130 is constant. Therefore
dA
dt
is also
constant.
Keplers Third Law: The square of the orbital period of a planet is proportional to the cube of the
semi-major axis of the elliptical orbit.
Having agreed that planets move in elliptical paths, period has its usual meaning: how much time it
takes for a planet to return to the same position at which it was initially observed. For example, it takes
one year (365
1
4
days) for the Earth to return to the same position with respect to the stars. Keplers third
law tells us that the period is proportional to the cube of the distance from the Earth at peri- or aphelion
to the centre of the elliptical orbit.
T
2
= Ka
3
The constant of proportionality K is
4
2
GM
, so
T
2
=
_
4
2
GM
_
a
3
(131)
7.2 The Law of Universal Gravitation
This Law was rst proposed by Sir Isaac Newton, to t empirical data from astronomical observations.
The force of gravity is sensible to mass, and, given any two objects of mass m
1
and m
2
, causes a
mutual attraction the magnitude of which is
F =
Gm
1
m
2
r
2
45
The constant G is the universal gravitational constant. r is the distance between the objects, measured
from their centres of mass.
Since the gravitational force between two objects is mutually attractive, an observer on mass m
1
will
measure the mass m
2
approaching (negative displacement); an observer on mass m
2
will measure mass
m
1
approaching (negative displacement). This is illustrated in gure 29. Therefore the direction of the
force is always negative, as the following equations indicate:

F
1,2
=
Gm
1
m
2
r
r
1,2

F
2,1
=
Gm
2
m
1
r
r
2,1
m
m
2
1
r
r
1, 2
r
2, 1
Figure 29: Mutual attraction
7.3 Gravitational Potential Energy
Let us now nd the potential energy possessed by a system of two masses interacting via the force of
gravity alone.
If we suppose that mass m
2
(usually the smaller of the pair) moves from position r
1
to r
2
within the
gravitational eld of mass m
1
, the potential energy of the system will change by the amount
U(r
2
) U(r
1
) =
_
r2
r1

F.dr
U(r
2
) U(r
1
) =
_
r2
r1
Gm
1
m
2
r
2
( r).dr
U(r
2
) U(r
1
) =
_
r2
r1
Gm
1
m
2
r
2
dr
U(r
2
) U(r
1
) =
_
Gm
1
m
2
r
_
r2
r1
U(r
2
) U(r
1
) = Gm
1
m
2
_
1
r
2

1
r
1
_
(132)
By convention, the reference point is usually chosen as the point at which the potential energy is zero.
If we set r
1
as the reference point, we will require that it is innitely far away from r
2
(i.e., r
1
).
Then the potential energy U(r
1
) = 0. Equation 132 becomes
U(r
2
) 0 = Gm
1
m
2
_
1
r
2
0
_
U(r
2
) =
Gm
1
m
2
r
2
(133)
46
In general the gravitational potential energy is dened as
U
g
=
Gm
1
m
2
r
. (134)
7.4 Escape Velocity
The total energy of a system composed of an object moving in a gravitational eld is given, as expected,
by the sum of the kinetic and potential energies of the particle. Consider a small object of mass m
moving in the gravitational eld of much larger object of mass M. An example could be a rocket carrying
a satellite into Earth orbit. This system will have total energy E
T
given by
E
T
= K.E +U(r)
E
T
=
1
2
mv
2

GMm
r
(135)
If the velocity of the rocket is such that the gravitational potential energy is larger than the kinetic energy
of the rocket, then the total energy of the system is negative (E
T
< 0) and the rocket will remain trapped
within the gravitational eld of the Earth.
In the case where E
T
> 0, the rocket will eventually escape the Earths gravitational eld. When
E
T
= 0, the initial velocity of the rocket is called the escape velocity: this is the smallest velocity the
rocket requires to break free from the Earths gravitational eld. From equation 135 we have
0 =
1
2
mv
2
esc

GMm
r
e
v
2
esc
=
2GM
r
e
v
esc
=
_
2GM
r
e
(136)
7.5 Example
1. Calculate the speed of a satellite in a circular orbit around the planet Mercury, 5.00 10
3
m above
the surface. Mercury has mass 3.24 10
23
kg and mean radius 2.42 10
6
m.
Solution
There are a number of ways to solve this problem, but we will use Keplers third law, replacing the
semi-major axis of an elliptical orbit a with the radius r of a circular orbit. We will recall that the
period T is related to the angular velocity as follows:
T =
2

47
By Keplers third law:
T
2
=
_
4
2
GM
_
r
3
_
2

_
2
=
_
4
2
GM
_
r
3
GM =
2
r
3
GM =
v
2
r
2
r
3
GM
r
= v
2
v =
_
6.67 10
11
3.24 10
23
2.42 10
6
+ 5 10
3
v = 2.99 10
3
ms
1
48
8 Thermal Physics
Heat is a form of energy. We nd out that the most appropriate denition for heat energy involves the
transfer of heat: we say that heat is the energy that ows fromone region to another because of a difference
in temperature. In addition, heat can be made to do work.
Temperature is a measure of how hot or how cold and object is. When heat energy is transferred
to or from an object, it is often the case that some observable physical property of the object will be
altered. For example, a metal object might change in size; the electrical resistance of a conductor might
change; the electromagnetic radiation emitted by the object might change, and so on. By standardising
such observations, a temperature scale can be created.
We will see shortly that the property of an object that is changed as a result of the loss or gain of heat
energy is called the thermometric property.
8.1 Thermal Equilibrium
Let us consdier objects A and B, that are in contact with each other, so that heat energy can ow between
them without hindrance. Then, so long ats there is a difference in temperature, heat energy will be
transferred from the object with the higher temperature to the object with the lower temperature. This
transfer will continue until both objects reach the same temperature. When this happens, the objects are
said to be in thermal equilibrium.
Suppose now that we have three objects A, B, and C placed such that A is in contact with C, and B is
in contact with C, but A and B are not in contact with each other. Suppose also that, heat energy cannot
ow from A to B, or from B to A.
We use this setup to state the zeroth law of thermodynamics: If objects A and B are separately in
thermal equilibrium with object C, then A is in thermal equilibrium with B.
8.1.1 Thermometric properties
Thermometric properties are the observable physical changes of a material that is undergoing a tempera-
ture change, or that is absorbing or rejecting heat, where the physical changes are caused by the transfer
of the heat energy. For example, in a metal the thermometric property could be a change in volume (ex-
pansion) or a change in electrical resistance; in a gas the thermometric property could be the change in
the volume or the change in the pressure of the gas, and so on. These observable properties are used in
making thermometers by relating the change in the temperature to the thermometric properties.
8.1.2 Temperature scales
There are a number of temperature scales in use, the most common of which, in our part of the world, is
the celcius (or centigrade) temperature scale, calibrated for melting ice at 0

C and boiling water at 100

C,
at standard atmospheric pressure.
There is also the Fahrenheit temperature scale, calibrated for melting ice at 0

F and boiling water at


212

F. The temperature difference between boiling water and melting ice is 180

for the Fahrenheit scale


and 100

for the celcius scale, so that a scale division of 5

C is equivalent to a scale division of 9

F. With
this in mind, we can write down the conversion between the two temperature scales as
T
F
= 32

+
9
5

F (137)
Finally, there is the Kelvin, or absolute temperature scale. This scale is independent of any system used to
measure the temperature, and can be derived directly from theory. To see this, we can consider a constant
volume ideal-gas thermometer used to take measurements of temperature say, of melting ice and boiling
49
water. Figure 30 shows that the straight line, when extended backwards, will cut the temperature axis at
273.15

C. This is taken to be the absolute zero of temperature, i.e., 0 K. At 0

C, the temperature in
Kelvin is 273.15 K. Theoretically, there can be no temperature less than absolute zero; on our scale, it
would imply the rather meaningless notion of a negative pressure.
The conversion between the celcius and the Kelvin temperature scales is
T
K
= T
C
+ 273.15 K (138)
where T
C
is the temperature in

C.
Figure 30: Temperature scale for constant-volume gas thermometer.
8.2 Thermodynamic Systems
A thermodynamic system can be understood by applying either a macroscopic description or a micro-
scopic description. A macroscopic desription involves a description of the state of the entire system
using what are usually called state variables: pressure p, volume V , temperature T, and quantity of mat-
ter, given by the number of moles, n. These variables are related to each other via an equation of state.
On the other hand, a microscopic description involves properties of the particles that make up the
system, for example, the kinetic energy, momentum, and speed of the particles.
There are three thermodynamic systems of interest:
1. Isolated system A system on which no work is done and which does no work on its surroundings.
In addition, no heat energy or radiation enters or leaves an isolated system. In essence, an isolated
system does not interact with its surroundings.
2. Open system It continuously interacts with its environment and can exchange energy and matter
with its surroundings.
3. Closed system In such a system, heat and work, but not matter can be exchanged with the surround-
ings.
We also dene a thermodynamic process as one in which the thermodynamic state of a system in
changed by a continuous variation of the state variables.
8.3 Quantity of heat
We have dened heat as the energy transferred between systems as a result of a temperature difference.
We would like to adress the question, how much heat energy is transferred, and what does the transfer
depend on?
50
The most fundamental way of measuring a quantity of heat energy transferred to a system is to dene
it in terms of a change in the temperature of the system. So, we have the denition of the calorie: 1
calorie is the amount of heat energy required to raise the temperature of 1 gram of water from 14.5

C to
15.5

C.
The S.I unit for heat, being a form of energy, is the joule, J. We relate the calorie to the joule by the
following conversion:
1 cal = 4.165 J (139)
If we represent the heat energy transferred by symbol Q, we can relate this to the mass m and the
temperature change T of the system using the equation
Q = mcT (140)
where c is a constant called the specic heat unique to the system.
Now for we can associate a small temperature change dT, with a small transfer of heat energy dQ,
so:
dQ = mc dT
c =
1
m
dQ
dT
(141)
We recall that the number of moles n is related to the mass m and molar mass M as follows:
n =
m
M
(142)
Then equation 141 can be written as
Mc =
1
n
dQ
dT
C =
1
n
dQ
dT
(143)
The quantity C = Mc is called the molar heat capacity. Experimentally, the molar heat capacity is
determined either at constant volume or at constant pressure, and are labeled C
V
or C
p
respectively. In
general, C
V
= C
p
.
8.4 Thermal Expansion
Most materials will expand when their temperature increases. We call such expansion under a change in
temperature a thermal expansion.
If we have a long, thin rod of length l
0
made of the same material, then for small changes in temper-
ature we can establish a linear relationship between the change in length and the change in temperature:
l
l
0
= T (144)
where is a constant called the co-efcient of linear expansivity. If the rod has an initial length l
0
and a
nal length l, then for a temperature change from T
0
ro T
1
, we have
l l
0
l
= T (145)
51
Figure 31: 1-D crystal
This tendency of materials to expand when the temperature increases leads to thermal stresses in
large buildings and other structures. For example, steel beams within a bridge will expand when the
temperature increases. If the design does not take care of this expansion, the thermal stresses can be high
enough to damage the bridge. Therefore, in large structures, use is often made of expansion gaps to
allow for thermal expansion.
Consider Figure 31, which shows a theoretical one dimensional crystal lattice, in which the atoms
are considered to be bound by elastic forces that obey Hookes law for small displacements from their
mean position. At any temperature T the atoms vibrate about their mean position as a result of the thermal
energy. As the temperature increases, the displacements fromthe mean position become larger and are no
longer proportional to the restoring force. This means that the vibrations are no longer simple harmonic,
and no longer symmetric about the mean position. The net effect of this is an extension in the length of our
one dimensional crystal. We can extend this description to three dimensions to gain a basic understanding
of the origin of thermal expansion of solids.
8.4.1 The anomalous expansion of water
There are some materials for which it is energetically more favourable for a re-arrangement of the crystal
lattice to occur when the temperature increases. This sometimes leads to a contraction in volume, instead
of an expansion.
Pure water shows an anomalous expansion, such that as the temperature increases from 0

to 4

, the
volume reduces. As the temperature increases beyond 4

, the volume starts to increase as well. This is


described as the anomalous expansion of water.
8.5 Calorimetry and Phase changes
We use the word calorimetry to describe the methods and processes used in measuring the amount of
heat transferred in a given process. It is observed that, apart from changes in volume, a solid metal object,
for example, will melt at a certain xed temperature. This indicates a change in phase from the solid
phase to the liquid phase of the metal. There are three phases of matter of interest to us: solids, liquids,
and gases. There are a few other phases, such as plasma and degenerate gases, but we will not consider
them.
During a change in phase, the system absorbs or rejects heat energy, but does not experience a change
in temperature. This is explained by noting that the heat energy absorbed or rejected during the phase
change is used in altering the structure of the system. When ice melts, the temperature remains at 0

C
while the energy absorbed breaks the bonds holding the molecules together as a solid.
The latent heat of fusion, L
f
is dened as the ratio between the energy transferred to the system per
unit mass, during a phase change from solid to liquid, or from liquid to solid.
L
f
=
Q
m
(146)
The latent heat of vaporisation L
v
is dened in a similar manner:
L
v
=
Q
m
(147)
52
A phase change from solid to gas or vice-versa without passing through the liquid phase is called
sublimation.
8.6 Work done by a gas
Consider a gas inside a cylindrical container of cross-sectional area A(Figure 32). One end of the cylinder
is sealed, and the other end is tted with a piston that is free to move. Now, when the gas expands, the
piston moves through a distance dx and the volume of the gas is increased by an amount dV , which from
the diagram is equal to Adx.
Figure 32: Gas inside sylinder with piston.
From our studies in mechanics, we have
W =
_

F dr (148)
But, pressure p is dened as the force per unit area:
p =
F
A
(149)
So, we have
W =
_
pAdx
W =
_
V2
V1
p dV (150)
where we have taken the integration from the initial volume V
1
to the nal volume V
2
.
Let us agree on the sign convention for the work done. When the gas expands and does work on the
piston, we label the work done as positive. When an external agent causes the gas to be compressed by
the piston, we label the work done by the gas as negative. In this manner, we only look at the work done
by the gas and label it as positive or negative according as the gas is expanding or contracting. We can
ignore the nature of the agent that is doing the compression.
The work done by a gas is also given by the area under the curve in a p-V diagram(the meaning of the
integral). The work done in a thermodynamic process depends on the path taken. This is because there
53
are many ways in which to change the state of a gas from p
i
, V
i
, T
i
to p
f
, V
f
, T
f
; i.e, on a p-V diagram,
many different paths could be followed, and since the work done is given by the area under the p-V curve,
it is clear that the work done is path dependent.
We will, at this time, dene different types of thermodynamic processes:
1. Isothermal processes. The temperature is constant throught this process. On a p-V diagram, lines
of constant T are called isotherms.
2. Isobaric processes. Here, the pressure remains constant throughout the process. The path followed
in such a process is called an isobar.
3. Isovolumetric or isochoric processes. In such a process the volume remains constant, and the
path followed is called an isochor.
4. Adiabatic processes. In such processes, there is no heat transferred to or from the gas. On a p-
V diagram, the path followed is called an adiabat. This ideal situation can be approximated by
systems which are thermally insulated from their surroundings or processes which take place so
rapidly that there is not enough time for heat transfer to occur.
5. Free expansion. This is an adiabatic process in which there is no work done on the system or by
the system.
Returning to equation 150, we nd that if the pressure remains constant, the integral becomes
W =
_
V2
V1
p dV
W = p(V
2
V
1
) (151)
This equation immediately suggests that the total work done in an isovolumetric process is zero. (It also
suggests that if the volume changes, but returns to its original volume, the work done will be zero. This
is in fact, false in general).
In the case where the volume of the gas changes along with the temperature, we can use equation 150
to nd the work done. We rst need to rewrite the pressure in terms of the volume and temperature using
the ideal gas law: p =
nRT
V
. Then,
W =
_
V2
V1
nRT
V
dV
W = nRTln
V
2
V
1
(152)
8.7 The Ideal Gas: Macroscopic Properties of a system.
We have already indicated that a macroscopic system can be described by the state variables p, V, T and
n. We will write down an equation of state for an ideal gas, which is a gas at low density (i.e., at low
pressure and high temperature). We have
pV = nRT (153)
where n is the number of moles, and R is a the ideal gas constant, with a value 8.314 J mol
1
K
1
. We
see immediately that for a process in which the number of moles is constant, we can write the equation
between two states, 1 and 2, as
p
1
V
1
T
1
=
p
2
V
2
T
2
(154)
54
Equation 154 is called the ideal gas laws. When the pressure of the gas is held constant through a process,
the volume of a xed mass of gas is related to the temperature through Charles law:
V
1
T
1
=
V
2
T
2
(155)
When the temperature and mass of the gas are kept constant, we have Boyles law:
p
1
V
1
= p
2
V
2
(156)
The state of the gas is completely specied by a knowledge of the state variables at the time of
measurement, and does not depend on the past history of the gas.
A continuous change in the state of a gas is called a process. Graphically, a process is indicated by
a line or a curve going from a certain initial state to a nal state. Such curves are drawn with p on the
ordinate and V on the abscissa and are called pV diagrams.
A process in which the gas is returned to its original state is called a cycle.
8.7.1 Standard temperature and pressure (STP)
As a reference standard, a gas at a temperature T = 273.15 K (0

)C and a pressure p = 1.013 10


5
Pa (which is the atmospheric pressure at the earths surface) is said to be at standard temperature and
pressure, S.T.P. Under such conditions, one mole of an ideal gas occupies a volume
V =
RT
V
V =
8.314 273.15
1.013 10
5
V = 2.24 10
2
m
3
(157)
9 The First law of Thermodynamics
We will now relate some of the quantities that we have dened to the quantity of heat energy transferred,
Q.
Let us rst suppose that we can dene an initial energy state U
i
of the ideal gas, and call this the
internal energy of the gas. Then, we expect that if energy is transferred to the gas, the internal energy
is likely to change to a certain nal value U
f
. Suppose also that the gas then uses some of this internal
energy to do work W (by expanding). Then, the internal energy will have increased as a result of the heat
supplied Q but will decrease as a result of the work done. We write this as
U
f
U
i
= QW
or Q = U +W (158)
In the form 158, we see that the heat supplied goes to change the internal energy of the system, and also
cause the system to do work.
In the case where we have a cyclic process that returns the ideal gas to its original state, U
f
= U
i
and
therefore W = Q.
For an isolated system, there is no transfer of heat and therefore Q = 0; in addition, our denition is
such that an isolated system does no work on its surroundings; W = 0. This means that
U
f
U
i
= 0
U
f
= U
i
(159)
55
i.e., the internal energy of an isolated system is constant.
We have seen already that in an adiabatic process, the system does not lose or gain heat energy.
Thus, Q = 0 and in an adiabatic process the change in internal energy is converted entirely into work,
U = W.
Let us rewrite the rst law of thermodynamics in terms of small changes in energy dQ:
dQ = dU + dW
dQ = dU +p dV (160)
Now, we will make an important addition to the ideal gas law: The internal energy of a gas depends
only on its temperature.
We now establish a relationship between the heat capacities for an ideal gas. We earlier dened C
V
and C
p
. Let us rst consider a constant volume process. Then by the rst law of thermodynamics,
dQ = p dV + U
dQ = U for a constant volume process. (161)
But we recall that, for the molar heat capacity,
dQ
dT
= nC
V
dQ = nC
V
dT (162)
So, U = nC
V
dT for a constant volume process.
For a constant pressure process,
dQ = pdV + U
nC
p
dT = p(V
2
V
1
) + U
nC
p
dT = p
_
nRT
2
p

nRT
1
p
_
+ U
nC
p
dT = nRdT + U (163)
But, the change in internal energy does not depend on the nature of the thermodynamic process. There-
fore, U = nC
V
dT can be used in equation 163:
nC
p
dT = nRdT +nC
V
dT
C
p
= R +C
V
(164)
This gives us the relation between the molar heat capacities. We go further to dene the ratio between
C
p
and C
V
as :
=
C
p
C
V
(165)
For an adiabatic process, the equation of state is given by
pV

= constant (166)
C
V
is found to depend on the nature of the gas. It has a constant value for monatomic gases, diatomic
gases, and so on. Well see more of this later.
56
9.1 The Microscopic description and the kinetic theory of gases.
We will assume that all the systems we are interested in are made of atoms and molecules (which is
perfectly reasonable, seeing as matter is made up of atoms and molecules). As earlier mentioned, a
microscopic description entails the use of variables such as the kinetic energy, speeds, and momenta of
the individual particles that make up a system. In this regard, we shall proceed to establish the ideal gas
law from an entirely microscopic description. We start off with our ideal gas within a cubical container
with smooth walls and volume l
3
. Next, we set down the following assumptions:
1. The molecules of a gas are separated by distances that are large compared with the size of the
molecules.
2. Collisions between molecules and between molecules and the walls of the container are perfectly
elastic.
3. The only forces that act on the molecules occur during collisons.
4. The duration of a collision is very short compared with the time between collisions.
5. Molecules obey Newtons laws of motion (i.e., they are subject to the laws of classical mechanics).
Now, let us consider the molecules travelling in only one direction out of the three possible co-
ordinates. When these molecules strike the wall with speed v, they rebound with the same speed (since
the collison was elastic) and therefore exert an average force F
av
on the walls:
F
av
=
2mv
t
(167)
The molecules are bouncing back and forth between the walls with speed v =
2l
t
. Equation 167 becomes
F
av
=
2mv v
2l
F
av
=
mv
2
l
(168)
Now, the pressure exerted on the wall is
p =
F
av
A
p =
mv
2
Al
p =
mv
2
V
(169)
We assume further that a third of the total number of molecules N is moving in the chosen direction.
So, the total pressure on the wall will be
p =
N
3
mv
2
V
pV =
N
3
mv
2
(170)
All the molecules are, however, not moving with the same velocity. Indeed, the range of velocities
is given by the Maxwell-Boltzmann distribution, as we shall see. At this time, we will introduce the
57
root-mean-square (or r.m.s) speed, which is a way of averaging the different speeds of the molecules.
We have
v
rms
=
_

v
2
(171)
So, the r.m.s value is the square root of the mean of the squares of the different molecular speeds. The
r.m.s value gives a better estimate of the velocity. Thus, equation 170 is better written as
pV =
1
3
Nmv
2
rms
(172)
Note that since
Nm
V
is simply the density of the gas, we can write equation 172 as
p =
1
3
v
2
rms
(173)
Comparing the equation of state for an ideal gas (pV = nRT) and equation 172 gives us a neat
expression for the average kinetic energy. In the following, we use the fact that the number of moles
n =
N
NA
, where N is the number of molecules and N
A
is Avogadros number, as per denition.
nRT =
1
3
Nmv
2
rms
N
N
A
RT =
1
3
Nmv
2
rms
mv
rms
=
3RT
N
A
1
2
mv
2
rms
=
3
2
_
R
N
A
_
T
1
2
m

v
2
=
3
2
_
R
N
A
_
T
K.E
av
=
3
2
_
R
N
A
_
T (174)
Boltzmanns constant k is dened as k =
R
NA
, so that equation 174 becomes
K.E
av
=
3
2
kT (175)
Boltzmanns constant is k = 1.38 10
23
J K
1
.
We now dene the internal energy of the gas as the sum of all the kinetic energies of the particles.
Since there are N particles, this gives
U =
3
2
NkT
or U =
3
2
nRT (176)
9.1.1 Mean free path.
Since the molecules make collisions with each other but are otherwise unimpeded in their motion (except
for when they collide with the walls of teh container), we would like to nd out how far a molcule can
travel before making another collision. This distance, averaged over many molecules, is called the mean
58
free path. Another quantity of interest is the collision frequency, which is the number of collisions made
per unit time. For a number of molecules N in a volume V , the formulae of interest are:
Mean free path =
V
4

2r
2
N
(177)
Collision frequency f =
4

2r
2
vN
V
(178)
9.1.2 Equipartition of energy.
We make a statement of the equipartition of energy as follows: For a large number of molecules at
temperature T, the total energy of the gas is distributed among all the molecules such that, for each
molecule, there is an average energy of
1
2
kT for each degree of freedom.
What is meant by a degree of freedom? Consider a single atom or molecule. It is free to move in
any of the three co-ordinate directions, and therefore we ascribe to it three degress of freedom. Since each
degree of freedom is worth energy
1
2
kT, we see that the gas has energy
3
2
kT per molecule. What about
a diatomic molecule? We will consider such a molecule as having two atoms bound together. There will
then be three possible ways in which this molecule can rotate. However, rotation about the line joining
the atoms will have negligible energy contribution to the system, and so we will admit two degrees of
freedom for rotation. So, the total energy per molecule due to rotation is kT.
We also will consider vibrations of each molecule. Now, as the vibrations occur, there are changes
in both the elastic potential energy and the kinetic energy. We will ascribe to each of these a degree of
freedom, such that for vibrations as well we will have a total of kT energy per molecule.
Table 1 summarises the foregoing.
Object Translation Rotation Vibration Total Energy U
Monatomic gas
3
2
kT 0 0
3
2
kT
3
2
nRT
Diatomic gas
3
2
kT
1
2
kT +
1
2
kT 0
5
2
kT
5
2
nRT
Diatomic gas
3
2
kT
1
2
kT +
1
2
kT
1
2
kT +
1
2
kT
7
2
kT
7
2
nRT
Table 1: Equipartion of energy
9.1.3 The Dulong and Petit law
We earlier on discussed the fact that a solid can be modelled as a crystal in which the atoms are bound
by forces that allow for vibrations. If each atom is allowed to vibrate in any of the three co-ordinate
directions, then we will have, a total energy of
3
2
kT per atom. But, each vibration has two components,
the potential energy and kinetic energy, as we saw in the previous section. We may then sum over all the
atoms to get the total energy:
U = N 2
3
2
kT
U = 3NkT
U = 3N
R
N
A
T
U = 3nRT (179)
59
The molar heat capacity is given by
C =
1
n
dU
dT
C = 3R (180)
Equation 180 is the Dulong and Petit law, which is veried by experiments at roomtemperature. However,
the explanation offered for (and the derivation of) the Dulong and Petit law is wrong. For one thing, we
notice that the contributions of the electrons in the solid were ignored, despite the fact that we should
expect, from our earlier discussion, that each electron also possess energy
1
2
kT per degree of freedom. A
proper explanation of what is happening here requires quantum mechanics, and therefore, well defer it
for the nonce.
9.1.4 The Maxwellian distribution of velocities.
We previously stated that all the molecules in the gas are not moving at the same velocities, and we
discussed the r.m.s speed in this regard. The velocity distribution of the molecules is described by the
Maxwell-Boltzmann function f(v), depicted in gure 33 for a given temperature T.
Figure 33: Maxwellian distribution of velocities.
We nd that, apart from the r.m.s value, we have the most probable velocity v
mp
and the average
velocity v
av
. These are derived from the function, but we will just state the results here:
Most probable speed v
mp
=
_
2kT
m
(181)
Average speed v
av
=
_
8kT
m
(182)
Root-mean-square speed v
rms
=
_
3kT
m
(183)
Note that we can use the relation molar mass (M) = mass (m) N
A
to rewrite the above equations.
For example, the r.m.s speed becomes.
v
rms
=
_
3RT
M
(184)
9.2 The Molecular properties of Matter.
We provide here a very basic denition of solids, liquids, and gases in terms of a molecular substructure.
In solids, the molecules are xed within a crystal lattice (for crystalline solids) by bonds that are quite
60
strong. The molecules have no freedom of movement through the lattice, but can vibrate about their
positions.
In a liquid, the molecules are free to move, being held by much weaker bonds than in the solid.
However, the molecules are close to each other, and this strongly inuences their movement.
In a gas, the molecules are widely separated and exert negligible forces on each other. The molecules
are free to move and do so until they collide with another molecule or the walls of the container, after
which they continue moving.
9.3 Transfer of heat
Heat enery may be transferred in any one or more of the following ways: conduction, convection, and
radiation.
When the system under consideration is a solid, heat energy transfer may occur as a result of two
mechanisms: (a) the transfer of energy as a result of increased vibrations of the molecules and (b) the
transfer of heat energy by the movement of electrons from the region of higher temperature to the region
of lower temperature. Item (b) would obviously only be relevant in materials that have large amounts of
free electrons available such as metallic conductors.
For uids in general, heat transfer involves the actual movement of molecules fromhigher temperature
regions to lower temperature regions.
Heat transfer by radiation is different from the other two in that it does not involve matter. Heat is
transferred as an electromagnetic wave, and can therefore travel through a vacuum and suffer reection,
refraction, and other optical phenomena.
61
10 Appendix A
A summary of some Mathematical Concepts
Trignometric Identities
cos(A +B) = cos Acos B sin Asin B
cos(A B) = cos Acos B + sin Asin B
sin(A +B) = sinAcos B + sin Acos B
sin(A B) = sinAcos B sin Acos B
sin
2
+ cos
2
= 1
sec =
1
cos
csc =
1
sin
cot =
1
tan
sec
2
= 1 + tan
2

Quadratic Equations
A quadratic equation is of the form
ax
2
+bx +c = 0
where a, b, c are constants and x is a variable. Solving for x gives two values determined by the sign of
the square root. If the number under the square root is negative, then the equation has no real solution.
x =
b

b
2
4ac
2a
Appendix B
Some Physical Constants
Acceleration due to gravity, g = 9.81ms
2
Universal Gravitational constant, G = 6.67 10
11
Nm
2
kg
2
Further Reading
1. C J Tranter, C G Lambe. Advanced Level Mathematics (Pure and Applied), 4th Ed. Hodder and
Stoughton (1979).
2. Richard T Weidner and Robert L Sells. Elementary Classical Physics, Vol.1. Allyn and Bacon
(1965).
3. M R Spiegel. Theory and Problems of Advanced Calculus McGraw Hill (1974)
62
4. R A Serway, J S Faughn. College Physics. Harcourt Publishers (1999).
5. Raymond ASerway and John WJewett, Jr. Principles of Physics, Fourth Edition, Vol. 1. Brookes/Cole
Thomson Learning (2006).
63

Vous aimerez peut-être aussi