Sie sind auf Seite 1von 81

Introductory Physics

Mechanics & Waves

Itay Yavin

Phys1B03 Website
http://www.physics.mcmaster.ca/phys1b03

Section C03 Website


http://www.physics.mcmaster.ca/~yavin/yavin/courses/phys1b03/

1
c 2011 by Dr. Itay Yavin,
Copyright –

All Rights Reserved.


Contents

1 Introduction 5

2 Kinematics 6

2.1 Motion in one dimension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2.2 Simple motion in one dimension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.3 Motion in two dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2.4 Simple motion in two dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

3 Forces 17

3.1 Newton’s Laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

3.2 Simple Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3.3 Friction and Drag . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

4 Energy 26

4.1 Conservative Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

4.2 Energy Conservation and Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

4.3 Work and Potential Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

4.4 Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

5 Momentum 34

5.1 Momentum Conservation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

5.2 Elastic collisions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

5.3 Explosions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

6 Simple Harmonic Motion 40

6.1 Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

2
6.2 The Pendulum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

6.3 Damped Harmonic Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

6.4 Driven Oscillations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

7 Waves 51

7.1 Characteristics of Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

7.2 Sinusoidal Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

7.3 Superposition of Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

7.4 Standing Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

7.5 Interference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

7.6 Reflection and Transmission Through a Boundary . . . . . . . . . . . . . . . . . . . 65

8 Fluids 71

8.1 Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

8.2 Incompressible Fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

8.3 Buoyancy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

8.4 Continuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

8.5 Fluid Statics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

8.6 Fluid Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

A Mathematical Appendix - Trigonometry 79

B Mathematical Appendix - Calculus 80

B.1 Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

B.2 Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

B.3 Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

3
1 Introduction

The laws of physics are simple. Their application and usage are not.

In these notes, I have attempted to state the laws as succinctly as I could without too much
interpretation and repetition. True understanding of these laws, their application, their usage,
and ultimately their limitation, is found only through repeated examination of different physical
systems. Together, we will analyze many phenomena from the natural world in order to gain
additional insights into the workings of the laws of physics. Separately, I hope you each spend most
of your time studying for this course by solving the problems I will suggest or other physics puzzles
you might encounter.

I urge you to solve as many problems as you can find the time for. Keep in mind that the notes
are not designed to teach problem solving techniques. We will go over problems together in class,
in tutorials, and in office hours so I encourage you to participate as much as possible. The notes
really only contain the principles of physics without much reference to their usage in problems.

Many of the sections are followed by a few problems. I hope these problems will help strengthen
the concepts learned and provide concrete examples for the notions discussed in the section itself.
It might be a good idea to try them on your own to test your understanding of the section. These
problems are not too simple so don’t get discouraged if it takes you a little longer to solve them.
There is no better way to learn physics than by solving problems by yourself in order to develop
your own way of understanding and interpreting the laws of physics.

The text we will use, Schaum’s Outline of College Physics, 11th or 10th edition by Bueche and
Hecht, contains many problems with detailed solutions. If you find that you would like another
textbook to consult with for additional explanations here are a few possibilities,

• Physics for Scientists and Engineers, by R.A. Serway and J. W. Jewett

• Physics: Principles with Applications, by D. C. Giancoli

• Fundamentals of Physics, by D. Halliday, R. Resnick, and J. Walker

The first one, by Serway and Jewett, is the one used by other sections of this course. These books
typically have several editions that differ from each other by trivial changes. So it does not really
matter which edition you pick, the physics is the same.

Simplicity is beautiful and powerful. But, the true appreciation of Nature’s majestic simplicity
requires frequent encounters with Her and under different circumstances so let us start.

4
2 Kinematics

Much of physics has emerged out of the investigation of motion. Before we understand why things
move the way they do, we will first develop the language necessary to describe motion, namely
kinematics. In what follows we will concentrate on the motion of point-like objects, an idealization
that proves powerfully useful in much of what we will encounter. We begin with motion in one
dimension.

2.1 Motion in one dimension

The motion of a particle in one dimension is described by a single function which we will denote by
x(t). The symbol t refers to time, which we can measure with a clock relative to some initial time.
The symbol x refers to the position of the particle, which we can measure with a ruler relative to
some point that we define as the origin. What we define as the initial time and the spatial origin
is arbitrary and just serves to define the coordinate system.

If we know the position function x(t), then we know everything about the motion of the particle.
The velocity of the particle is defined as the derivative of its position,
dx(t)
v(t) = . (2.1)
dt
It is also a function, describing how fast and in which direction the particle is moving at any
moment. It is sometimes known as the instantaneous velocity since it yields the velocity at every
instant of time. The magnitude of the velocity —v(t)— is known as the speed. It is also possible
to define the instantaneous acceleration of the particle, which is the rate of change of its velocity,
dv(t) d2 x(t)
a(t) = = . (2.2)
dt dt2
Together, the position, velocity, and acceleration form the most important aspects of the motion
of particles.

Knowing the velocity of the particle you can reconstruct its position by integrating the velocity
function,
Z
x(t) = v(t)dt . (2.3)

The constant of integration that appears when effecting the indefinite integral corresponds to the
need to specify the particle’s position at some instance of time in order to completely specify its
position. Similarly, the velocity itself can be found by integrating the acceleration,
Z
v(t) = a(t)dt . (2.4)

A note about units: it is very important to keep in mind that time, position, velocity, and accel-
eration all have different units. Time is typically measured in seconds, but we will also encounter

5
minutes, hours, days, years, and etc. Position is typically measured in meters, but centimetres,
millimetres, kilometres, and etc. are also used. From its definition, velocity is measured in meters
per second. Similarly, acceleration is measured in meters per second per second (or equivalently in
meters per second squared).

6
Problems

1. The motion of a spring about its equilibrium position is given by x(t) = A cos(t/T ) where
A = 1m, T = 1s.

(a) Plot the spring’s motion.


(b) Find its velocity as a function of time.
(c) Find its acceleration as a function of time.

* We will meet this system again later in the course and show that this is indeed the correct
motion.

2. The acceleration of a ball you throw vertically upwards is given by a(t) = −g, where g =
9.82m/s2 . We will define the moment you throw it upwards as the initial time t = 0, and its
position at that moment as the spatial origin. Its initial velocity is v0 = 10m/s.

(a) Find its velocity function, v(t)


(b) Find its position function, x(t)

* The positive x-axis here is defined as the upwards vertical direction. Which way is the
acceleration pointing?

3. The velocity of an object moving through a fluid slowly is given by v(t) = v0 exp(−t/T ) where
v0 = 1mm/s is its initial velocity and T = 1s is some characteristic timescale for the motion
of that object in the fluid. We will define the origin, x = 0, as the object’s position at t = 0.

(a) Find its acceleration, a(t). Is it accelerating or decelerating?


(b) Find its position function, x(t).
(c) What is the asymptotic distance of the object from the origin as time→ ∞.

7
2.2 Simple motion in one dimension

There are two special cases of the general motion, which are particularly simple. They are motion
with constant velocity and motion with constant acceleration. Despite their simplicity, these two
cases play an important role in many physical situations.

We begin with the case of a constant velocity, v(t) = vc , where vc is a constant with the units of
velocity. Using the general relation between the velocity function and the acceleration, Eq. (2.2),
we immediately find the result,
dv(t)
a(t) = =0. (2.5)
dt
That is to say, the body neither accelerates nor decelerates. The position function is found through
Eq. (2.3) by effecting a simple integration,
Z
x(t) = vc dt = vc t + x0 , (2.6)

where x0 is an integration constant. Its physical significance is immediately apparent since x(0) = x0
is the initial position of the body along the x-axis. The position function x(t) is a general linear
function of time. The slope is the velocity and the intercept with the x-axis is the initial position.

Let us move on now to consider the second special case of uniformly accelerated motion, namely
a(t) = ac , where ac is a constant with the units of acceleration. Using the relation between the
acceleration and velocity, Eq. (2.4), we have
Z
v(t) = ac dt = ac t + v0 , (2.7)

where v0 is an integration constant that represents the initial velocity. The position function is
recovered with yet another integration,
1
Z
x(t) = v(t)dt = ac t2 + v0 t + x0 , (2.8)
2
where x0 is a second integration constant with the same meaning as before. As I emphasized above,
this last equation describing the position as a function of time completely specifies the state of the
system at every instance. However, it requires knowledge of the time and hence some measuring
device in the form of a clock. There are situations where we do not want to or cannot measure
the time that elapsed between two instances, but we can measure the velocity and position, or
the velocity and acceleration, or any other two kinematic functions. Then we can express the
kinematics without reference to time. In the case of uniformly accelerated motion it is particularly
simple to do so. From Eq. (2.7) we solve for the time1 ,
t = (v − v0 )/ac . (2.9)
This can be used in the position function, Eq. (2.8), to yield,
1 v 2 − v02
x= + x0 . (2.10)
2 ac
1
You can now think about the time as a function of the velocity, t(v). The time does not have to be the independent
variable.

8
Problems

1. Plot a graph of the position (ordinate) versus the time (abscissa) for the following motions:

(a) constant velocity of magnitude 3.14m/s to the right, starting at x = 0.


(b) constant acceleration of magnitude 9.81m/s2 pointing downwards, starting 2m above
ground.
(c) light bouncing between two mirrors separated by 1m starting at the leftmost mirror.

2. The speed of light is a fundamental quantity of nature. The fact that it is finite was experi-
mentally established in the 17th century by Ole Romer (this fact deflated many philosophical
musings dating as far back as the 4th century BC and possibly earlier). This ingenious ex-
periment consisted of observing the change in the periodicity of Io, one of Jupiter’s moons,
over the course of a year as the Earth moves around the Sun. Without going into the details
of the observation, Romer managed to estimate that it takes light about 22 minutes to cross
the diameter of the Earth’s orbit around the Sun.

(a) What would he conclude is the speed of light?


(b) How does it compare with the 20th century established value of c = 299, 792, 458m/s ?
(c) Using the established value for the speed of light (not Romer’s estimate), calculate how
long does it take for a signal from a GPS satellite to reach the receiver. GPS satellites
are at a distance of 26560km from the Earth’s centre.

3. We can rearrange Eq. (2.10) to read,

1 v 2 − v02
D ≡ x − x0 = , (2.11)
2 ac
where D is known as the displacement, which is basically the distance covered by the body
from its initial position. Note that the displacement grows quadratically with the velocity.
This seemingly innocuous fact has important ramifications for your stopping abilities when
driving your car too fast.

(a) If you slam on your brakes at 90 km/hr you will come to a full stop after about 70 meters.
Assuming that your brake provide a constant acceleration (not a bad assumption), find
that acceleration.
(b) How does this acceleration compare to near Earth gravitational acceleration, g = 9.81m/s2 ?
Is your answer reasonable?
(c) What will be your stopping distance if you drove at 120 km/hr?

4. Think of an experiment to estimate the human reaction time.

9
2.3 Motion in two dimensions

If a body experiences no acceleration than the velocity remains constant both in magnitude and
direction. In that case we can describe the motion as one-dimensional as we did above. If a body
does experience acceleration, but the acceleration is along the direction of the motion then this is
still a one-dimensional motion. The magnitude of the velocity will change, but not its direction
and so we can simply choose the x-axis along the direction of the motion. If the body accelerates
and the acceleration is not in the same direction as the velocity, but the two vectors form a plane
then the resulting motion is a two-dimensional motion.

The motion of a body in two dimensions is described by two position functions, say x(t) and y(t),
for each of the dimensions. For motion in the plane, it is often convenient to choose a cartesian
coordinates system with two perpendicular fixed axes. The origin of the coordinate system is then
defined by the intersection of the axes.

The position of a point-like body is then given by the ordered pair (x(t), y(t)) or as the vector,

~r(t) = x(t)î + y(t)ĵ , (2.12)

where î and ĵ are unit vectors pointing in the positive x and y axes, respectively. The vector
notation is useful since it makes it clear that the two directions, î and ĵ, are independent. As in
the case of motion in one dimension, the velocity of a body is the rate of change of its position and
is defined as,

d~r(t)
~v (t) = . (2.13)
dt
In the case of the cartesian coordinates which we employ this is simply,

~v (t) = vx (t)î + vy (t)ĵ , (2.14)

where,

dx(t)
vx (t) = , (2.15)
dt
dy(t)
vy (t) = . (2.16)
dt
Similarly, the acceleration is the rate of change of the velocity,

d~v (t) d2~r(t)


~a(t) = = , (2.17)
dt dt2

or in components ~a(t) = ax (t)î + ay (t)ĵ with

dvx (t)
ax (t) = , (2.18)
dt
dvy (t)
ay (t) = . (2.19)
dt

10
It is often more convenient to denote a vector with the use of brackets as in,

~v = (vx , vy ) , (2.20)

instead of

~v = vx î + vy ĵ . (2.21)

The two notations mean the same thing and can be used interchangeably as convenient.

11
Problems

1. A particle is moving with a velocity ~v1 (t) = (0, 2)m/s starting at t = 0 somewhere along the
x-axis. A second particle is moving with a velocity ~v2 (t) = (−1, 2)m/s also starts along the
x-axis, but a distance of 10 meters to the right of the first particle.

(a) Will the particles collide? prove it.


(b) How far from the first particle’s original position will the collision happen?
(c) How far from the second particle’s original position will the collision happen?

2. The velocity vector of a particle is given by ~v (t) = (− sin(t), cos(t)) with time being measured
in seconds and velocity in m/sec. The particle is initially found on the x-axis at x=1 m.

(a) Find the position function, ~r(t).


(b) Draw the vector connecting the origin to the particles position when t = π/4, π/2, 3π/4,
π, 5π/4, 3π/2, 7π/2, and 2π.
(c) What kind of motion is the particle executing?
(d) Find the acceleration function, ~a(t).
(e) Draw the acceleration vector ~a(t) with its origin on the particle at the same times as
above.

12
2.4 Simple motion in two dimensions

Motion in a circle

A very important type of motion is that of a body in circular motion with a constant speed. By
choosing the origin of time appropriately, such motion can always be described by,
~r(t) = (R cos(2πt/T ), R sin(2πt/T )) , (2.22)
where R is the radius of the circle, and T is the period of the motion. By differentiating once we
obtain the velocity,
~v (t) = (−V sin(2πt/T ), V cos(2πt/T )) , (2.23)
where V = 2πR/T as can be expected by consideration of the geometry. We note that the velocity
vector is perpendicular to the position vector everywhere along the path. Also, its magnitude or
speed is constant,
q
|~v (t)| = V 2 sin2 (2πt/T ) + V 2 cos2 (2πt/T ) = V . (2.24)
In this last step I made use of the basic identity of trigonometry.

Affecting a second differentiation we deduce the acceleration of the particle,


~a(t) = (−a cos(2πt/T ), −a sin(2πt/T )) , (2.25)
with a = 2πV /T . The magnitude of the acceleration is also constant and is equal to a. The
acceleration is always pointing towards the centre of the circle, perpendicular to the velocity vector
which is always tangential to the circle. By considering the definition of V and a we deduce the
following simple relation between them,
V2
a= . (2.26)
R
This is known as centripetal acceleration. It is always pointing towards the center of the circle and
its magnitude is given by Eq. (2.26). It is necessarily present when a body follows a circular path
since this acceleration is responsible for changing the direction of the velocity along the path.

Projectile Motion

Another very important type of motion, both for its practical application and theoretical simplicity,
is that of a projectile. As a first approximation we will neglect the effects of air resistance and
consider only the acceleration due to the Earth’s gravity. Choosing a coordinate system where the
x-axis is associated with the horizontal direction and the y-axis with the vertical we can write the
two components of the acceleration vector,
ax (t) = 0 , (2.27)
ay (t) = −g . (2.28)

13
where the minus indicates that I have chosen the upwards direction as the positive axis (recall that
gravity causes objects to accelerate downwards towards the Earth). There are no forces acting
horizontally and so ax = 0. Since the acceleration is the derivative of the velocity, the velocity is
the anti-derivative of the acceleration and so we have,
Z
vx (t) = ax (t)dt = vx0 , (2.29)
Z
vy (t) = ay (t)dt = vy0 − gt , (2.30)

where vx0 , and vy0 are the initial velocities along the x-axis and y-axis respectively. Since there is
no acceleration in the x-axis the velocity remains constant. On the other hand, in the y-axis the
body is experiencing an acceleration downwards which causes the velocity to tend towards large
negative values.

The position of the body is obtained through an additional integration,


Z
x(t) = vx (t)dt = x0 + vx0 t , (2.31)
Z
y(t) = vy (t)dt = y0 + vy0 t − 21 gt2 , (2.32)

where x0 and y0 are the initial positions of the body along the x-axis and y-axis, respectively.
Projectiles are typically shot from the ground upwards at some speed with a certain angle with
respect to the ground. So, let us choose x0 = 0 and y0 = 0 as our initial position (origin of the
coordinate system). Let us also choose the initial projectile to have an initial velocity v0 with an
angle θ with respect to the ground. The initial velocity is therefore,

vx0 = v cos θ , (2.33)


vy0 = v sin θ . (2.34)

With these relations in mind, the equations describing the projectile’s motion are now particularly
simple,

x(t) = (v cos θ)t , (2.35)


1 2
y(t) = (v sin θ)t − 2 gt . (2.36)

14
Problems

1. Imagine yourself driving your car heading North. At a certain point you exit East on a ramp
that bends as a quarter circle of radius 200 m. It takes you 10 s to complete the turn.

(a) How fast are you moving?


(b) What is the centripetal acceleration of the car as you take the turn?
(c) Suppose you slow down to 60 km/hr, how long does it take you to complete the turn
and what is the centripetal acceleration in this case?

* This problem embodies the danger of taking turns. The time it takes to complete a turn
only decreases linearly with speed. But, the acceleration, and therefore as we will learn below
also the forces on the car, increases quadratically with the speed.

2. Astronaut Harrison Schmitt from the Apollo 17 mission to the moon threw his hammer on
the moon. From the footage, I guesstimated that the hammer was thrown from an initial
height of 1.5 m off the ground with an initial velocity of 10m/s at roughly 45◦ . Since there
is no air resistance on the moon, the only force acting on the hammer is the moon’s gravity
pulling the hammer down at a constant acceleration of amoon = 1.6m/s2 .

(a) What is the hammer’s initial speed in the horizontal and vertical directions?
(b) What is the maximum height the hammer reaches?
(c) How far from astronaut Schmitt did the hammer land?

15
3 Forces

When studying kinematics we assumed a position function, x(t), and derived related quantities
such as the velocity, v(t), and the acceleration, a(t). However, the real business of mechanics is
the determination of x(t) or its derivatives from basic principles. These principles, or the laws of
mechanics, can be stated in several different ways, but in this course we will concentrate on their
expression in terms of forces.

3.1 Newton’s Laws

The effect of forces on the motion of a body is embodied in Newton’s second law of motion,

F~tot = m ~a , (3.1)

where ~a is the acceleration of the body, m is its inertial mass, and F~tot is the total force acting on
it. The standard unit for measuring forces is called a Newton,

1 N ≡ kg × m/s2 . (3.2)

The force, F~tot , is in general a function of all the kinematical quantities as well as the time. By
itself this law does not provide us with enough information to determine the motion since the force
is left unspecified. Over the last several centuries, experimentation, guesswork, and insight have led
to a comprehensive understanding of the forces present in different physical situations. We shall
encounter several of them in what follows. It is important to recognize that often several distinct
forces are acting concurrently on a body. In this case one must consider their vectorial sum as
affecting the motion and the total force is given by,

F~tot = F~1 + F~2 + . . . = m ~a . (3.3)

When forces are absent, namely F~ = 0, a body is left unperturbed and Newton’s second law then
asserts that its acceleration also vanishes. This immediately leads to Galileo’s principle of inertia,
namely that a body’s velocity remains constant in the absence of any external forces. This is also
known as Newton’s first law of motion. When the net force on a body vanishes, it is sometimes
said that the body is in equilibrium. A distinction is then drawn between static equilibrium, which
refers to situations where the net force is zero and the body is at rest, and dynamic equilibrium,
which refers to situations where the net force is zero and the body is moving. This distinction is
important in cases where the environment exerts forces on the body, such as friction and drag, that
depend on its velocity.

Now consider two separate bodies that exert forces on each other. Newton’s third law of motion
states that the force exerted by the first body on the second is equal in magnitude, but opposite
in direction to the force exerted by the second body on the first. Symbolically we can write,

F~(12) = −F~(21) , (3.4)

16
where the subscript (12) designates the force exerted by the first body on the second. Similarly,
(21) designates the force exerted by the second body on the first. Importantly, these two forces,
F~(12) and F~(21) are not acting on the same body. F~(12) is acting on the second body, whereas F~(21)
is acting on the first one.

When working out the dynamics of a system, it is important to identify the different bodies involved.
It is often helpful to draw a separate diagram for each body and indicate the forces acting on that
body alone. Such diagrams, drawn for each body separately depicting the forces that act on it, are
known as force diagrams. It is essential to make sure that Netwon’s third law is respected when
there are internal forces between the different bodies. If two bodies, A and B, are exerting forces
on each other, then a force is drawn as acting on body A (in the diagram depicting body A), and
an equal but opposite force must be drawn as acting on body B (in the diagram depicting body
B).

Newton’s second law applies to each body separately and it can be read off from the diagrams.
One sums up the different forces acting on a body and write the LHS of Newton’s second law. The
RHS is simply the body’s mass and acceleration. Written explicitly this equation is known as the
equation of motion. The dynamics of the system is then obtained by considering the equations of
motion associated with all the bodies together.

17
Problems

1. Two bodies with masses m1 and m2 are placed next to each other on an inclined plane that
is at an angle θ with the horizon, as shown in the figure.

(a) Draw the force diagrams assuming the experiment is done on the surface of the Earh
and that friction with the plane is present.
(b) Write down equations of motion for each body.
(c) Solve for the acceleration.

2. A body of mass m1 is sitting on top of another body of mass m2 , which is resting on a


frictionless table. One pushes on the lower body of mass m2 with a force F to the right.
The force of friction between the two bodies prevents the upper one from slipping and so it
accelerates together with the lower one.

(a) Draw only the horizontal forces acting on the bodies


(b) What is the resulting horizontal acceleration?

* Note: Not the precise nature of the friction force nor its relation to the vertical forces acting
on the bodies matter for this problem.

3. The force on a particle moving through a fluid at low velocities is proportional to its velocity,
but opposite in direction (as to oppose the motion). We shall consider only 1-dimensional
motion and write the force as,

F = −b v(t) , (3.5)

where b is some constant of proportionality with dimensions of kg/s.

(a) Write down the equation of motion for a particle of mass m


(b) Show that v(t) = v0 exp(−mt/b) is a solution to this equation.

18
3.2 Simple Forces

Near Earth Gravity

Near to the surface of the earth, a situation which applies to every day life, objects are attracted
to the Earth with a constant force,

F = −mg , (3.6)

where m is known as the gravitational mass, g = 9.8m/s2 , and the minus sign is a matter of
choosing the upwards direction as positive. The force an object feels due to the local gravitational
force is known as the object’s weight. Notice the sharp distinction between the gravitational mass
of an object, which is an intrinsic quantity, and the weight of an object which depends on the local
gravitational field, which is an extrinsic quantity that is equal to g = 9.8 m/s2 in the case of the
Earth. On the moon, the local gravitational field is much weaker and gmoon = 0.166 g = 1.63 m/s2 .
So, on the moon your weight is about 1/6 of what it would be on the Earth, but your mass remains
the same.

A very deep and remarkable observation is that the gravitational mass is always equal to the inertial
mass. We have never found an exception to this rule, and by applying Newton’s second law it allows
us to conclude that,

− mg = ma → a = −g . (3.7)

The acceleration is therefore constant and is pointing downwards. Throughout this course we will
take g to be a constant. This is a good approximation for our purposes, but you should know
that g in fact varies on the Earth’s surface due to the non-sphericity of the Earth, local density
inhomogeneities (important for oil explorations), and other effects. Ultimately, the force law itself is
only a good approximation near the surface of the Earth. At higher altitudes and further away from
the Earth’s centre gravity is more accurately described by Newton’s law of universal gravitation.

Hooke’s Law

When pulling on an elastic material, such as a spring, so as to stretch it, you have undoubtedly felt
a force pulling back on you in an attempt to restore the material back to its equilibrium position.
For many materials, the dependence of this restoring force on the displacement from equilibrium is
known to follow Hooke’s law, namely,

F = −kx , (3.8)

where x denotes the displacement away from equilibrium, k is a constant with units of N/m and is
known as the spring constant, and F is the force. The minus sign represents the fact that the force
opposes the stretch. Stretch the spring to the right and the force on your hand will be to the left.
Stretch the spring to the left and the force on your hand will be to the right.

19
Normal Forces

When a solid object A presses against another solid object B it applies a normal force on it. By
Newton’s third law, object B must be applying an equal but opposite normal force on A. But, what
is the direction and magnitude of a normal force?

The normal force on A is always perpendicularly away from B (as B is opposing the pressure from
A). The magnitude is as large as needed so as to keep A from penetrating into the surface of B.
So, for example, consider a book (object A) resting on a table (object B). The book is being pulled
downwards towards the ground by the Earth’s gravity. Since it is pressing against the table it also
feels a normal force due to the table. The normal force is pointing upwards (perpendicularly away
from the surface of the table) and is equal in magnitude to the weight of the book. The total force
on the book in the vertical direction is then,
(book)
Ftot = −mbook g + N = 0 . (3.9)

Therefore, the book is not accelerating in the vertical and keeps its position at rest with respect to
the table.

Tension

Let’s consider a lamp of mass m hanging from the ceiling by a massless string or rope. Gravity is
pulling the lamp downwards with a force equal to the lamp’s weight,

W = −mg . (3.10)

Since the lamp is stationary the rope must be pulling the lamp upwards with a force opposing the
weight of the lamp Frope−on−lamp = −W so that the sum of forces acting on the lamp vanishes. By
Newton’s third law, the lamp is applying an equal but opposite force on the rope and so the rope
is being pulled downwards by Flamp−on−rope = W . This force is the tension in the rope and since
the rope is stiff and massless, this tension must be the same throughout,

T = Flamp−on−rope = W . (3.11)

The ceiling is also pulling the rope upwards and so the pull of the ceiling must exactly equal
the tension in the rope since the total force must vanish. By Newton’s third law, the rope must
therefore be applying an equal but opposite force on the ceiling, pulling the ceiling downwards with
a force equal to the tension T . Since T = W , we conclude that the ceiling is experiencing a pull
downwards equal to the weight of the lamp. This is intuitively the result we expect because the
rope is massless and only acts as a connector between the lamp and the ceiling.

More generally, ropes and strings provide a connection between two separate objects and the tension
on them is uniform throughout and applies to both object on the two ends. These considerations
apply to many situations even when ropes and strings are not obviously present. For example, the
stem of an apple connecting the apple to the branch is acting like a string with tension equal to
the weight of the apple.

20
Problems

1. Gravity provides a central force that allows objects to execute circular motion around the
Earth, that is to say, it allows satellites in orbit. Let’s consider an object moving fast enough
to be in a circular orbit close to the Earth’s surface.

(a) Draw the force diagram on the object as it moves in its orbit.
(b) How fast must the object be moving in order to stay in orbit?

Such low orbits are not practical because air resistance will cause the satellite to heat up and
burn. Satellites are therefore flown at much higher altitudes where there is no atmosphere.
There are special types of satellites call gyrostationary satellites that cruise at an altitude of
about 42 × 103 km. They are moving at a speed of 3054m/s2 .

(a) What is the centripetal acceleration due to gravity at such high altitudes?
(b) How many days does it take the satellite to complete one revolution around the Earth.
(c) What do you think is the significance of such a period?

2. It is possible to measure the spring constant of a given spring by using it to hang a small
mass against gravity. Denote the spring length before it is hanged by l0 and after by l.

(a) Draw the force diagram on the mass assuming it is in equilibrium.


(b) Write down Newton’s second law associated with the mass.
(c) Derive an expression for the spring constant in terms of the spring’s length before and
after and g.

3. When considering the effect of more than one spring acting, it is useful to think about the
following two simple situations,

(a) Two springs, with spring constant k1 and k2 are attached in parallel to a mass. As
long as the motion is rigid and remains 1-dimensional, it is possible to think about this
system as a single spring of spring constant k. Express k in terms of k1 and k2 .
(b) Two springs, with spring constant k1 and k2 are attached in series to a mass. As long as
the motion is rigid and remains 1-dimensional, it is possible to think about this system
as a single spring of spring constant k. Express k in terms of k1 and k2 .

4. A simple weighing device involves a spring attached to a flat board on which you can stand.
Before you place your weight on the board, the spring’s length is l0 . As you weigh on the
board the spring contracts to a shorter length l.

(a) Draw the forces acting on yourself.


(b) Draw the forces acting on the board on which you stand.
(c) Neglect the weight of the board itself. Assuming it is stationary, express your mass in
terms of the spring’s constant k, its original and final length, and g.

21
3.3 Friction and Drag

Friction and drag are both forces that result from the motion of an object against its environment.
Unlike the other forces we have encountered so far, friction and drag always oppose the motion of
an object so as to slow it down and bring it to rest with respect to its environment. These forces
are extremely important in practice because they are ever present in our daily life. Almost every
practical device must account for the presence of friction and drag.

Friction

As we saw above, when a solid object, A, presses against a surface (of another object), it feels a
normal force pointing perpendicularly away from the surface. If A moves along the surface then
generally friction will result, which is a force parallel to the surface acting to oppose the motion of
A. It is known from experience that the magnitude of this force is,
F = µk N , (3.12)
where N is the normal force acting on A due to the surface, and µk is a dimensionless constant
coefficient characterizing the materials involved. The subscript k stands for kinetic friction as this
force is known in this case where A is moving with respect to the surface.

If A presses against a surface, but is at rest relative to it then in general static friction is present.
This is a force acting on A due to the surface that opposes any other force acting on A that tries
to move A along the surface. It is again known from experience that the maximum magnitude of
the static friction is,
F = µs N , (3.13)
where µs is the coefficient of static friction, which in general is different from the coefficient of kinetic
friction, µk . I want to emphasize to you that static friction is present only when another force (such
as your own push) is pressing on A as to move it from rest along the surface. The static friction
force quoted in Eq. (3.13) is the maximum force available to oppose any other force attempting to
move the object. In contrast, kinetic friction is present in the form quoted in Eq. (3.12) as long as
the object moves along the surface independently of other forces acting on the object (so long as it
maintains contact with the surface so that the normal force is not zero).

Drag

Like friction, drag always opposes the motion of an object on which it is acting. Unlike friction,
drag increases with the velocity of the object in the medium. Drag is present when objects move in
liquids and gases. There are two main types of drag, one called linear drag and the other quadratic
drag, denoting the linear and quadratic dependence on velocity, respectively. Linear drag, which is
appropriate at low velocity, is represented by the relation
F~ = −b~v , (3.14)

22
where ~v is the velocity of the object in the medium, and b is some constant coefficient with the
dimensions of Ns/m characterizing the strength of the drag force.

Quadratic drag is appropriate when an object moves at high velocities in the medium and is given
by,

F = 21 ρACd v 2 . (3.15)

Here, ρ is the mass density of the medium, A is the reference area, v is the magnitude of the
velocity, and Cd is a dimensionless characterizing the quadratic drag in the medium. The direction
of the force is opposite to the motion.

23
Problems

1. A banked curve is one that is positioned at some angle with the horizon to allow for a safer
turn. Consider then a circular banked curve with a radius R making an angle θ with the
horizon. The surface of the road has a static coefficient of friction µs .

(a) Draw the force diagram associated with a car taking a turn. Identify the force of gravity,
the normal force, and the friction.
(b) Find the maximum velocity a car can have on the road as a function of the curve’s
radius R, the angle of the bank θ, the coefficient of friction µs , and the gravitational
acceleration g.
(c) Find the maximum velocity for a car of mass a radius R = 100 m, static coefficient
µs = 0.4, and an angle θ = 20◦ . Compare this maximum velocity to the maximum
velocity on a flat circular turn of the same radius.

* I encourage you to imagine yourself driving in these different speeds to gain an appreciation
for the difference.

2. Consider an object of mass m falling from a considerable height. Air presents quadratic drag
on the object.

(a) Draw the force diagram for the object during its fall.
(b) Before the object hits the ground, it may stop accelerating if the drag becomes sufficiently
large. How large must the drag be for this to happen?
(c) How large must the velocity be for the object to stop accelerating towards the ground?
Express your answer in terms of m, g, ρair , Cd, air , and A. This maximal velocity is
known as the terminal velocity.
(d) Estimate how much larger is the terminal velocity of a human being compared to that
of an ant.

24
4 Energy

The notion of energy permeates much of physics and allows us to understand both qualitatively
and quantitatively a tantalizingly diverse range of phenomena. In this chapter we will begin to
explore the notion of energy through the consideration of mechanical systems. These humble origins
are unlikely to convince you of its amazing utility. We will encounter this concept several more
times during this course and in different physical systems, so you might begin to develop some
appreciation for its usefulness by the end of the course. At any rate, let us move on and begin by
introducing energy in simple mechanical systems.

4.1 Conservative Forces

The dynamics of many systems is governed by forces that result in a constant of the motion known
as the energy. Such forces are called conservative forces and they are defined through a function
referred to as the potential energy.

Energy in one dimension

In one-dimension, the situation is particularly simple. Consider a force acting on a single body,
where the force takes the following form,
dU (x)
F =− , (4.1)
dx
where U (x) is the potential energy. From Newton’s second law we then have,
dv dU
F =m =− , (4.2)
dt dx
and multiplying both sides by v = dx/dt we obtain,
dv dx dU
mv =− . (4.3)
dt dt dx
If U (x) is not an explicit function of time, and only depends on time through the coordinate x(t),
then the chain rule of calculus allows us to simplify the RHS,
dv dU
mv =− . (4.4)
dt dt
Moving the potential energy to the LHS and noting that vdv/dt = 12 dv 2 /dt we find,
d 1 2

dt 2 mv + U (x) = 0 . (4.5)

Therefore, the energy


1
E = mv 2 + U (x) , (4.6)
2

25
is a constant of the motion. The part of the energy which is not the potential energy is known as
the kinetic energy and is customarily labeled K,
1
K = mv 2 . (4.7)
2
It is the part of the energy that is associated with the motion itself. Together, the kinetic and
potential energies form the total energy, Eq. (4.6). I would like to emphasize again that its only
the total energy that is conserved.

The standard unit by which energy is measured is known as the joule and is defined as,

kg × m2
joule ≡ . (4.8)
s2

Energy in three dimensions

The generalization to 2 or 3-dimensional motion is straightforward. A potential energy function


U (~r) defines the force,
 
dU (~r) dU (~r) dU (~r)
F~ = , , , (4.9)
dx dy dz

and following similar steps to the one above we find that the energy,

E = 21 m vx2 + vy2 + vz2 + U (~r) ,



(4.10)

is conserved. The part that involves the velocity is again the kinetic energy of the system,

K = 12 m vx2 + vy2 + vz2 .



(4.11)

The zero of the potential energy

The equations of motion, which determine the dynamics of the system, are insensitive to any
constant shift of the potential energy since the force is the derivative of the potential. Therefore,
the zero of the potential can be chosen at our convenience. However, once a choice is made, it must
be respected throughout the analysis of the motion.

26
Problems

1. The potential associated with a one-dimensional spring is given by U (x) = 12 kx2 , where x is
the displacement from equilibrium.

(a) Calculate the force.


(b) As we saw in a previous problem, the solution for the motion can be written as x(t) =
A cos(ωt). Show that the energy E is indeed a constant of the motion.
(c) Plot the two components of the energy, the kinetic and potential, as a function of time.
Comment on the relation between the two.

2. The potential associated with the gravitational force near the surface of the earth is given
by U (h) = mgh, where h is the height from the surface (you should think about it as a one
dimensional problem along the vertical with h being the coordinate).

(a) Calculate the force.


(b) A small ball is rolling down an incline with very little friction starting from rest at a
height of 2 m above ground. What is the ball’s speed when it reaches ground level?
(c) How fast is the ball moving when it is still 1 m above the ground?

27
4.2 Energy Conservation and Work

In the previous section we discussed the dynamics of a single body and saw that when only con-
servative forces are acting on it then its energy is conserved. Its energy is made of kinetic energy
and potential energy and the sum of these two types of energies is known as the mechanical energy
of the body. Gravity, electricity, and magnetism are all examples of important forces that are also
conservative. So the conservation of energy is useful for understanding the motion of a particle
under their influence. But there are other forces, such as friction and drag, which are not conser-
vative, they cannot be written in terms of a potential energy. What do we do in those cases? Is
energy still a useful quantity?

Over the last couple of centuries the following empirical observation has been made: when a body
is slowed down by friction or drag (or other dissipative forces), heat is transferred to the stopping
medium. If the body itself is made of many smaller parts then heat may also be transferred to
itself in the process. If this heat quantity is included in the balance of energy, then the energy
of the system as a whole, body + medium, is conserved. Thus, conservation of energy is still a
useful notion, but we must include heat in the balance of energy. Importantly, while the mechanical
energy is associated with the body under consideration, heat is the energy quantity transferred to
the medium the body is interacting with or to the body’s own internal components. 2

But, whether the forces are conservative or dissipative, the kinetic energy of a body is something
that we can measure directly. Conservation of energy tells us that the change in the kinetic energy
is either associated with a change in potential energy or the transfer of heat, either one is important.
So, let us define a quantity which we call work that is the change in a body’s kinetic energy,

W = 12 mvf2 − 12 mvi2 , (4.12)

where m is the mass of the body, and vi and vf are its velocities at some initial and final times,
respectively3 . When the kinetic energy of a body increases, the work is positive and we say that
work was done on the body. Alternatively, when the body’s kinetic energy decreases, the work is
negative and we say that work was done by the body.

The notion of work can be directly related to the forces acting on a body. Work can also be defined
as the time integral over the force acting on a body in the direction of the body’s velocity,
Z tf
W = F~ · ~v dt , (4.13)
ti

where the integral is over the total force acting on the body at every point along its motion from
the initial moment ti to the final moment tf . The dot product, F~ · ~v is defined as,

F~ · ~v ≡ |F ||v| cos θF v , (4.14)


2
The study of heat, how it is exchanged between different systems, and how it affects other properties of the
medium is known as thermodynamics. It is an extremely important subject with far reaching implications for both
fundamental and applied science and technology, but alas it is beyond the scope of this course.
3
Keep in mind that in more than one dimension, the velocity squared involves all the different components of
velocity. So for example, in three dimensions vf2 = vf2 x + vf2 y + vf2 z involving the x,y, and z components of the
velocity.

28
where |F | and |v| are the magnitudes of the force and velocity, respectively, and θF v is the angle
between these two vectors. To see that these two definitions of work, Eqs. (4.12) and (4.13), are
equivalent, we need Newton’s second law, F = ma,
Z tf Z tf Z tf
~ d~v
W = F · ~v dt = (m~a) · ~v dt = m · ~v dt (4.15)
ti ti ti dt
Z tf
1 d 2
= m 2 dt |v| dt
ti
2
= 1
2 mvf − 12 mvi2 .

In the second line I have used the chain rule while in the last line I employed the fundamental
theorem of calculus. This result is very important as it tells us how to relate the forces acting on
a body to the change in its kinetic energy, in other words the work.

There is yet a third way to understand work, which is closely related to the integral definition,
Eq. (4.13), but without reference to the velocity of the body,
Z ~
xf
W = F~ · d~x , (4.16)
~
xi

where ~xi and ~xf are the initial and final positions of the body, respectively. You can prove that
this is an equivalent definition to Eq. (4.13) by using the relation ~v = d~x/dt and the chain rule of
calculus.

29
Problems

1. A body is rotating about a center by a tight string of radius R and tension T .

(a) Draw the force diagram on the body


(b) Find the velocity of the particle in terms of the radius and tension of the string.
(c) Find the period of the motion in terms of the radius and tension of the string.
(d) How much work is done on the body during one period?

* For the last part, think carefully about the direction of the velocity and the direction of the
force. Consult Eqs. (4.13) and (4.14).

30
4.3 Work and Potential Energy

When the forces acting on an object are conservative, there exists a very simple relation between
the work done and the potential energy. For simplicity I will illustrate it to you in one dimension,
but the proof in any number of dimensions is very similar. I remind you that in one dimension a
conservative force can be written in terms of some potential function U (x), namely F = −dU (x)/dx.
Using the definition of work given by Eq. (4.16) for the case of one dimension,
Z xf Z xf  
dU (x)
W = F dx = − dx = − (U (xf ) − U (xi )) , (4.17)
xi xi dx
where I have used the fundamental theorem of calculus in that last step. This says that the work
is equal to the change between the initial potential energy and the final potential energy. It does
not matter what happened in between!

To better appreciate how powerful this relation is, let us consider for example hydropower. The
basic principle of hydropower is to utilize the speed water gains as they drop from a height in
order to drive turbines. What is the maximal amount of work that can be extracted from a certain
amount of water dropping a given height? The force acting on the water as they drop is that of
gravity and so the potential energy is given by,
U (h) = mgh . (4.18)
If the hydro plant is constructed such that water drops from a height h2 to a lower height h1 , then
the work done by the water on the turbines is at most the work they gained by falling under gravity
(neglecting their initial speed),
W = − (mg h1 − mg h2 ) = mg(h2 − h1 ) . (4.19)
Now, the actual amount of work extracted is always lower because machines are not perfectly
efficient. But, think how impressive this relation is. With all its simplicity, it is giving us a very
sharp upper limit on how much energy we can hope to extract from a given power source. With
the ever-increasing demand for energy in the world, it is important to be able to clearly identify
and analyzed the different energy resources and their potential output.

Suppose now that there is both a conservative force Fcons as well as another unspecified external
force, Fext . The external force may or may not be conservative and so we will not assume anything
about its form. Using the above we can arrive at a handy relation between the amount of work
done by the external force and the change in potential and kinetic energies. From the definition of
work, Eq. (4.13), we have that,
W = ∆K , (4.20)
where ∆K is the difference in kinetic energy. From the discussion above we have,
Z xf
W = (Fext + Fcons ) dx = Wext − ∆U , (4.21)
xi
where ∆U is the difference in potential energy. Combining Eqs. (4.20) and (4.20) we have,
Wext = ∆K + ∆U . (4.22)

31
4.4 Power

As we learned in the previous section, the amount of work done by a body or on a body is related
to the change in the body’s motion (more precisely to its kinetic energy). For example, if a car’s
engine does a certain amount of work on the car it can accelerate it to a certain high speed. So
the notion of work is important in understanding how much we could potentially change a body’s
motion. However, it makes a big difference whether we could accelerate the car to 100 km/hour
in a year vs. if we could accelerate it to 100 km/hour in 6 seconds. You will not think much of a
car that takes a year to accelerate to 100 km/hour, but you might be interested in one that does it
in 6 seconds. And yet, the amount of work involved in both cases is the same. So what is it that
distinguishes these two obviously different situations?

Clearly, it is the amount of work for a given amount of time which is important. This is known as
power. It is the rate of change of work with time,
dW
P = . (4.23)
dt
For illustration, let us consider a situation where the power delivered over some duration, say T , is
a constant, say Pc . We can then integrate both sides of Eq. (4.23) to get,
Pc T = W → Pc = W/T . (4.24)
We see now why a car that takes 6 seconds to accelerate to 100 km/hour is more powerful than a
car that takes a year to do that. The amount of work, W , is the same, but the duration T is much
shorter. So it must deliver more power to affect this change. The standard dimension by which
power is measured is watt, which is defined as,
1 watt ≡ joule/s . (4.25)
Another very common unit of power that you are likely to encounter when dealing with locomotives,
engines, turbines, and etc. is the horsepower. One unit of horsepower is roughly the amount of
power that can be delivered by an average horse. It became useful to define this unit when people
began to make comparisons between horses and steam engines. In terms of the standard watt, a
horse power is given by,
1 horsepower = 745 watt . (4.26)
Cars’ power is typically measured in units of horsepower. A standard measure of the power of a car
is how long it takes it to accelerate from rest to 100 km/hour. Using Eq. (4.24) and the definition
of work as the change in kinetic energy we see that,
1m 2
T = v . (4.27)
2 Pc
This makes it clear that the larger the power to mass ratio (also known as ”power-to-weight”), the
less time it will take the car to reach 100 km/hour. A high-performance car, like a Ferrari, may
have in the neighborhood of 350-400 horsepower and a mass of about 1,400 kg. Using Eq. (4.27),
you can estimate that it should take it about 2 seconds to accelerate to 100 km/hour. In realty
it is more like 4-5 seconds because the engine can not transmit its power to the wheels with 100%
efficiency.

32
5 Momentum

If we know the position as a function of time, ~r(t), for all the bodies in the system then we have
solved the dynamics of the system completely and we know all of its past, present, and future
motion. If all the forces in the problem are specified, it is possible to find this solution through a
careful analysis of the equations of motion. However, in many cases we can easily identify certain
quantities which remain constant through the motion, i.e. they are independent of time. These
are known as constants of the motion and they are particularly useful since knowing their value at
any moment, we know their values at all times, irrespective of the motion. In the last chapter we
encountered the notion of energy. I showed you that when only conservative forces are involved,
energy is a constant of the motion. In this chapter I will describe to you one more such constant
of the motion, known as the momentum.

5.1 Momentum Conservation

Newton’s second law applied to a single body states that,


d~
p
F~ = m~a = , (5.1)
dt
where in the last step I identified p~ = m~v . This is a vector quantity known as the momentum. In
general it is a function of time. However, if any of the components of the force vanishes Fx = 0,
Fy = 0, or Fz = 0 then the corresponding component of the momentum is conserved,
dpi
0= → pi = constant . (5.2)
dt
If there are no external forces on the system altogether, then all the components of the momentum
are conserved, which is simply Newton’s first law of motion. Thus momentum changes only when
a force is applied. The change in momentum is known as the impulse,
Z tf Z tf
~ ~ d~
p
J= F dt = dt = p~f − p~i . (5.3)
ti ti dt

In this last step I made use of the fundamental theorem of calculus once again. The impulse, or the
change in momentum, depends on the time integral of the force over some duration. Therefore the
same impulse may result from applying a lot of force over a short period of time as when applying
a little bit of force over a long period of time.

The situation becomes more interesting when we consider more than a single body. Consider then
two bodies, with masses m1 and m2 , that exert some force on each other, but otherwise have no
forces applied to them. Writing Newton’s second law we have,
d~
p1
F~1 = , (5.4)
dt
d~
p2
F~2 = , (5.5)
dt

33
where p~1 = m1~v1 is the momentum associated with the first particle, and p~2 = m2~v2 is the
momentum associated with the second particle. Since we assumed that no external forces are
present, then from Newton’s third law we have that F~1 = −F~2 . Adding the two equations we then
find,
d
0= (~
p1 + p~2 ) → P~tot = p~1 + p~2 is conserved! (5.6)
dt
The total momentum of the system, which is the sum of the individual momenta, is conserved.

This powerful conclusion can be extended to systems with arbitrary number of interacting bodies
and no external forces. Consider N bodies with different masses m1 , m2 , . . . , mN with individual
momenta p~1 = m1~v1 , p~2 = m2~v2 , . . . , p~N = mN~vN . Each of these bodies feels a force due to each of
the others. I will denote by F~ij the force on particle i due to particle j. Then writing Newton’s
second law for each of the bodies we have
d~
p1
F~12 + F~13 + . . . + F~1N = (5.7)
dt
d~
p2
F~21 + F~23 + . . . + F~2N =
dt
.. ..
. .

d~
pN
F~N1 + F~N2 + . . . + F~(N−1)N =
dt
Adding all these equations together the left hand side cancels exactly because of Newton’s third
law, F~ij = −F~ji . We are then left with the equation

d
0= (~
p1 + p~2 + . . . + p~N ) → P~tot = p~1 + p~2 + . . . + p~N is conserved! (5.8)
dt
Thus when no external forces are present the total momentum of any system is exactly conserved.

These simple considerations are also useful when some of the components of the motion are subject
to external forces.

34
Problems

1. There is a particularly simple type of collision that can be analyzed using the conservation
of momentum alone. It is known as a totally inelastic collision. Consider an object with
mass m1 moving with velocity v1i towards another object with mass m2 moving with velocity
v2i . The two objects undergo a collision during which they stick, moving out together with
velocity vf . Mass is conserved in the process.

(a) v1i and v2i may have either sign and thus may be pointing to the right or to the left. Draw
a diagram indicating all the possible relative orientations of the velocities that would
result in a collision. Consider all three cases |v1i | > |v2i |, |v1i | < |v2i |, and |v1i | = |v2i |.
(b) What is the momentum of each of the objects before the collision?
(c) What is the total momentum before the collision?
(d) What is the total momentum after the collision?
(e) What is the velocity of the two objects after the collision in terms of the initial velocity?

* While this particular process is rarely of much interest, the time reversed process is of great
interest in propulsion.

35
5.2 Elastic collisions

Bodies, especially microscopic ones, can collide and recoil off each other without losing energy to
heat or other internal degrees of freedom. Such collisions are known as elastic collisions. Many
collisions are approximately elastic, that is to say the energy lost is a small fraction of the ki-
netic energy and does not affect the dynamics. When no external forces are present we can use
conservation of energy and momentum to solve for the motion of the bodies involved.

We begin with one-dimension. Since no forces are present before the collision, Galileo’s principle
informs us that the two particles move with constant velocities, which I will denote with v1i and v2i .
Similarly, since no forces are present after the collision the two particles again move with constant
velocities, v1f and v2f . Finding the relation between the initial and final velocities is the solution
to the collision problem.

Since no external forces are present at any time, including the collision itself, the total momentum
of the system is conserved and so,

m1 v1i + m2 v2i = m1 v1f + m2 v2f . (5.9)

By assumption, this is an elastic collision and so the total kinetic energy is conserved,
1 2 2 2 2
2 m1 v1i + 12 m2 v2i = 12 m1 v1f + 21 m2 v2f . (5.10)

There are two solutions to these equations. The first is not very interesting, v1i = v1f and v2i = v2f .
It is physical and just means that no collision occurred. The second solution is more interesting
and is given by,

m1 − m2
   
2m2
v1f = v1i + v2i , (5.11)
m1 + m2 m1 + m2
m2 − m1
   
2m1
v2f = v1i + v2i . (5.12)
m1 + m2 m1 + m2

There are several simple cases where this general solution is particularly simple. First, if the two
objects have the same mass, m1 = m2 , then after the collision they simply exchange their velocities,
v1f = v2i and v2f = v1i .

Second, if the first object is much heavier than the second one, m1  m2 , then

v1f = v1i , (5.13)


v2f = −v2i + 2v1i . (5.14)

The heavier object does not change its motion, while the lighter one is greatly affected by the
collision. In particular, if v1i = 0 then the lighter object just bounces off the heavier one, v2f = −v2i .

36
5.3 Explosions

There are many natural as well as human-made phenomena where a body at rest disintegrates into
its components due to some internal release of energy. In the case of explosions, for example, the
internal energy (be it chemical, nuclear, or otherwise) results in heat transfer to the components’s
kinetic energy, which causes them to fly apart, namely to explode. How fast do the pieces move
outwards? How much energy was released in the explosion?

Let’s begin with the simplest system involving a single body disintegrating into two components of
mass m1 and m2 . The disintegration process involves the release of heat Q, which is transferred to
the kinetic energy of the outgoing pieces, so conservation of energy dictates,

Q = 21 m1 v12 + 12 m2 v22 , (5.15)

where v1 and v2 are the velocities of the two pieces. Since there are no external forces acting on
the system, momentum is also conserved during the process,

0 = m1 v1 + m2 v2 , (5.16)

where the left hand side is zero because the initial body is at rest so its momentum vanishes.
Momentum conservation, Eq. (5.16), and the positivity of the inertial mass make it immediately
evident that the outgoing pieces must fly apart from each other in opposite directions,

m1 v1 = −m2 v2 . (5.17)

Using the conservation of energy, Eq. (5.15), we can solve for the outgoing velocities,
r r
m2 2Q
v1 = ± , (5.18)
m1 m1 + m2
r r
m1 2Q
v2 = ∓ . (5.19)
m2 m1 + m2

The signs represent the physical point we made above, namely that if one of the pieces is flying out
in the positive direction then the other must be flying out in the negative direction and vice-versa.
The problem as posed does not provide enough information to further determine which piece goes
in what direction.

Let’s consider a couple of simple cases. If the masses are equal, m1 = m2 = m, then
p the velocities
are equal in magnitude and opposite in direction, v1 = −v2 , and |v1 | = |v2 | = Q/m. On the
other hand, if the first piece is much heavier than the second one, m1  m2 , we have,
r
m2 2Q
v1 → ± , (5.20)
m m2
r1
2Q
v2 → ∓ . (5.21)
m2
Therefore, the heavier object moves out much slower than the lighter one by a factor of m2 /m1 .

37
Problems

1. A firecracker of mass 100 gram is launched vertically upwards at a speed of 20m/s. After 2
seconds, the firecracker explodes in midair and breaks into two pieces, a heavier piece of mass
80 gram, and a lighter piece of mass 20 gram. The explosion released 4 J of energy that went
into propelling the pieces in the horizontal direction only.

(a) What is the vertical velocity of the pieces directly before and directly after the explosion?
(b) Draw the forces on the firecracker before the explosion.
(c) Draw the forces on the pieces after the explosion.
(d) Which component of what velocity is conserved throughout the motion?
(e) What is the velocity of the lighter piece directly after the explosion?
(f) How far from the launch pad does the lighter piece land?

38
6 Simple Harmonic Motion

6.1 Dynamics

As we saw above, conservative forces can be described through a potential function. In one-
dimension the relation between the force and kinetic energy is given by,

dU (x)
F =− . (6.1)
dx
The simplest function is U (x) = constant, which results in no force.

The next simplest function is a linear function, U (x) = Fc x, with Fc some constant with dimensions
of Newton. This potential results in a constant force F = −Fc , and a constant acceleration
a = −Fc /m. We have encountered such a potential in the case of gravity near the surface of the
earth where Fc = mg, and x was a measure of the height from the surface, h. In that case the
potential energy is simply U (h) = mgh.

At the next order of complexity is a quadratic function, U (x) = 12 kx2 , where k is some constant
with units of Newton/m. This is a very important potential that appears in a wide variety of
phenomena. The force associated with such a potential is,

dU (x)
F =− = −kx(t) . (6.2)
dx
You might recognize this force as that of a simple spring with spring constant k, namely Hooke’s
law. This is indeed the case, but it turns out that this force also serves as a good model for many
other physical systems. The equation of motion associated with such a force is,

d2 x(t) k
2
= − x(t) . (6.3)
dt m
As usual, this is a differential equation whose solution is the position as a function of time, x(t).
The general solution to the equation of motion, Eq. (6.3) is given by,

x(t) = A cos(ωt + φ) , (6.4)

where
p
ω= k/m (6.5)

is called the angular frequency. It is an intrinsic quantity of the system. A is the amplitude of
the oscillations and φ is the phase, which defines the initial position. Both A and φ are extrinsic
quantities, which vary depending on the initial conditions.

Such a motion is known as simple harmonic motion. It is an extremely important system, not just
because of its simplicity, but because it shows up in a countless number of situations in all fields of
science.

39
The motion is periodic with a period of T = 2π/ω. To see this we note that,

x(t + T ) = A cos(ω(t + T ) + φ)
= A (cos(ωT ) cos(ωt + φ) − sin(ωT ) sin(ωt + φ))
= A cos(ωt + φ)
= x(t) .

The frequency is the inverse of the period f = 1/T and hence we have a relation between the
frequency and the angular frequency,

ω = 2πf = . (6.6)
T

Since we have the full solution for the position as a function of time, we can obtain all the other
kinematical functions such as velocity and acceleration through differentiation,

dx(t)
v(t) = = −Aω sin(ωt + φ) , (6.7)
dt
dv(t)
a(t) = = −Aω 2 cos(ωt + φ) . (6.8)
dt

Since the force is derived from a potential, it is a conservative force and thus the total energy is a
conserved quantity. As usual, it is composed of kinetic energy and a potential energy,
1 1
K(t) = mv(t)2 = mA2 ω 2 sin2 (ωt + φ) , (6.9)
2 2
1 1
U (t) = kx(t)2 = kA2 cos2 (ωt + φ) . (6.10)
2 2
Each of these quantities varies with time, but their sum, which is the total energy, is a constant of
the motion,

Etot = K(t) + U (t) = 12 kA2 = 12 mω 2 A2 . (6.11)


p
This last equation is proved using the relation ω = k/m and the basic identity of trigonometry.

40
Problems

1. A body of mass m is hanging from the ceiling by a spring with spring constant k.

(a) Draw the force diagram on the object.


(b) If the object and spring are vertically at rest, what is the elongation of the spring as
compared with its resting position horizontally?

2. A horizontal spring with spring constant k = 10.0 N/m is attached to an object with mass
m = 1 kg. The mass is displaced by x0 = 10 cm from the spring’s equilibrium point and
released.

(a) Derive an expression for the speed of the mass as a function of its displacement x from
equilibrium. Express the answer in terms of the displacement x, the initial displacement
x0 , the spring’s constant k, and the object’s mass m.
(b) What is the maximum speed of the object?
(c) How far from equilibrium is the object displaced when its speed is half its maximum
speed?
(d) What is the acceleration of the body at each of these points ?

41
6.2 The Pendulum

So far in this chapter I discussed x(t) as the linear displacement from equilibrium. But simple
harmonic motion is not restricted to linear motion. Let’s consider a long string of length l and
negligible mass that is used to hang an object of mass m from the ceiling. This is known as a
pendulum and you are surely familiar with it from everyday life. Unperturbed, the object remains
in equilibrium along the vertical with the pull of gravity downwards balanced against the tension
of the string upwards.

mg

Figure 1: A pendulum consisting of a massive object hanging by a string. The forces acting on the
object, namely its weight and the string’s tension, are drawn as the object is moved away from the
vertical by an angle θ.

Now suppose the object is displaced so as to make an angle θ with the vertical, as shown in Fig. 1.
The forces are still gravity and the string’s tension, but as is clear from the picture, a horizontal
component is now present due to the string tension that acts to push the object back towards the
vertical. When the angle is small, the displacement in the vertical is negligible and we can write
Newton’s second law for the component of the force acting in the vertical,
0 = −mg + T cos θ(t) → T = mg/ cos θ(t) , (6.12)
where I have written the angle θ explicitly as a function of time, θ(t). From geometry we see that
the horizontal displacement is x(t) = l sin θ(t). Therefore, the horizontal component of Newton’s
second law reads,
d2 (l sin θ(t))
T sin θ(t) = −m , (6.13)
dt2
where the minus sign comes about because I chose positive angular displacements to correspond
to negative horizontal displacements from the vertical. When the angular displacement is small,

42
|θ(t)| ≈ 0 we can use the approximations sin θ(t) ≈ θ(t) and cos θ(t) ≈ 1. Then, using the results
for both the vertical and horizontal components, Eqs. (6.12) and (6.13), we obtain the important
relation,

d2 θ(t) g
2
= − θ(t) . (6.14)
dt l
This equation describes the motion of the pendulum when the angle remains small. It is same
equation as the equation for simple harmonic motion, Eq. (6.3) with the change of notation x(t) →
θ(t), k → g and m → g. Therefore, we can immediately write down the solution for the motion,
p
θ(t) = Θ cos( g/l t + φ) , (6.15)

where Θ is the maximum angular displacement and φ is the phase. The period associated with the
pendulum’s motion is,
s
−1 2π l
T =f = = 2π . (6.16)
ω g

You can convince yourself that on the Earth, a pendulum of length one meter will complete half
a period (exactly one swing) in almost exactly one second. This was the original definition of the
meter, namely it was the length of a pendulum that takes a second to complete one swing.

43
Problems

1. An object of mass mo = 50 kg is hanging from a massless string of length l = 1 m. A bullet of


mass mb = 100 gram is moving horizontally with velocity vb = 500 m/s towards the hanging
object. As it hits the hanging object, the bullet is embedded in the object resulting in a
totally inelastic collision.

(a) What is the velocity of the combined bullet + object system directly after the collision?
(b) How high in the air will the combined system move?
(c) What is the maximum angular displacement from equilibrium?Do you think the com-
bined system can be safely modelled as an oscillating pendulum? If so, what is the
period of oscillation?
(d) If you have a second bullet of the same mass as the first, how fast should it be moving
so that when it hits the pendulum at its equilibrium point it will bring it back to rest?

2. A pendulum of length l = 1 m is used to define the second in Paris where the local gravitational
acceleration is gparis = 9.81 m/s2 . A day is measured to be 86,400 swings of the pendulum.
A second identical pendulum on the equator measures that same day to be 86,269 swings.
What is the local gravitational acceleration on the equator? A third identical pendulum on
the north pole measures the day to be 86,488 swings. What is the gravitational acceleration
on the pole? Can you think of a few reasons for this variation?

44
6.3 Damped Harmonic Motion

Simple harmonic motion is characterized by sinusoidaly varying motion about some equilibrium
point. The oscillating spring about its position of rest, the oscillating pendulum about the vertical.
Such motion is surely familiar to you from everyday life, but you will likely contend that such
motion eventually decays and stops. This is true, and under certain assumptions it is not too
hard to model that aspect of the motion as well. As we saw when we discussed resistive, such as
friction and drag, these forces act to dampen the motion of object and bring them to a stop. Let’s
consider a pendulum once again then, but this time include the effect of the surrounding air and
its resistance as well as the mechanical friction on the. We will model the combined effect of the
resistive forces as linear drag,

Fr = −bv(t) = −bl , (6.17)
dt
where b is some constant of dimensions Ns/m, and I have used the fact that the velocity of the
pendulum is v = ldθ/dt. The equation of motion is then the same as before, Eq. (6.14), with the
combined resistive force added to the sum of forces,

d2 θ(t) g b dθ
2
= − θ(t) − , (6.18)
dt l m dt
where I have divided by l throughout. The solution to this equation is not very simple and depends
on the exact relation of the different parameters involved. But, in the limit that the resistence is
very weak it is given by,

θ(t) = Θ e−bt/2m cos (ωt + φ) , (6.19)


p
where ω = l/g − b2 /(4m2 ). In Fig. 2 I plotted the behavior of the angular displacement as a
function of time. Notice that while it oscillates in a periodic fashion as you expect from a pendulum,
the amplitude of the oscillations also diminishes over time, as you are familiar with from real life.

45
0.10

0.05

ΘHtL
0.00

-0.05

-0.10
0 50 100 150 200
tHsecL

Figure 2: The angular displacement of a pendulum when the effect of air resistance is accounted
for, Eq. (6.19). For the purpose of illustration, I have chosen the initial amplitude Θ = 0.1, the
length of the pendulum l = 1 m, and the ratio b/m = 0.03 N/kg.

6.4 Driven Oscillations

Because of drag and friction, simple harmonic motion will generally not persist unless aided by an
external force. If we denote such an external force by Fext , then the equations of motion for the
spring take the form,

d2 x(t)
Ftot = −kx(t) + Fext = m . (6.20)
dt2
The solution to this equation depends on the precise form of Fext and we will now consider a couple
of important examples.

Constant External Force - The Vertical Spring

Let’s first analyze the case of a constant external force. As a concrete example consider a mass m
hanging vertically from the ceiling by a spring with spring constant k. For a change we will take
the downwards direction to be positive. The equation of motion, Eq. (6.20), then takes the form,

d2 x(t)
− kx(t) + mg = m . (6.21)
dt2
Recall that x(t) denotes the displacement from the length of the spring at rest, so a positive value
actually implies an extension downwards. To solve for the motion, note that we can redefine the
displacement function as,

x0 (t) = x(t) − x0 , (6.22)

46
where x0 = mg/k. Then the equation takes the form,

d2 x0 (t)
− kx0 (t) = m (6.23)
dt2
(the RHS is easy because the derivative of a constant vanishes). But, we already encountered the
solution to this equation, it is simply Eq. (6.4) with a trivial change

x0 (t) = A cos(ωt + φ) , (6.24)

where ω = k/m. Therefore, using the relation between x0 (t) and x(t), Eq. (6.22), the solution
p

for the original displacement function is,

x(t) = x0 + A cos(ωt + φ) . (6.25)

This is the same behavior as usual as far as the oscillations are concerned, but the displacement
suffers a constant elongation, x0 = mg/k. This is as it should, the gravitational pull downwards is
causing the spring to adopt a new equilibrium length, which is longer by an amount x0 as compared
with the horizontal length of the spring at rest.

Periodic External Forces - Resonance Phenomena

It should be intuitively clear to anyone has who pushed a swing before that one can most efficiently
increase the amplitude of oscillations by pushing at the right moment. The periodic push must be
done in coordination with the periodic motion of the swing. This leads us to the idea of a periodic
external force,

Fext = F0 cos (ωext t) , (6.26)

where F0 is the amplitude of the external force, ωext is the angular frequency, and φext is the phase.
Using Eq. (6.20), we can again write the equation of motion for the oscillator,

d2 x(t)
− kx(t) + F0 cos (ωext t) = m . (6.27)
dt2
As you can confirm by direct substitution, there is one particularly simple solution to this equation,
namely,
 
F0 1
x(t) = 2 cos (ωext t) . (6.28)
m ω 2 − ωext
The presence of dissipative forces will generally dampen any other motion of the system and the
solution described by Eq. (6.28) represents the asymptotic motion of the system after a long time.
As can be expected intuitively, the system simply follows the periodicity of the driving force. But,
this solution also exhibits the importance of driving the system at the right frequency. Notice that
the amplitude of the oscillations in Eq. (6.28) grows as the frequency
p of the external driving force
comes close to the natural frequency of the oscillator, ωext → ω = k/m. Away from the natural
frequency the oscillator is driven inefficiently with a reduced amplitude.

47
This phenomenon of driving a system in its natural frequency is known as resonance. The rapid
growth of the amplitude as the driving frequency approaches the natural frequency is the hallmark
of resonance. You may be troubled by the divergence of the amplitude with ωext → ω. This
divergence is present because we neglected the important effects of dissipative forces present in
physical system. When those are accounted for, this divergence is softened. Nevertheless, it is still
true that the amplitude is largest when the frequency of the external driving force is on resonance
with the natural frequency of the system, ωext = ω.

The phenomenon of resonance is used in many applications. For example, radio receivers can be
thought of as an oscillator driven by the periodic force associated with radio waves in the air. Most
of the waves are out of resonance with the receiver and cannot be heard. When you tune your radio
to your favourite radio station you effectively tune the natural frequency of your radio receiver to
that station’s frequency. Hence, you amplify the receiver’s response to the ever-present radio waves
associated with that station.

48
Problems

1. Show that even in the presence of drag, the effect of a constant external force Fc is simply to
shift the equilibrium position by an amount Fc /k.

2. Consider an external force with an arbitrary phase, Fext = F0 cos (ωext t + φext ). Show that
the solution given in Eq. (6.28) is still a solution provided add the phase to the argument of
the cosine function.

49
7 Waves

We have so far considered a small number of objects, one or two bodies that may feel external forces
or may exert forces on each other. We found the motion of the system by analyzing the objects
individually, writing down the different forces that act on each object separately, using Newton’s
second law to arrive at the equations of motion, and finally solved these equations. Ultimately, we
were after the position as a function of time, x(t).

But what happens, if we have many many many bodies, all interacting with each other? We could
certainly try to write down a position function for each one, x1 (t), x2 (t), . . ., x1000 (t) and so on.
But, it is very difficult to solve all the uncountable number of equations associated with all the
different bodies. This is not just an academic question, many important and interesting systems
in nature are like this. Is there anything we can still hope to learn about such systems even if we
cannot solve them exactly? If the number of bodies is so large that they can be treated collectively
as some sort of medium, then indeed there is much we can learn about the medium. It turns out that
such systems often exhibit collective behavior that can be analyzed and understood without having
to understand all the internal interactions between the different bodies making up the medium.

For example, a tight string is composed of many atoms, but we may just want to describe its
displacement away from its stretched position. In that case, we can do so with a function y(x, t),
where x denotes the position along the string, t is the time, and y(x, t) denotes the displacement
at position x at time t. This is shown in Fig. (3). But, what is the explicit form of y(x, t)? When

y
t = t0

y(x0, t0)

x0
x

Figure 3: A general displacement along a stretched string is denoted by y(x, t). The figure shows
the displacement at a particular point, x0 , and at a particular time, t0 .

dealing with a single body, in order to solve for the explicit form of x(t) we drew a force diagram,
we used Newton’s second law to write down the equation of motion for x(t), and finally we solved
the equation of motion. It is possible to follow similar steps, applying Newton’s laws to every point
in the medium, to arrive at the equation of motion for y(x, t). The resulting equation for the case
of a string takes the form,
∂ 2 y(x, t) ∂ 2 y(x, t)
T = µ , (7.1)
∂x2 ∂t2
where T is the string’s tension, and µ is the mass density of the string (mass density = mass
of string/length of string). You can think of the right hand side as the usual md2 x(t)/dt2 term

50
in Newton’s second law, and the left hand side is the force, that is F = ma. But, this is now
applied at every point along the string, x. This equation is known as the wave equation and it finds
application in many fields of science and technology. As we shall explore in the next few sections,
the solutions of this equation are waves.

7.1 Characteristics of Waves

There are many types of media that support waves, and what is ”waving” in each one could be dif-
ferent. Waves are disturbances that propagate through the medium about some unperturbed state.
Water waves are disturbances in the height of the water’s surface. Sound waves are disturbances
in the local density of materials. Waves in a string are disturbances of the string’s displacement
about its stretched position. Electromagnetic waves are disturbances in the local strength of the
electromagnetic field. In each of these cases, the function y(x, t) represents something different
(height, displacement, density, and etc.), but the mathematical description is the same.

The most general solution to the wave equation, Eq. (7.1), is


y(x, t) = f (x − vw t) + g(x + vw t) , (7.2)
where f (x − vw t) is any continuous function of the single variable x − vw t, and g(x + vw t) is any
continuous function of the variable x + vw t. The constant vw has units of velocity and it is given
by,
s
T
vw = . (7.3)
µ

The first function, f (x − vw t) represents waves traveling in the positive x-direction. Notice that if
we simultaneously translate time by some amount ∆t and distance by an amount ∆x = vw ∆t then
f ((x + ∆x) − vw (t + ∆t)) = f (x − vw t). Similarly, the function g(x + vw t) represents waves traveling
in the negative x-direction. Therefore, a general wave in one dimension is propagating through the
medium in either one way or the other with velocity vw . This is the physical importance of the
constant vw . The velocity of propagation of waves through the medium depends on the properties
of the medium.

What is vibrating also depends on the medium. In the case of a stretched string for example, the
displacement of the string is vibrating about its stretched position. In water waves, the height
of the water is vibrating. Notice that in both of these cases, the vibrations are perpendicular to
the direction of the motion of the wave itself. Such waves are known as transverse waves, since
the vibrations at every point in the medium are transverse to the direction of the wave’s motion.
Importantly, we still call it a one-dimensional wave because the propagation of the wave is confined
to one dimension. Its vibration, however, is happening in a transverse dimension. For example,
a water wave in a narrow canal can be described as moving in the horizontal direction, but the
water’s displacement is in the vertical direction.

In contrast, the sound waves in a material are associated with compression and rarefaction of the
material’s density. The vibration in this case is along the direction of motion of the wave itself.

51
Such waves are known as longitudinal waves. The vibrations induced in a long spring are another
example of longitudinal waves. Earthquakes produce different type of waves, but the primary ones
(p-waves) and least destructive are longitudinal waves that travel fastest and therefore are first to
arrive. These correspond to compression and rarefactions in the earth’s density.

Sound

Sound waves are the compression and rarefaction of the local density of materials and as such
they are longitudinal waves. You are likely familiar with sound as it travels through air, a gas.
As you might know, for a fixed temperature, the density of a gas is related to its pressure. So
the compression and rarefaction of density associated with sound waves in air result in changes of
pressure, which is what your hear drum senses as sound. The speed of sound in dry air at 20◦ is

vs = 343 m/s . (7.4)

Note the particular way I quoted this velocity as depending on the temperature of the air as well
as its humidity. This is an example of the very general point mentioned above, namely that the
speed of waves in a medium depends on the properties of the medium. In the case of gases, the
general expression for the speed of sound is given by an expression similar to the speed of waves in
a string, Eq. (7.3), with the string’s tension replaced by the gas’ average pressure.

A gas is not the only medium to support sound waves. A Liquid or a solid can also serve as
media where density fluctuations result in the propagation of sound waves. Considering the very
different average densities and pressures found in liquids and solids, it is no surprise that sound
waves in liquids and solids propagate at very different speeds as compared to sound speed in air.
The dependence of the speed on the properties of the medium can actually be of use in learning
about the interior of materials. For example, p-waves, which are a type of sound-wave produced
by an earthquake, are used to study the Earth’s composition.

Light

One of the triumphs of 19th physics is the realization by James C. Maxwell that his theory of
electric and magnetic fields predicts the existence of light. He worked out the speed of waves in
the electromagnetic field and found to his amazement that it is the speed of light in vacuum,

c = 299, 792, 458 m/s . (7.5)

This helped identifying light as an electromagnetic wave. Light is a transverse wave, but this fact is
by no means obvious. It is the oscillations of magnetic and electric fields in a plane perpendicular
to the direction of motion. But, unlike other waves where the medium is manifest (water for
water-waves, air for sound-waves, and etc.), light requires no medium to propagate. The absence
of any clear mechanical medium associated with light has perplexed 19th century physicists who
attempted to introduce such a medium, known as the ether. These attempts, and the ether, were
abandoned with the advent of Einstiein’s special theory of relativity.

52
When light travels through a medium other than vacuum, it slows down. The ratio of its velocity
in vacuum, Eq. (7.5) to its velocity in the medium is known as the index of refraction,
c
n= , (7.6)
v
where v is the light velocity in the medium. The index of refraction depends on the properties of
the medium as well as the wavelength of light.

53
7.2 Sinusoidal Waves

In this section we will consider a very particular form for the general function describing a wave,
f (x − vw t). It is known as a sinusoidal wave,

y(x, t) = A sin (kx − ωt + φ) , (7.7)

where ω is the angular frequency and k is known as the wave-number. This function is periodic in
both time and space. Concentrating on a particular point along the x-axis, say x = 0, we see that
the point in the string is executing a simple harmonic motion about the equilibrium position,

y(x0 , t) = A sin (ωt) . (7.8)

You can prove that y(0, t + 2π/ω) = y(0, t) and so the period of this motion is

T = . (7.9)
ω

The sinusoidal wave also exhibits spatial periodicity, repeating itself at a distance,

λ= , (7.10)
k
where λ is known as the wavelength. Indeed, you can show that at any given moment, say t = t0 ,
we have y(x + λ, t0 ) = y(x, t0 ). If it is a wave, then it must have some velocity, which can obtain
by writing Eq. (7.7) as,
 ω 
y(x, t) = A sin k x − t . (7.11)
k
Comparison with the defining equation of a general wave, Eq. (7.2), we see that,
ω λ
vw = = = fλ . (7.12)
k T
The velocity of a sinusoidal wave is its frequency times its wavelength.

Power and Intensity

The power of a wave, which is the rate that energy is transported by the wave, has units of kg m2 /s2 .
For a wave on a string, there are four independent quantities that carry units, the mass density µ
with units of kg/m, the wave speed vw with units of m/s, the amplitude of the wave A with units
of m, and the angular frequency ω with units of s−1 . We expect the power to be delivered at the
same rate as the wave is propagating, hence P ∝ vw . Also, on dimensional grounds, the power
must involved one power of the mass density P ∝ µ because of the units associated with mass.
Therefore, again on dimensional grounds we can conclude that the power in a wave must be,

Pstring ∝ µA2 ω 2 vw , (7.13)

54
where the constant of proportionality turns out to be 1/2. Very generally, the power in waves
increases quadratically with the wave’s amplitude and linearly with its speed,

Pwave ∝ A2 vw . (7.14)

This result holds in general for waves whether they are sinusoidal or not.

As a wave propagates away from the source it spreads out and diminishes in amplitude. Instruments
that detect the wave are only sensitive to the area of the wavefront covered by the detector.
Therefore, it proves useful to define the intensity of the wave, which is its power divided by the
area over which the power is distributed. For example, the intensity of a uniform spherical wave at
a distance r from a source of power P is given by,
P
I= . (7.15)
4πr2
This relation between the intensity measured, the power emitted, and the distance to the source
is of great importance in many fields of science and technology. For example, the intensity of star
light or other astronomical objects is used by astronomers to infer the distance to other stars if
their power output is known or vice versa.

55
Problems

1. The density of Helium at room temperature is about ρHe = 0.16 kg/m3 . That of air is
ρAir = 1.2 kg/m3 .

(a) What is the speed of sound in Helium?


(b) The range of wavelength of sound you can produce is limited to some extent by the size of
your throat and mouth cavity. Using the relation between the velocity, the wavelength,
and the frequency, Eq. (7.12), what do you expect will happen to a person’s pitch after
inhaling helium?
(c) Compare predictions to observation, http://www.youtube.com/watch?v=jUgEMGYC8wI
.

56
7.3 Superposition of Waves

An important property of the wave equation, Eq. (7.1), is known as the superposition principle.
If y1 (x, t) and y2 (x, t) are two solutions of the wave equation, then so is their sum, ysum (x, t) =
y1 (x, t) + y2 (x, t). This simple observation leads to a variety of interesting phenomena.

Beats

The phenomenon of beats arises when two waves of similar, but not equal, frequencies overlap at
some region of space. Let’s consider then two sources of sound, a piano and a tuning fork, that
produce sound waves of different frequencies, say f1 and f2 , respectively. When the sound waves
reach our hear they combine together,

y(0, t) = ypiano (0, t) + yfork (0, t)

= sin (2πf1 t) + sin (2πf2 t) , (7.16)

where I have assumed our hear or listening device is positioned at x = 0. Also, I have chosen the
phases and amplitudes for convenience and simplicity, but the effect we are about to discuss does
not depend on these choices. Using a trigonometric identity we find that,

f1 − f2
       
f1 + f2
sin (2πf1 t) + sin (2πf2 t) = 2 cos 2π t sin 2π t . (7.17)
2 2

The second factor represents a note with frequency,


f1 + f2
fsum = . (7.18)
2
If f1 and f2 are similar then this note more or less coincides with the fork and piano’s notes. More
interestingly, the first factor in Eq. (7.17) can be thought of as the modulating amplitude of this
note. The frequency associated with this modulation is very small since it is the difference between
f1 and f2 which are close to each other,

fbeat = f1 − f2 . (7.19)

Typical notes are measured in hundreds of Hz and the human hear cannot discern the rapid oscil-
lations associated with the notes. However, fbeat can easily be only a few Hz or less as one tunes
the piano to the fork and bring f1 → f2 . Such amplitude modulations are certainly detectable by
our hears and one can perceive the note’s intensity increasing and decreasing in a periodic fashion.
The piano is tuned when fbeat = f1 − f2 → 0. This can be detected by listening to the beat as its
period grows larger and larger.

57
7.4 Standing Waves

As we saw above, general sinusoidal waves are characterized by their wavelength, λ, and frequency,
f . The only restriction on these quantities is that their product must equal the wave velocity,
Eq. (7.12). Further restrictions arise when the medium is forbidden from vibrating at particular
points. For concreteness, let us consider a long string of length ` whose ends are clamped at two
points x = 0 and x = `, see Fig. (4). The displacement, y(x, t), must then satisfy the following
conditions,

y(0, t) = 0 , (7.20)
y(`, t) = 0 . (7.21)

The string cannot be displaced at its ends at any time. Clearly, this cannot be satisfied by a
traveling wave, because even if it vanishes at a point at a given moment, it will not vanish there
a moment later. But, what if we consider two waves traveling in opposite directions? The general
expression is,

y(x, t) = A sin k (x − vw t) + B sin k (x + vw t) . (7.22)

This sort of wave, which is the superposition of a wave moving to the right and a wave moving to
the left, can certainly satisfy the first condition, Eq. (7.20),

y(0, t) = A sin k (−vw t) + B sin k (+vw t) = 0 → A=B . (7.23)

With a simple trigonometric identity we can now write the wave as follows,

y(x, t) = 2A cos (ωt) sin (kx) . (7.24)

The string is also clamped at x = ` and so we must impose the second condition (Eq. (7.21), namely
y(`, t) = 0 at all times,

y(`, t) = 2A cos (ωt) sin (k`) = 0 → k` = nπ n = 1, 2, 3, . . . , (7.25)

where n is a positive integer. This condition results because for the function to vanish at all times,
the argument of the sinus function must vanish at x = `. This can be satisfied if it is an integer
multiple of π. We thus arrive at an important condition on the wavelengths supported by the string
clamped at both ends,
2`
λn = n = 1, 2, 3, . . . . (7.26)
n
Since the frequency is related to the wavelength through the speed of the wave, this condition also
translates to discretized values for the frequency (and angular frequency),
v nv  nv 
fn = w = w ⇒ ωn = 2πfn = 2π w
. (7.27)
λn 2l 2l
The longest wavelength supported is twice the string’s length, λ = 2`, corresponding to n = 1.
It is the known as the fundamental. Shorter wavelength are also supported, corresponding to

58
n = 2, 3, . . ., and are known as harmonics. The fundamental and the first three harmonics are
depicted in Fig. (4). So finally, the waveform satisfying all the boundary conditions is simply,

y(x, t) = 2A cos (ωn t) sin (2πx/λn ) . (7.28)

Remarkably, even though it was formed from the combination of two traveling waves, the resulting
wave is a standing wave! It has a spatial profile of sin (2πx/λn ) with an amplitude that oscillates
in time as 2A cos (ωn t).

λ = 2!

λ = !/2

x=0 λ = 2!/3 x=!


λ=!

Figure 4: A string clamped on both ends, x = 0 and x = `, supports standing waves of wavelength
λ = 2n/`, with n a positive integer. The figure depicts the fundamental vibration λ = 2` (n = 1)
as well as the first three harmonics λ = `, 2`/3, and `/2, corresponding to n = 2, 3, 4, respectively.

If you are familiar with any musical string instrument, then you have undoubtedly encountered the
phenomenon of standing waves. The different notes on the musical scale are produced by plucking
on a string of a well-defined length. On a violin, or a guitar, you can create a string of a particular
length by clamping down on the string with your finger at a particular place. One typically excites
the fundamental vibration, λ = 2`. The thickness and tension in the string determine the speed of
waves in the string via Eq. (7.3) and hence the vibration frequency of the fundamental is given by
Eq. (7.27) with n = 1,
vw
f= . (7.29)
2`
This is the frequency of the note produced. Alongside the fundamental, the higher harmonics are
also excited, f = vw /`, 3vw /2`, . . .. The contribution of the higher harmonics is essential in creating
a full-bodied sound. The same note, produced by two instruments, sounds differently because the
contribution of the higher harmonics is different between the two instruments. There are also ways
to clamp the string so as to produce one of the higher harmonics directly without exciting the
fundamental. The notes produced in this way are known as flageolets and are much purer since
there are fewer harmonics contributing.

Finally, I would like to briefly mention a phenomenon known as sympathetic


p resonance. In chapter 6
we discussed an harmonic oscillator with natural frequency ω = k/m driven by an external force
with frequency ωext . We saw that the best way to drive the oscillator is to do so on resonance,
that is by tuning the external force to the same frequency, ωext → ω. Now consider several close by

59
strings, like the strings of a guitar, each of them vibrating at a certain frequency. If the fundamental
(or harmonics) of one of the strings happens to coincide with the fundamental (or harmonics) of a
second string then it is possible to excite the second by simply plucking the first. The sound waves
produced by the first string act as an external force driving the second string on resonance in its
natural frequencies, the fundamental or harmonics. This has the very pleasant effect of creating a
very rich sound and is called sympathetic resonance.

60
7.5 Interference

Let us now consider what happens when waves of the same frequency, but a different source, arrive
at the same location. If the first wave travelled a distance d1 from its source, and the second wave
travelled a distance d2 , then we have,
y(x, t) = y1 (x, t) + y2 (x, t) = A sin (kd1 − ωt + φ01 ) + A sin (kd2 − ωt + φ02 )

= A sin (φ1 − ωt) + A sin (φ2 − ωt) , (7.30)


where φ01 and φ02 are the initial phases of the two waves, and
φ1 = kd1 + φ01 , (7.31)
φ2 = kd2 + φ02 . (7.32)
can be thought of as the phases of the waves at this particular location. Using the properties of the
sinus function, you can convince yourself that the oscillations of the two components in Eq. (7.30)
exactly cancel each other if they are out of phase. This is known as perfectly destructive interference
and requires,
∆φ = φ1 − φ2 = m + 12 · 2π ,

(7.33)
where m = 0, ±1, ±2, . . ., is an integer. On the other hand, the amplitudes of the two components
add coherently to yield perfectly constructive interference if the oscillations are exactly in phase
∆φ = φ1 − φ2 = 2 m π , (7.34)
where again, m = 0, ±1, ±2, . . .. More generally, the interference results in an amplitude that
depends on the phase shift. Using a trigonometric identity for expressing the sum of two sinus
functions as the product of two others, the waveform can be recast into,
y(x, t) = A sin (φ1 − ωt) + A sin (φ2 − ωt)
       
∆φ d1 + d2 φ01 + φ02
= 2A cos × sin k + − ωt . (7.35)
2 2 2
When the variations in the average distance, (d1 + d2 ) /2 are small compared with the wavelength,
the amplitude of the wave can be thought of as the coefficient of the oscillating piece, namely,
 
∆φ
Amplitude = 2A cos . (7.36)
2
In many instances the initial phases of the two waves are equal since they emerge from the same
source. In those cases, the conditions for interference can be stated as requirements on the path-
length difference. From the relations between the phases and distance travelled, Eqs. (7.31) and
(7.32) we see that,
∆d = d1 − d2 = mλ (constructive interference) , (7.37)
1

∆d = d1 − d2 = m+ 2 λ (destructive interference) . (7.38)
Here I used the relation between the wavenumber and the wavelength, k = 2π/λ.

61
Young’s Double Slit Experiment

One of the vexing questions of classical physics was whether light is a wave or a particle. The
beautiful experiment by Young seemed to have settle this question in favour of waves. Young
considered a single light source impinging on a screen with two slits separated by a distance D as
shown in Fig. 5. A second screen is placed a distance L behind the first. Using simple trigonometry,
the distance of the upper light ray to a point on the second screen is given by
L
d1 = (7.39)
cos θ
where the θ is the angle of the light ray with the horizontal as shown in the figure. Geometry then
dictates that the distance of the lower light ray to that same point on the screen is,
L
q
d2 = L2 + (D + L tan θ)2 ≈ + D tan θ , (7.40)
cos θ
Where the approximation holds when the distance to the screen is much larger than the slits’
separation L  D. The difference in distance between the two rays, which is the relevant quantity
for interference, is then

∆d = d2 − d1 ≈ D tan θ . (7.41)

The vertical distance along the screen is usually also much smaller than the distance L and so the
angle θ is small, θ  1 and so,
y
∆d ≈ D θ ≈ D . (7.42)
L
So far this is just geometry. But, now comes the crucial observation. If light is really a wave, then
we expect to see bright fringes when the path difference satisfies the condition for constructive
interference, Eq. (7.37),

λ λL
θm = m or ym = m . (7.43)
D D
Similarly, when the path difference satisfies the condition for destructive interference, Eq. (7.38)
we expect to see dark fringes,

1
 λ 1
 λL
θm = 2 +m or ym = 2 +m . (7.44)
D D
This pattern of bright and dark fringes is precisely what was observed by Young. This experiment
has conclusively evinced the wave-like nature of light. ”Light as a wave” stood as a paradigm
for about a century until Einstein demonstrated that the photoelectric effect can be understood
by assuming light to be a particle nevertheless. The clear paradox that emerged from the two
conflicting views of the nature of light foreshadowed the advent of quantum mechanics.

62
θ

Figure 5: A single light source on the left hand side of the figure illuminates a screen with two slits
separated by a distance D. The light rays emerging from the two slits arrive at a second screen
positioned at a distance L to the right of the first screen. Each point along the second screen is
illuminated by two rays, one from each slit, as shown in the figure for a particular location. If
light is a wave then the difference in distance travelled between the two rays would result in the
interference pattern shown on the right.

63
7.6 Reflection and Transmission Through a Boundary

The last topic we will consider regarding waves concerns what happens when a wave in a medium
arrives at the boundary and either gets reflected or crosses over to another medium. Let’s start
by considering the simplest cases of a long string which is attached to a fixed boundary or a free
boundary, Fig. (6). In the case of a free boundary, as the wave reaches the boundary the end-point
just slides up and then slides back down as the wave is reflected. Aside from the reflection, the
wave continues to move as if it never encountered the boundary - it experiences no phase-shift.

On the other hand, in the case of a fixed boundary, the end-point tries to move up but it encounters
the infinite resistance of the wall and so it bounces back down and the wave flips its orientation.
Aside from being reflected it gets flipped upside down - it experiences a π phase-shift 4 . In these
two cases of a perfectly free or a perfectly fixed boundaries the waves are perfectly reflected since
there is nowhere else for the energy to go. Now that we understand these limiting cases we can use
this intuition to understand what happens in the more general case.

(1a) (1b)

(2a) (2b)

(3a) (3b)

Figure 6: A wave propagating on a string approaches a free boundary on the left pane (1a) and a
fixed boundary on the right pane (1b). When the wave is reflected from the free-boundary (2a and
3a) it suffers no phase-shift. On the other hand, when reflecting off the fixed-boundary it suffers a
π phase-shift.
4
To under why it is a π phase-shift recall that waves are described by cosine and sine functions. We know that
for any argument θ we have that cos(θ + π) = − cos(θ) and sin(θ + π) = − sin(θ) and so the functions are flipped.
This is precisely what happens to the general wave when it is reflected by a fixed boundary.

64
The more general case is shown in Fig. (7) where two different strings (two different media) are
connected at a boundary. The tension in the strings has to be the same throughout by Newton’s
third law, but the mass density could be different. Let’s start by considering the case of a wave
initially in a thick string impinging on a boundary with a thin string. What will happen? If the
thin string was simply massless then it acts as a free-boundary as in Figs. (6a) and we know that
the reflected wave will not suffer any phase-shift. But it is not massless and so not perfectly free,
and we consequently expect the reflection not to be perfect. Where does that energy go? It goes
into the transmitted wave that now propagates in the new medium along the thinner string!

What happens in the opposite case of a wave propagating in a thin string which is connected by
a boundary to a thicker string? If the thicker string was much thicker and much more massive
we expect it to act as a fixed boundary as in Fig. (6b) and so the reflected wave will suffer a π
phase-shift. This is precisely what happens, the wave reflected back in the thin string is flipped
upside down. But, since the thicker string is not perfectly fixed some of the wave it transmitted
through and propagates in the new medium.

In the above, I was not sufficiently precise in my statements. Really what matters is not the
thickness of the string but its mass density. When moving from a medium with high mass density
to a new medium with low mass density there is no phase-shift in the reflected wave. On the
other hand, when moving from low to high the reflected wave undergoes a π phase-shift. I must
emphasize that the transmitted wave does not suffer a phase-shift in any of these cases.

(1a) (1b)

(2a) (2b)

(3a) (3b)

Figure 7: On the left is a wave moving in a string with a higher mass density which is connected
by a boundary to a second string with a lower mass density. Upon arrival at the boundary part
of the wave is reflected and part is transmitted. No phase-shift results. On the right is a wave
moving from a string with low mass density to another one with a high mass density. In this case
the reflected wave suffers a π phase-shift and flips over.
p
Recall that the speed of the wave in the string is given by Eq. (7.3), vw = T /µ. Therefore,

65
the transmitted wave changes its speed relative to the incoming wave. When moving from a
high-density to low-density string the speed increases and vice-versa. Therefore, another way to
understand when a π phase-shift occurs and when it doesn’t is to think about the change in velocity
between the two media:

• When crossing from low-speed to high-speed medium the reflected wave suffers no phase-shift.

• When crossing from high-speed to low-speed medium the reflected wave suffers a π phase-shift.

Before we move on to understand how all of this applies to light it is worth mentioning one last
general point about waves crossing a boundary. We saw that when moving from one type of string
to another the mass density changes, but the tension remains the same. That implied that the
velocity also changes across the boundary. What about the wavelength and the frequency, do they
change or remain the same? A moment of reflection will convince you that the frequency of the
wave remain the same as it crosses a boundary. Imagine knocking on thick wood every second. Any
animal leaving inside the wood will hear your knocks with the same frequency, one every second.

What about the wavelength then? We know that the velocity of the wave is vw = λf . We just saw
that the frequency, f , does not change across the boundary. Therefore the wavelength, λ, increases
when moving from low-speed medium to high-speed medium and decreases when moving from
high-speed to low-speed. This has important ramifications and we shall see how it manifests itself
when we consider interference phenomena due to the reflection and transmission of light through
different materials.

Thin-film interference

The beautiful colors displayed by oil-spills or soap bubbles are a familiar sight5 . But, what is
their origin? In particular, how come we see this multitude of colors when the material itself (oil
or soap) certainly does not possess all these colors and daylight is regular white light with no
special features? Moreover, we see that the color pattern repeats itself. What is the origin of this
regularity? This is a fairly involved phenomenon, but we are now in possession of all the physics
necessary to understand it in detail. Fundamentally it is an interference effect.

We note that this effect happens when light moving through one medium (e.g. air) encounters a
thin-film made of a different medium (e.g. oil or soap) as shown in Fig. (8a). As we know, part of
the light is reflected and part is transmitted through, and this is shown in Fig. (8b). The incident
light and reflected light both make an angle θ1 with the normal to the surface. The transmitted
light is refracted according to Snell’s law with angle θ2 obeying,

n1 sin θ1 = n2 sin θ2 . (7.45)

5
See the following webpages for images of this effect in oil-spills https://en.wikipedia.org/wiki/File:
Dieselrainbow.jpg, soap bubbles, https://en.wikipedia.org/wiki/File:Thinfilmbubble.jpg, and nail polish
https://en.wikipedia.org/wiki/File:Interferencia_lakkretegen2.jpg

66
incident reflected reflected
light light light

n1 θ1 n1 θ1 n1 θ1

transmitted θ2
n2 d n2 light n2 reflected
θ2 light

(a) (b) (c)

reflected
light
observer

D transmitted
n1
θ1 light
A C
θ2
n2 d

B
(d)
(a)

Figure 8: In figure (a) light moving through a medium with index of refraction n1 is incident on a
surface with a thin-film of thickness d and index of refraction n2 . In figure (b) part of the wave is
reflected with an angle θ1 and another part is transmitted with an angle θ2 . From Snell’s law we
know that n1 sin θ1 = n2 sin θ2 . In figure (c) the transmitted light is reflected from the bottom of
the thin-film. Finally, figure (d) shows the light ray that went through the thin-film emerging and
interfering with the first light-ray that was reflected directly from the top.

67
The transmitted light moves through the thin-film and reflects off the bottom surface with an angle
θ2 as shown in Fig. (8c) (the light is partly transmitted through, but that portion is irrelevant for
our purpose). That reflected ray then refracts from the thin-film back to air and emerges parallel
to the original reflected ray, see Fig. (8d). The two emerging rays both arrive at your eye and
they interfere. We know that whether they interfere constructively or destructively depends on the
difference between their phase. What is the difference between their phase then?

Looking at Fig. (8d) it is clear that any difference in phase between the two rays is due to the phase
difference between the path AD and the path ABC. We therefore have that,
φABC − φAD = k 0 ABC − kAD
2π 2π
= 0
ABC − AD , (7.46)
λ λ
where λ is the wavelength of light in air and
n1
λ0 = λ (7.47)
n2
is the wavelength of light in the medium. With the aid of some trigonometry we can express the
two paths in terms of the thickness of the thin-film, d, and the angles θ1 and θ2 as follows,
d
ABC = 2 , (7.48)
cos θ2
AD = (2d tan θ2 ) sin θ1 . (7.49)
Using these expressions for the paths’ length Eqs. (7.48) and (7.49) in Eq. (7.46) together with
Snell’s law, Eq. (7.45), and the wavelength of light in the thin-film, Eq. (7.47), we arrive at,
 
2π n2 d 2π
φABC − φAD = 2 − (2d tan θ2 sin θ1 )
λ n1 cos θ2 λ
sin2 θ2
 
2π n2 1
= 2d −
λ n1 cos θ2 cos θ2
2π n2
= 2d cos θ2 . (7.50)
λ n1
We are almost done. What we just calculated is the difference in phase due to the optical path
length difference. But, both rays we considered were reflected from some surface and we need
to account for any possible phase-shift that happened. The first ray reflected from the air-oil
boundary and so it underwent a π phase-shift since oil has a higher index of refraction. The second
ray reflected from the oil-air boundary at the bottom of the thin-film and so suffered no phase-shift
since air has a lower index of refraction than oil. Therefore, the total phase difference is,
∆φ = φ2 − φ1 = φABC − (φAD + π)

2π n2
= 2d cos θ2 − π . (7.51)
λ n1

As we already know very well, constructive interference occurs when the phase difference is an
integer multiple of 2π, that is ∆φ = 2πm where m is any integer. Destructive interference happens

68
when the phase difference is a half integer of 2π, that is ∆φ = 2π(m + 21 ). We thus arrive at the
following conditions for constructive and destructive interference of the two rays,
n2 1

2d cos θ2 = m+ 2 λ constructive interference , (7.52)
n1
n2
2d cos θ2 = m0 λ destructive interference , (7.53)
n1

where both m and m0 are any integers.

What do these conditions have anything to do with the colorful patterns displayed by soap bubbles
and other thin-films? Notice that the indices of refraction, n1 and n2 , and the thickness of the thin-
film, d, are all fixed. So for different colors (different wavelength) the condition for constructive
interference, Eq. (7.52), is satisfied at different angles. Similarly for destructive interference. This
is why different patches of the thin-film seem to have different colors - only the colors (wavelengths)
that result in constructive interference would be visible and so as your observation angle changes
the apparent color changes 6 .

6
Your observation angle is related to θ1 , which in turn is related to θ2 through Snell’s law n1 sin θ1 = n2 sin θ2 .

69
8 Fluids

So far in this course we have considered solid objects. We saw how Newton’s laws of motion apply
to solid objects and how they are used to determine the motion of these objects. In this last part of
the course we will consider a certain type of fluids, known as incompressible fluids, and see how the
very same Newton’s laws can be used to determine the behavior of such fluids. Let us commence
then.

8.1 Pressure

The pressure applied on an object is the force per unit area perpendicular to the surface,
F
P = , (8.1)
A
where F is the applied force and A is the area onto which it is applied. The standard unit of
pressure is the Pascal and it is defined as,
N
1 Pa = . (8.2)
m2
Maybe the most familiar example of pressure is atmospheric pressure caused by the weight of the
atmosphere (recall that weight is a force). Air, albeit very light still has a mass. If we consider a
column of air of cross-sectional area of 1 cm2 , starting at sea level and extending all the way up
to the atmosphere, that column weighs W = 10.1 N, and hence that entire column has a mass of
about W/g = 1 kg. It exerts a pressure of 1.01 × 105 Pa, also known as 1 atmosphere.

8.2 Incompressible Fluids

We call materials that easily deform under stress and easily flow so as to take the shape of a
container they are placed in, fluids. Examples of fluids are liquids, gases, and plasmas. Let us
consider a typical fluid that completely fills a closed container. We can measure the mass of the
fluid by measuring the mass of the container with and without it. We can also measure the volume
of the container by measuring its dimensions. This allows us to compute the density of the fluid,
mfluid
ρ= . (8.3)
V
The density is the mass of the fluid divided by its volume. We will deal only with fluids whose
density does not change under applied pressure. Such fluids are known as incompressible fluids. As
a concrete example, think about a balloon filled with water. As you press on it, it may deform, but
you will not change the density of water inside. As an example of a fluid that is not incompressible,
think about the gas in an engine’s piston. As the piston compresses it increases the gas density
inside it. So to recap, incompressible fluids are fluids whose density does not change under pressure.

70
Problems

1. The simplest type of pipettes use atmospheric pressure to draw liquids up the pipette. With
your finger you close off the upper part while immersing the lower part into the liquid you
wish to draw. Explain the mechanism by which the liquid is pushed up the pipette.

2. A cubic meter of water has a mass of 1 Ton. What is the density of water? The density
of Uranium at room temperature is 19.1 g/cm3 . How heavy is a cubic meter of Uranium at
room temperature.

3. Not unlike a pipette, a manual water pump draws water from a well by creating a vacuum
inside the pump that the water are then pushed into. What is pushing the water into and up
the water pump? Assuming it is simply a long cylinder, how high can a water pump be and
still draw water from a well?

71
8.3 Buoyancy

Let’s consider now the interaction of solids with fluids. When a solid object is placed in a liquid,
will it sink or will it float?

Consider the solid at a moment when a part of it, with volume Vl , has been submerged as shown in
Fig. (9a). Now, imagine the solid was not present and replace the submerged part of the solid by
liquid of the same volume, Vl , as shown in Fig. (9b). The force of gravity on that volume of liquid
is simply,

Fgliquid = −md g = −Vl ρl g , (8.4)

where ρl is the density of the liquid. This configuration must be in equilibrium since Vl is the same
liquid and the boundary drawn is entirely fictitious. So the force of gravity must be balanced by
an upward force exerted on this volume of liquid by the rest of the liquid,

FBliquid = Vl ρl g . (8.5)

But, the surrounding liquid could not know whether Vl is liquid or any other material. So, the
upward force exerted on the solid by the surrounding liquid is the same,

FBsolid = Vl ρl g . (8.6)

This force is known as buoyancy. The forces acting on the solid are then the buoyancy, FB = Vl ρl g
pushing it upwards, as well as its weight Fg = −msolid g pulling it downwards. This is shown in
Fig. (9c). The volume, Vl is simply the volume of liquid displaced by the solid. Therefore, a simple
way to state the law of buoyancy is to say that ”Buoyancy is equal to the weight of the liquid
displaced by the solid ”. This is Archimedes’ law.

72
Vl ρl g Vl ρl g

Vl Vl

Fgliquid = −Vl ρl g Fgsolid = −msolid g

(a) (b) (c)

Figure 9: (a) A volume Vl of a solid is submerged in liquid, what are the forces on the solid? (b)
If we imagine replacing the submerged solid by liquid, then the system must be in equilibrium.
Therefore, the upward force exerted on the volume Vl of liquid by the surrounding liquid must
equal its weight. (c) The buoyancy force exerted on the solid due to the surrounding liquid is
therefore FB = Vl ρl g.

Problems

1. A solid made of steel is suspended by a massless string above a bucket of water.

(a) Draw the force diagram on the solid.


(b) What is the tension in the string?

The solid is then lowered into the water until it is entirely submerged.

(c) Draw the force diagram on the solid.


(d) What is the tension in the string?
(e) What is the density of still, in terms of the tension in the string before and after the
solid is submerged?

*A variant of this construction can be used to measure the density of gases. Can you construct
such a device?

73
8.4 Continuity

The conservation of mass implies an important relation between the speed of a fluid and the cross-
sectional area it is flowing through. Consider a long tube with cross-sectional area of A1 on one
end and A2 on the other. Using the usual relation between mass and density, the amount of mass
moving into the tube through a distance dx1 is

dm = ρl A1 dx1 . (8.7)

The rate at which the mass is flowing into the tube is given by,
dm dx1
R1 = = ρl A1 = ρ l A1 v 1 . (8.8)
dt dt
Similarly, the rate at which mass is flowing out of the tube is,

R2 = ρl A2 v2 . (8.9)

Since mass is conserved, the rate at which mass is flowing into the tube must be equal to the rate
at which mass is flowing out of the tube. This is the statement of continuity. Therefore we must
have,

R1 = R2 . (8.10)

In the case of incompressible fluids we consider here, the density is unchanged throughout the tube
and so we can drop it on both sides to arrive at the conservation of volume flow rate,

A1 v 1 = A2 v 2 . (8.11)

Thus, as the fluid moves from a more constricted tube to a large tube it slows down. Similarly, as
the fluid moves from a more spacious tube to a more constricted one it speeds up.

R1 R2

A1 A2

dx1

dx2

Figure 10: Conservation of mass implies that the rate of mass flow through the tube must be
constant. Further, since we are dealing with an incompressible, its density remains constant.
Therefore, the volume flow rate, or A v, is a constant of the motion.

74
8.5 Fluid Statics

Let’s consider the forces acting on a small static cube of liquid, of area A and height h. The forces
and hence the pressure from the horizontal sides must all be equal since the cube is in equilibrium.
In the vertical, the force must include the effect of the weight. At the bottom of the cube the force
is,

Fbottom = A Pb − m g = A Pb − (A h ρl ) g , (8.12)

where Pb is the pressure at that height, pointing upwards and into the cube. At the top it is only
the pressure pointing downwards,

Ftop = −A Pt . (8.13)

Since this volume of liquid is in equilibrium, the forces must cancel,

Ftop + Fbottom = 0 → Pb = Pt + ρl g h . (8.14)

This is known as Pascal’s principle. The pressure of an incompressible fluid in equilibrium is the
same at all points of the same depth and increases linearly with depth. The pressure at greater
depths must be larger in order to hold the weight of the liquid on top. The linear relationship is
maintained as long as the liquid is incompressible and the density does not change.

75
8.6 Fluid Dynamics

There are many aspects of fluid dynamics that can be understood by simply applying the notions
of particle mechanics such as energy, forces, and work. For incompressible fluids in particular, there
is a simple relation between the pressure on the fluid, its height, and its velocity. Let us consider a
small amount of mass entering a tube on one end and exiting on the other. The two ends are not
necessarily at the same height, nor do they necessarily have the same area. Using Eq. (4.16) for
the work done by a force over a distance, the work done on the fluid by the external pressure as it
enters is,

W1 = A1 p1 dx1 , (8.15)

where A1 is the area of the entrance, p1 is the inward pressure, and dx1 is the distance the fluid
was pushed through by the pressure. Similarly, the work done by the external pressure upon exit
is,

W2 = −A2 p2 dx2 , (8.16)

where the minus sign accounts for the opposite direction of the velocity (outwards) as compared
with the pressure (inwards). The total work done on the fluid is then,

Wext = W1 + W2 = A1 p1 dx1 − A2 p2 dx2


= (p1 − p2 ) V . (8.17)

Conservation of volume flow rate implies that over the same time the volume flowing inwards is
equal to the volume flowing outwards and thus A1 dx1 = A2 dx2 = V . As this work is done, there
are changes in the fluid’s kinetic and potential energies. These changes are,

∆K = 1
2m v22 − 21 m v12 , (8.18)
∆U = mg h2 − mg h1 . (8.19)

The relation between the external work done and the change in kinetic and potential energy,
Eq. (4.22), informs us that,

Wext = ∆K + ∆U → (p1 − p2 ) V = 12 m v22 − v12 + mg (h2 − h1 ) .



(8.20)

Dividing through by the volume and rearranging, we arrive at a relation among the pressure,
velocity, and height upon entrance and exit,

p1 + 12 ρl v12 + ρl gh1 = p2 + 12 ρl v22 + ρl gh2 . (8.21)

In other words, the quantity on both sides is unchanged throughout the motion,

p + 12 ρl v 2 + ρl gh = constant . (8.22)

This relation is known as Bernoulli’s equation.

76
Problems

1. Water coming out of the faucet have a circular cross-section with radius r0 and velocity v0 .
The faucet is at height h0 and the water falls freely under gravity. The density of water is
ρw = 103 kg/m3 .

(a) Find the velocity of the water as they fall through a height ∆h (so the present height
is h = h0 − ∆h). Express your answer in terms of the drop in height ∆h, the initial
velocity v0 , and the density of water ρw .
(b) Continuity forces a certain relation between the initial velocity v0 and the velocity at
height h = h − ∆h. Express that relation in terms of the initial radius r0 at height h0
as well as the radius r at height h. Use this relation together with the one you found in
the previous part to express the radius of the cross-section as a function of the drop in
height ∆h.
(c) Plot the function you found in the previous part, that is the radius r vs. the drop in
height ∆h.
(d) Open your faucet slowly and compare your observations to the plot you found in the
previous part. You might want to rotate your plot by 90◦ to facilitate the comparison.

77
A Mathematical Appendix - Trigonometry

Throughout this section, I will use θ and φ to represent angles in general. Angles are measured in
radians unless otherwise stated.

The fundamental relation of trigonometry, which is another way of stating Pythagorus’ Theorem
is

cos2 θ + sin2 θ = 1 . (A-1)

A few important relations of cosines and sines are

cos (θ ± 2π) = cos θ , sin (θ ± 2π) = sin θ periodicity (A-2)


cos(−θ) = cos θ , sin(−θ) = − sin θ symmetry (A-3)
cos π2 − θ = sin θ , sin π2 − θ = cos θ
 
(A-4)

The angle addition formulas state,

cos (θ1 + θ2 ) = cos(θ1 ) cos(θ2 ) − sin(θ1 ) sin(θ2 ) (A-5)

sin (θ1 + θ2 ) = sin(θ1 ) cos(θ2 ) + cos(θ1 ) sin(θ2 ) (A-6)

From these two formulas several others formulas follow. First are the double angle formula you are
probably familiar with

cos 2θ = cos2 θ − sin2 θ , (A-7)


sin 2θ = 2 sin(θ) cos(θ) . (A-8)

Also, by taking linear combinations of the above, you can arrive at the following formulas for the
additions of sines and cosine functions (useful when discussing the interference of waves)

φ1 − φ2
   
φ1 + φ2
sin φ1 + sin φ2 = 2 sin cos , (A-9)
2 2
φ1 − φ2
   
φ1 + φ2
cos φ1 + cos φ2 = 2 cos cos . (A-10)
2 2

Any other formula can be obtained with some of the defining relations of cosines and sines, Eqs. (A-
2)-(A-4).

78
B Mathematical Appendix - Calculus

B.1 Functions

Functions are a mapping from some set of numbers to another set of numbers. The simplest type
of functions are ones that take a single variable, like f (x). This function, f (x), takes as argument
a single number and returns a single number. For example
1
f (x) = kx2 , (B-1)
2
where k is some constant. The independent variable x may have many different meanings and
symbolic representations. Similarly, the function f (x) may be represented by many other symbols.
But, the meaning is always the same: you give it some number x and it will give you some other
number f (x). A few examples are,

f (t) some function of time, t. (B-2)

f (vx ) some function of the velocity along the x-direction, vx . (B-3)

f (x) some function of the position x. (B-4)

You get the point.

B.2 Derivatives

I will introduce the notion of derivatives by discussing a function of time f (t), but everything below
applies equally well to functions of other variables, just copy and paste.

The derivative of a function f (t) is defined as

df f (t + ) − f (t)
= lim . (B-5)
dt →0 

Practically, what you care about is the result, so let’s see some simple cases that will be useful for
you:
df
f (t) = Atn → = (A n) tn−1 , (B-6)
dt

df
f (t) = A cos (ωt + φ) → = −Aω sin (ωt + φ) , (B-7)
dt
df
f (t) = A sin (ωt + φ) → = Aω cos (ωt + φ) . (B-8)
dt

79
Finally, maybe the prettiest function of all, the exponential,
df
f (t) = A eλt → = λA eλt . (B-9)
dt
Notice that for λ = 1 this function is its own derivative! We don’t cover this function in this
course, but I thought I’ll include it for your amusement.

In all the expressions above, all the symbols which are not t or the function itself are some constants
(e.g. the amplitude A, the angular frequency ω, the phase φ, the exponent n, the lifetime λ)

B.3 Integrals

The fundamental theorem of calculus states that if a function f (t) is the derivative of another so
that f (t) = dg(t)/dt, then its definite integral is given by
Z b
f (t)dt = g(b) − g(a) . (B-10)
a

Here are some indefinite integrals for you to enjoy


A n+1
Z
(Atn ) dt = t as long as n 6= −1 . (B-11)
n+1
When n = −1 we arrive at the beautiful natural logarithm function
Z  
A
dt = A ln t . (B-12)
t

For Trigonometric functions the integrals can be easily guessed from the derivatives above, Eqs. (B-
7)
A
Z
A cos (ωt + φ) dt = sin (ωt + φ) , (B-13)
ω
A
Z
A sin (ωt + φ) dt = − cos (ωt + φ) . (B-14)
ω

Finally, the integral of the exponential function is


A λt
Z
A eλt dt = e . (B-15)
λ

If you know the indefinite integral, then you can always evaluate the definite integral. For example,
from Eq. (B-11) we see that the definite integral of a power is
Z b
A n+1 A n+1
(Atn ) dt = b − a as long as n 6= −1 . (B-16)
a n+1 n+1

80

Das könnte Ihnen auch gefallen