You are on page 1of 175

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/304943405

Book Five: Vibrations, Waves and Sounds

Book · September 2013

CITATIONS READS
0 12,558

1 author:

Felix Nwachinemelu Chinedum Anyaegbunam


Federal University Ndufu Alike Ikwo
36 PUBLICATIONS 91 CITATIONS

SEE PROFILE

All content following this page was uploaded by Felix Nwachinemelu Chinedum Anyaegbunam on 22 July 2016.

The user has requested enhancement of the downloaded file.


UNDERSTANDING PHYSICS BOOK 5

Vibrations Waves and Sounds

BY
ANYAEGBUNAM F.N.C. (B.Sc., M.Eng., Ph.D.)

1
Copyright @ 2013

ISBN:978-978-935-377-5

Second Edition, 2013

All rights reserved. No part of this book may

be Reproduced or transmitted in any form or

by any Means without written permission

from the Author or the Publisher.

Published and Printed in Abuja FCT Nigeria

By

SOLID ROCK PRESS

08023542950, 08059459210

2
UNDERSTANDING PHYSICS BOOK 5

Vibrations Waves and Sounds

BY

ANYAEGBUNAM F.N.C. (B.Sc., M.Eng., Ph.D.)

3
TABLE OF CONTENTS

CONTENTS PAGE 4

Acknowledgement 6

Preface 7

Forward 8

CHAPTER ONE 9

Vibrations and Waves 9


1.1: Vibrations 9

1.1.1 Vibrational Motion 9


1.1.2 Properties of Periodic Motion 13

1.1.3 Pendulum Motion 19


1.1.4 Motion of a Mass on a Spring 30

1.2: The Nature of a Wave 39

1.2.1 Waves and Wavelike Motion 39


1.2.2 What is a Wave? 42
1.2.3 Categories of Waves 46

1.3: Properties of a Wave 51

1.3.1 The Anatomy of a Wave 51


1.3.2 Frequency and Period of a Wave 54
1.3.3 Energy Transport and the Amplitude of a Wave 57
1.3.4 The Speed of a Wave 60
1.3.5 The Wave Equation 64

1.4: Behavior of Waves 68

1.4.1 Boundary Behavior 68


1.4.2 Reflection, Refraction, and Diffraction 78
1.4.3 Interference of Waves 81
1.4.4 The Doppler Effect 85

4
1.5: Standing Waves 87

1.5.1 Traveling Waves vs. Standing Waves 87


1.5.2 Formation of Standing Waves 89
1.5.3 Nodes and Anti-nodes 91
1.5.4 Harmonics and Patterns 95
1.5.5 Mathematics of Standing Waves 98

CHAPTER TWO 102

Sound Waves and Music 102


2.1: The Nature of a Sound Wave 102

2.1.1 Sound is a Mechanical Wave 102


2.1.2 Sound is a Longitudinal Wave 105
2.1.3 Sound is a Pressure Wave 106

2.2: Sound Properties and Their Perception 109

2.2.1 Pitch and Frequency 109


2.2.2 Intensity and the Decibel Scale 112
2.2.3 The Speed of Sound 116
2.2.4 The Human Ear 121

2.3: Behavior of Sound Waves 123

2.3.1 Interference and Beats 123


2.3.2 The Doppler Effect and Shock Waves 129
2.3.3 Boundary Behavior 133
2.3.4 Reflection, Refraction, and Diffraction 135

2.4: Resonance and Standing Waves 139

2.4.1 Natural Frequency 139


2.4.2 Forced Vibration 142
2.4.3 Standing Wave Patterns 145
2.4.4 Fundamental Frequency and Harmonics 149

2.5: Musical Instruments 156

2.5.1 Resonance 156


2.5.2 Guitar Strings 159
2.5.3 Open-End Air Columns 163
2.5.4 Closed-End Air Columns 169

5
ACKNOWLEDGEMENT

I am thankful to the Almighty God, His son our Lord Jesus Christ and The Holy Spirit who gave me the
strength and courage to embark on this research in spite of obvious challenges in this trying period of my
life, for Devine inspiration and good health throughout the long period dedicated to writing the
UNDERSTANDING PHYSICS SERIES.

I am sincerely grateful to Prof. S. Okoye, who in spite of his busy schedule, found time to painstakingly
read through the understanding physics series and graciously wrote the forward. I am grateful to my wife,
Dr. Mrs. Ngozi Anyaegbunam, and my son Felix Jr. Chinazam Anyaegbunam for their useful suggestion.
I also appreciate my daughters, Zikora Ada, Tito Chinenye, Mary Gift, Akachukwu Miracle, and David
for their understanding and support during the period of writing this book.

6
PREFACE TO BOOK 5

The need to provide students with a textbook that will not only make the study of Physics simple
but also interesting is the driving force that gave rise to this book. The book is tailored to meet
the needs of both beginners and advanced students of Physics. It provides a seamless transition
from secondary school to the University undergraduate Physics with very clear understanding of
the subject matter. Even a student who has never entered the Physics classroom will find this
book very interesting and easy to understand. Every Child who has the ambition to make a career
in any field of Science, technology, Engineering and Medicine should, as a matter of priority,
possess a copy of this book for a clearer understanding of the subject of physics. Understanding
Physics is filled with explanations of various concepts, practical examples, every-day life
experiences in our physical world, practice questions and numerous exercises to help the
students’ skills development.

The aim of this book is to enable students:


A. Have a seamless transition from senior secondary school Physics to university
undergraduate Physics.
B. Have basic knowledge of Physics that will help them in other fields of science,
technology, engineering & medicine.
C. Develop a sense of enthusiasm for Physics, and appreciate its applications in different
contexts.
D. Develop the ability to apply their knowledge and skills in Physics and Electronics to
the solutions of theoretical and practical problems in all areas of Science &
Technology.
E. Acquire knowledge and skill base for further studies in specialized areas of Physics or
multi-disciplinary areas of science & technology.
F. Acquire knowledge and skill base for studies in future chosen special areas e.g.,
science, technology, engineering and medicine.

The book commenced in chapter one with waves and vibrations, introducing the
periodic/sinusoidal motion, describing the nature of waves, various wave pattern and properties
of waves. Standing wave, energy transport and Doppler effects were clearly discussed including
harmonics. Mathematical analysis of waves, reflection, refraction and diffraction of waves were
simplified. Chapter two introduces Sound waves, properties of sound and perception. Wave
behavior, Longitudinal wave as sound wave and music, wave interference and beats are
explained. Boundary behavior of sound wave, pitch and frequency are discussed. The human ear
is also illustrated.

Indeed it is a complete physics book in the series that provide the students with a clear
understanding of all branches of physics.

Anyaegbunam F.N.C.

7
FORWARD

I have seen many Physics books by many writers but I am yet to see physics books written with
as much clarity, readable and understandable as this text. The writer started from the basics and
created a clearer picture of the subject which makes the reader feel a home with the books in the
series. I think that the books should be recommended not only to the University undergraduates-
levels one and two, but also to senior secondary school students preparing to enter the university
to read Science, Engineering and Medicine. This book will definitely encourage more students
to love and venture into physics and indeed provide the critical mass required for research in
science and engineering. The foundation created by the books shall lead to production of future
Nobel Prize winners in the country. The writer is creative and simplistic in his approach of
Physics which makes it readable and understandable by anyone who can read and write. Physics
has never been made so easy that even our secondary school beginners of physics will find the
books very interesting and indeed will treasure them. The university levels one and two
undergraduates who usually run away from physics lectures shall find out from the books that
Physics is, after all, not as difficult as they thought, except those who had the opportunity of
reading the books in their secondary schools.

The writer has a first class brain and has indeed brought the subject of physics down to earth for
average readers to understand the subject. I further recommend that various libraries in the
country, especially in the universities and secondary schools should order the books in series for
their audience and students.

In the series, you have Books 1 to 6 and each creates a clearer understanding of the subject
matter. Book 1 presents motion and forces in the simplest understandable format. Various ways
of describing motion and the laws governing motion are presented and simplified. Satellite
motions and universal gravitation are simplified. Various forces and modes of operation and
equilibrium as well as real life applications are described. Momentum and its Conservation are
properly described. Book 2 presents modern electricity and magnetism in simplified details
beginning with electrostatics and current electricity and concluding with modern electronics and
magnetism. Book 3 deals thermal physics, energy and power. Book 4 presents the solutions to all
the exercises and problems encountered in all the books. Book 5 presents light waves and optics
in a very readable and understandable format while Book 6 deals with vibrations and waves. The
six books cover all aspects of physics in simplified and clear format with illustrations necessary
for the children to follow and understand.

I sincerely recommend that everyone should read all the books in the series to have a better
perspective of physics.

Professor S. Okoye.

8
CHAPTER ONE
Vibration and Waves

1.1 Vibrations
Vibrational Motion | Properties of Periodic Motion | Pendulum Motion | Motion of a Mass on a
Spring

1.1.1 Vibrational Motion


Things wiggle. They do the back and forth. They vibrate; they shake; they oscillate. These
phrases describe the motion of a variety of objects. They even describe the motion of matter at
the atomic level. Even atoms wiggle - they do the back and forth. Wiggles, vibrations, and
oscillations are an inseparable part of nature. In this chapter, we will make an effort to
understand vibrational motion and its relationship to waves. An understanding of vibrations and
waves is essential to understanding our physical world. Much of what we see and hear is only
possible because of vibrations and waves. We see the world around us because of light waves.
And we hear the world around us because of sound waves. If we can understand waves, then we
will be able to understand the world of sight and sound.

Bobble head Dolls - An Example of a Vibrating Object

To begin our ponderings of vibrations and waves, consider one of those


crazy bobble head dolls that you've likely seen at baseball stadiums or
novelty shops, Fig. 10.1. A bobble head doll consists of an oversized replica
of a person's head attached by a spring to a body and a stand. A light tap to
the oversized head causes it to bobble. The head wiggles; it vibrates; it
oscillates. When pushed or somehow disturbed, the head does the back and
forth. The back and forth doesn't happen forever. Over time, the vibrations
tend to die off and the bobble head stops bobbing and finally assumes its
usual resting position. Fig. 1.1

The bobble head doll is a good illustration of many of the principles of vibrational motion. Think
about how you would describe the back and forth motion of the oversized head of a bobble head
doll. What words would you use to describe such a motion? How does the motion of the bobble
head change over time? How does the motion of one bobble head differ from the motion of

9
another bobble head? What
quantities could you measure to describe the motion and so distinguish one motion from another
motion? How would you explain the cause of such a motion? Why does the back and forth
motion of the bobble head finally stop? These are all questions worth pondering and answering if
we are to understand vibrational motion. These are the questions we will attempt to answer in
Section 2 of this chapter.

What Causes Objects to Vibrate?

Like any object that undergoes vibrational motion, the bobble-head has a resting position. The
resting position is the position assumed by the bobble-head when it is not vibrating. The resting
position is sometimes referred to as the equilibrium position. When an object is positioned at its
equilibrium position, it is in a state of equilibrium. As discussed in the Newton's Law, book 1, an
object which is in a state of equilibrium is experiencing a balance of forces. All the individual
forces - gravity, spring, etc. - are balanced or add up to an overall net force of 0 Newtons. When
a bobble head is at the equilibrium position, the forces on the bobble head are balanced. The
bobble head will remain in this position until somehow disturbed from its equilibrium.

Fig. 1.2

If a force is applied to the bobble-head, the equilibrium will be disturbed and the bobble-head
will begin vibrating. We could use the phrase forced to describe the force which sets the
otherwise resting bobble-head into motion. In this case, the force is a short-lived, momentary
force that begins the motion. The bobble-head does its back and forth, repeating the motion over
and over. Each repetition of its back and forth motion is a little less vigorous than its previous
repetition. If the head sways 3 cm to the right of its equilibrium position during the first
repetition, it may only sway 2.5 cm to the right of its equilibrium position during the second
repetition. And it may only sway 2.0 cm to the right of its equilibrium position during the third
repetition. And so on. The extent of its displacement from the equilibrium position becomes less
and less over time. Because the forced vibration that initiated the motion is a single instance of a
short-lived, momentary force, the vibrations ultimately cease. The bobble-head is said to
experience damping. Damping is the tendency of a vibrating object to lose or to dissipate its

10
energy over time. The mechanical energy of the bobbing head is lost to other objects. Without a
sustained forced vibration, the back and forth motion of the bobble-head eventually ceases as
energy is dissipated to other objects. A sustained input of energy would be required to keep the
back and forth motion going. After all, if the vibrating object naturally loses energy, then it must
continuously be put back into the system through a forced vibration in order to sustain the
vibration.

The Restoring Force


A vibrating bobble-head often does the back and forth a number of times. The vibrations repeat
themselves over and over. As such, the bobble-head will move back to (and past) the equilibrium
position every time it returns from its maximum displacement to the right or the left (or above or
below). This begs a question - and perhaps one that you have been thinking of yourself as you've
pondered the topic of vibration. If the forces acting upon the bobble-head are balanced when at
the equilibrium position, then why does the bobble-head sway past this position? Why doesn't
the bobble-head stop the first time it returns to the equilibrium position? The answer to this
question can be found in Newton's first law of motion. Like any moving object, the motion of a
vibrating object can be understood in light of Newton's laws. According to Newton's law of
inertia, an object which is moving will continue its motion if the forces are balanced. Put another
way, forces, when balanced, do not stop moving objects. So every instant in time that the bobble-
head is at the equilibrium position, the momentary balance of forces will not stop the motion.
The bobble-head keeps moving. It moves past the equilibrium position towards the opposite side
of its swing. As the bobble-head is displaced past its equilibrium position, then a force capable of
slowing it down and stopping it exists. This force that slows the bobble-head down as it moves
away from its equilibrium position is known as a restoring force. The restoring force acts upon
the vibrating object to move it back to its original equilibrium position.

Vibrational motion is often contrasted with translational motion. In translational motion, an


object is permanently displaced. The initial force that is imparted to the object displaces it from
its resting position and sets it into motion. Yet because there is no restoring force, the object
continues the motion in its original direction. When an object vibrates, it doesn't move
permanently out of position. The restoring force acts to slow it down, change its direction and
force it back to its original equilibrium position. An object in translational motion is permanently
displaced from its original position. But an object in vibrational motion wiggles about a fixed
position - its original equilibrium position. Because of the restoring
force, vibrating objects do the back and forth. We will explore the
restoring force in more detail later in this Section.

Other Vibrating Systems

As you know, bobble-head dolls are not the only objects that vibrate.
It might be safe to say that all objects in one way or another can be

11
Fig. 1.3
forced to vibrate to some extent. The vibrations might not be large enough to be visible. Or the
amount of damping might be so strong that the object scarcely completes a full cycle of
vibration. But as long as a force persists to restore the object to its original position, a
displacement from its resting position will result in a vibration. Even a large massive skyscraper
is known to vibrate as winds push upon its structure. While held fixed in place at its foundation
(we hope), the winds force the length of the structure out of position and the skyscraper is forced
into vibration.

A pendulum is a classic example of an object that is considered to vibrate. A simple pendulum


consists of a relatively massive object hung by a string from a fixed support. It typically hangs
vertically in its equilibrium position. When the mass is displaced from equilibrium, it begins its
back and forth vibration about its fixed equilibrium position. The motion is regular and
repeating. In the next part of this Section, we will describe such a regular and repeating motion
as a periodic motion. Because of the regular nature of a pendulum's motion, many clocks, such as
grandfather clocks, use a pendulum as part of its timing mechanism.

An inverted pendulum is another classic example of an object that undergoes vibrational motion.
An inverted pendulum is simply a pendulum which has its fixed end located below the vibrating
mass. An inverted pendulum can be made by attaching a mass (such as a tennis ball) to the top
end of a dowel rod and then securing the bottom end of the dowel rod to a horizontal support.
This is shown in the diagram below. A gentle force exerted upon the tennis ball will cause it to
vibrate about a fixed, equilibrium position. The vibrating skyscraper can be thought of as a type
of inverted pendulum. Tall trees are often displaced from their usual vertical orientation by
strong winds. As the winds cease, the trees will vibrate back and forth about their fixed positions.
Such trees can be thought of as acting as inverted pendula. Even the tines of a tuning fork can be
considered a type of inverted pendulum

Fig. 1.4

Another classic example of an object that undergoes vibrational motion is a


mass on a spring. The animation at the right depicts a mass suspended from a

12
Fig. 1.5
spring. The mass hangs at a resting position. If the mass is pulled down, the spring is stretched.
Once the mass is released, it begins to vibrate. It does the back and forth, vibrating about a fixed
position. If the spring is rotated horizontally and the mass is placed upon a supporting surface,
the same back and forth motion can be observed. Pulling the mass to the right of its resting
position stretches the spring. When released, the mass is pulled back to the left, heading towards
its resting position. After passing by its resting position, the spring begins to compress. The
compressions of the coiled spring result in a restoring force that again pushes rightward on the
leftward moving mass. The cycle continues as the mass vibrates back and forth about a fixed
position. The springs inside of a bed mattress, the suspension systems of some cars, and
bathroom scales all operated as a mass on a spring system.

In all the vibrating systems just mentioned, damping is clearly evident. The simple pendulum
doesn't vibrate forever; its energy is gradually dissipated through air resistance and loss of
energy to the support. The inverted pendulum consisting of a tennis ball mounted to the top of a
dowel rod does not vibrate forever. Like the simple pendulum, the energy of the tennis ball is
dissipated through air resistance and vibrations of the support. Frictional forces also cause the
mass on a spring to lose its energy to the surroundings. In some instances, damping is a favored
feature. Car suspension systems are intended to dissipate vibrational energy, preventing drivers
and passengers from having to do the back and forth as they also do the down the road.

Hopefully a lot of our original questions have been answered. But one question that has not yet
been answered is the question pertaining to quantities that can be measured. How can we
quantitatively describe a vibrating object? What measurements can be made of vibrating objects
that would distinguish one vibrating object from another? We will ponder this question in the
next part of this Section on vibrational motion.

1.1.2 Properties of Periodic Motion


A vibrating object is wiggling about a fixed position. Like the mass on a
spring in the animation at the right, a vibrating object is moving over the
same path over the course of time. Its motion repeats itself over and over
again. If it were not for damping, the vibrations would endure forever (or at
least until someone catches the mass and brings it to rest). The mass on the
spring not only repeats the same motion, it does so in a regular fashion. The
time it takes to complete one back and forth cycle is always the same
amount of time. If it takes the mass 3.2 seconds for the mass to complete
the first back and forth cycle, then it will take 3.2 seconds to complete the
seventh back and forth cycle. It's like clockwork. It's so predictable that you
could set your watch by it. In Physics, a motion that is regular and repeating
is referred to as a periodic motion. Most objects that vibrate do so in a
regular and repeated fashion; their vibrations are periodic.

Fig. 1.6
13
The Sinusoidal Nature of a Vibration

Suppose that a motion detector was placed below a vibrating mass on a spring in order to detect
the changes in the mass's position over the course of time. And suppose that the data from the
motion detector could represent the motion of the mass by a position vs. time plot. The graphic
below depicts such a graph. For discussion sake, several points have been labeled on the graph to
assist in the follow-up discussion.

Fig. 1.7

Before reading on, take a moment to reflect on the type of information that is conveyed by the
graph. And take a moment to reflect about what quantities on the graph might be important in
understanding the mathematical description of a mass on a spring. If you've taken time to ponder
these questions, the following discussion will likely be more meaningful.

One obvious characteristic of the graph has to do with its shape. Many students recognize the
shape of this graph from experiences in Mathematics class. The graph has the shape of a sine
wave. If y = sine(x) is plotted on a graphing calculator, a graph with this same shape would be
created. The vertical axis of the above graph represents the position of the mass relative to the
motion detector. A position of about 0.60 m cm above the detector represents the resting position
of the mass. So the mass is vibrating back and forth about this fixed resting position over the
course of time. There is something sinusoidal about the vibration of a mass on a spring. And the
same can be said of a pendulum vibrating about a fixed position or of a guitar string or of the air
inside of a wind instrument. The position of the mass is a function of the sine of the time.

A second obvious characteristic of the graph may be its periodic nature. The motion repeats itself
in a regular fashion. Time is being plotted along the horizontal axis; so any measurement taken
along this axis is a measurement of the time for something to happen. A full cycle of vibration
might be thought of as the movement of the mass from its resting position (A) to its maximum
height (B), back down past its resting position (C) to its minimum position (D), and then back to
its resting position (E). Using measurements from along the time axis, it is possible to determine
the time for one complete cycle. The mass is at position A at a time of 0.0 seconds and completes
its cycle when it is at position E at a time of 2.3 seconds. It takes 2.3 seconds to complete the
first full cycle of vibration. Now if the motion of this mass is periodic (i.e., regular and
repeating), then it should take the same time of 2.3 seconds to complete any full cycle of
vibration. The same time-axis measurements can be taken for the sixth full cycle of vibration. In
the sixth full cycle, the mass moves from a resting position (U) up to V, back down past W to X
and finally back up to its resting position (Y) in the time interval from 11.6 seconds to 13.9

14
seconds. This represents a time of 2.3 seconds to complete the sixth full cycle of vibration. The
two cycle times are identical. Other cycle times are indicated in the table below. By inspection of
the table, one can safely conclude that the motion of the mass on a spring is regular and
repeating; it is clearly periodic. The small deviation from 2.3 s in the third cycle can be
accounted for by the lack of precision in the reading of the graph.

Table 1.1 Spring Displacement and Cycle time

Times at Beginning and Cycle Time


Cycle Letters
End of Cycle (seconds) (seconds)
1st A to E 0.0 sto 2.3 s 2.3
2nd E tp I 2.3 s to 4.6 s 2.3
3rd I to M 4.6 s to 7.0 s 2.4
4th M to Q 7.0 s to 9.3 s 2.3
5th Q to U 9.3 s to 11.6 s 2.3
6th U to Y 11.6 s to 13.9 s 2.3

Students viewing the above graph will often describe the


motion of the mass as "slowing down." It might be too early to
talk in detail about what slowing down means. We will save
the lengthy discussion of the topic for the page later in this
lesson devoted to the motion of a mass on a spring. For now,
let's simply say that over time, the mass is undergoing changes
in its speed in a sinusoidal fashion. That is, the speed of the
mass at any given moment in time is a function of the sine of
the time. As such, the mass will both speed up and slow down
over the course of a single cycle. So to say that the mass is "slowing down" is not entirely
accurate since during every cycle there are two short intervals during which it speeds up. (More
on this later.)

Students who describe the mass as slowing down (and most observant students do describe it this
way) are clearly observing something in the graph features that draws out the "slowing down"
comment. Before we discuss the feature that triggers the "slowing down" comment, we must re-
iterate the conclusion from the previous paragraphs - the time to complete one cycle of vibration
is NOT changing. It took 2.3 seconds to complete the first cycle and 2.3 seconds to complete the
sixth cycle. Whatever "slowing down" means, we must refute the notion that it means that the
cycles are taking longer as the motion continues. This notion is clearly contrary to the data.

A third obvious characteristic of the graph is that damping occurs with the mass-spring system.
Some energy is being dissipated over the course of time. The extent to which the mass moves
above (B, F, J, N, R and V) or below (D, H, L, P, T and X) the resting position (C, E, G, I, etc.)
varies over the course of time. In the first full cycle of vibration being shown, the mass moves

15
from its resting position (A) 0.60 m above the motion detector to a high position (B) of 0.99 m
cm above the motion detector. This is a total upward displacement of 0.29 m. In the sixth full
cycle of vibration that is shown, the mass moves from its resting position (U) 0.60 m above the
motion detector to a high position (V) 0.94 m above the motion detector. This is a total upward
displacement of 0.24 m cm. The table below summarizes displacement measurements for several
other cycles displayed on the graph.

Table1.2: Displacement for several cycles

Maximum Upward Maximum Downward


Cycle Letters
Displacement Displacement
1st A to E 0.60 m to 0.99 m 0.60 m to 0.21 m
2nd E to I 0.60 m to 0.98 m 0.60 m to 0.22 m
3rd I to M 0.60 m to 0.97 m 0.60 m to 0.23 m
6th U to Y 0.60 m to 0.94 m 0.60 m to 0.26 m

Over the course of time, the mass continues to vibrate - moving away from and back towards the
original resting position. However, the amount of displacement of the mass at its maximum and
minimum height is decreasing from one cycle to the next. This illustrates that energy is being
lost from the mass-spring system. If given enough time, the vibration of the mass will eventually
cease as its energy is dissipated.

Perhaps, this observation of energy dissipation or


energy loss is the observation that triggers the "slowing
down" comment discussed earlier. In physics (or at
least in the English language), "slowing down" means
to "get slower" or to "lose speed". Speed, a physics
term, refers to how fast or how slow an object is
moving. To say that the mass on the spring is "slowing
down" over time is to say that its speed is decreasing
over time. But as mentioned (and as will be discussed
in great detail later), the mass speeds up during two
intervals of every cycle. As the restoring force pulls the
mass back towards its resting position (for instance, from B to C and from D to E), the mass
speeds up. For this reason, a physicist adopts a different language to communicate the idea that
the vibrations are "dying out". We use the phrase "energy is being dissipated or lost" instead of
saying the "mass is slowing down." Language is important when it comes to learning physics.
And sometimes, faulty language (combined with surface-level thinking) can confuse a student of
physics who is sincerely trying to learn new ideas.

16
Period and Frequency
So far in this part of the lesson, we have looked at measurements of time and position of a mass
on a spring. The measurements were based upon readings of a position-time graph. The data on
the graph was collected by a motion detector that was capturing a history of the motion over the
course of time. The key measurements that have been made are measurements of:

• the time for the mass to complete a cycle, and


• the maximum displacement of the mass above (or below) the resting position.

These two measurable quantities have names. We call these quantities period and amplitude.

An object that is in periodic motion - such as a mass on a spring,


a pendulum or a bobblehead doll - will undergo back and forth
vibrations about a fixed position in a regular and repeating
fashion. The fact that the periodic motion is regular and
repeating means that it can be mathematically described by a
quantity known as
the period. The
period of the object's
motion is defined as
the time for the
object to complete one full cycle. Being a time, the
period is measured in units such as seconds,
milliseconds, days or even years. The standard metric
unit for period is the second.

An object in periodic motion can have a long period


or a short period. For instance, a pendulum bob tied to a 1-meter length string has a period of
about 2.0 seconds. For comparison sake, consider the vibrations of a piano string that plays the
middle C note (the C note of the fourth octave). Its period is approximately 0.0038 seconds (3.8
milliseconds). When comparing these two vibrating objects - the 1.0-meter length pendulum and
the piano string which plays the middle C note - we would describe the piano string as vibrating
relatively frequently and we would describe the pendulum as vibrating relatively infrequently.
Observe that the description of the two objects uses the terms frequently and infrequently. The
terms fast and slow are not used since physics types reserve the words fast and slow to refer to an
object's speed. Here in this description we are referring to the frequency, not the speed. An object
can be in periodic motion and have a low frequency and a high speed. As an example, consider
the periodic motion of the moon in orbit about the earth. The moon moves very fast; its orbit is
highly infrequent. It moves through space with a speed of about 1000 m/s - that's fast. Yet it
makes a complete cycle about the earth once every 27.3 days (a period of about 2.4x105 seconds)
- that's infrequent.

Objects like the piano string that have a relatively short period (i.e., a low value for period) are
said to have a high frequency. Frequency is another quantity that can be used to quantitatively
describe the motion of an object is periodic motion. The frequency is defined as the number of

17
complete cycles occurring per period of time. Since the standard metric unit of time is the
second, frequency has units of cycles/second. The unit cycles/second is equivalent to the unit
Hertz (abbreviated Hz). The unit Hertz is used in honor of Heinrich Rudolf Hertz, a 19th century
physicist who expanded our understanding of the electromagnetic theory of light waves.

The concept and quantity frequency is best understood if you attach it to the everyday English
meaning of the word. Frequency is a word we often use to describe how often something occurs.
You might say that you frequently check your email or you frequently talk to a friend or you
frequently wash your hands when working with chemicals. Used in this context, you mean that
you do these activities often. To say that you frequently check your email means that you do it
several times a day - you do it often. In physics, frequency is used with the same meaning - it
indicates how often a repeated event occurs. High frequency events that are periodic occur often,
with little time in between each occurrence - like the back and forth vibrations of the tines of a
tuning fork. The vibrations are so frequent that they can't be seen with the naked eye. A 256-Hz
tuning fork has tines that make 256 complete back and forth vibrations each second. At this
frequency, it only takes the tines about 0.00391 seconds to complete one cycle. A 512-Hz tuning
fork has an even higher frequency. Its vibrations occur more frequently; the time for a full cycle
to be completed is 0.00195 seconds. In comparing these two tuning forks, it is obvious that the
tuning fork with the highest frequency has the lowest period. The two quantities frequency and
period are inversely related to each other. In fact, they are mathematical reciprocals of each
other. The frequency is the reciprocal of the period and the period is the reciprocal of the
frequency.

This reciprocal relationship is easy to understand. After all, the two quantities are conceptual
reciprocals (a phrase I made up). Consider their definitions as restated below:

• period = the time for one full cycle to complete itself; i.e., seconds/cycle
• frequency = the number of cycles that are completed per time; i.e., cycles/second

Even the definitions have a reciprocal ring to them. To understand the distinction between
period and frequency, consider the following statement:

QUIZ 10.1 According to Wikipedia (and as of this writing), Tim Ahlstrom of Oconomowoc,
WI holds the record for hand clapping. He is reported to have clapped his hands 793 times in
60.0 seconds.
What is the frequency and what is the period of Mr. Ahlstrom's hand clapping during this 60.0-
second period?

18
Amplitude of Vibration

The final measurable quantity that describes a vibrating object is the amplitude. The amplitude is
defined as the maximum displacement of an object from its resting position. The resting position
is that position assumed by the object when not vibrating. Once vibrating, the object oscillates
about this fixed position. If the object is a mass on a spring (such as the discussion earlier on this
page), then it might be displaced a maximum distance of 35 cm below the resting position and 35
cm above the resting position. In this case, the amplitude of motion is 35 cm.

Over the course of time, the amplitude of a vibrating object tends to become less and less. The
amplitude of motion is a reflection of the quantity of energy possessed by the vibrating object.
An object vibrating with a relatively large amplitude has a relatively large amount of energy.
Over time, some of this energy is lost due to damping. As the energy is lost, the amplitude
decreases. If given enough time, the amplitude decreases to 0 as the object finally stops
vibrating. At this point in time, it has lost all its energy.

Exercise 1.1
1. A pendulum is observed to complete 23 full cycles in 58 seconds. Determine the period and
the frequency of the pendulum.

2. A mass is tied to a spring and begins vibrating periodically. The distance between its highest
and its lowest position is 38 cm. What is the amplitude of the vibrations?

1.1.3 Pendulum Motion


A simple pendulum consists of a relatively massive object
hung by a string from a fixed support. It typically hangs
vertically in its equilibrium position. The massive object is
affectionately referred to as the pendulum bob. When the bob
is displaced from equilibrium and then released, it begins its
back and forth vibration about its fixed equilibrium position.
The motion is regular and repeating, an example of periodic
motion. Pendulum motion was introduced earlier in this

19
Section as we made an attempt to understand the nature of vibrating objects. Pendulum motion
was discussed again as we looked at the mathematical properties of objects that are in periodic
motion. Here we will investigate pendulum motion in even greater detail as we focus upon how a
variety of quantities change over the course of time. Such quantities will include forces, position,
velocity and energy - both kinetic and potential energy.

Force Analysis of a Pendulum


Earlier in this Section we learned that an object that is vibrating is acted upon by a restoring
force. The restoring force causes the vibrating object to slow down as it moves away from the
equilibrium position and to speed up as it approaches the equilibrium position. It is this restoring
force that is responsible for the vibration. So what forces act upon a pendulum bob? And what is
the restoring force for a pendulum? There are two dominant forces acting upon a pendulum bob
at all times during the course of its motion. There is the force of gravity that acts downward upon
the bob. It results from the Earth's mass attracting the mass of the bob. And there is a tension
force acting upward and towards the pivot point of the pendulum. The tension force results from
the string pulling upon the bob of the pendulum. In our discussion, we will ignore the influence
of air resistance - a third force that always opposes the motion of the bob as it swings to and fro.
The air resistance force is relatively weak compared to the two dominant forces.

The gravity force is highly predictable; it is always in the same direction (down) and always of
the same magnitude - mass*9.8 N/kg. The tension force is considerably less predictable. Both its
direction and its magnitude change as the bob swings to and fro. The direction of the tension
force is always towards the pivot point. So as the bob swings to the left of its equilibrium
position, the tension force is at an angle - directed upwards and to the right. And as the bob
swings to the right of its equilibrium position, the tension is directed upwards and to the left. The
diagram below depicts the direction of these two forces at five different positions over the course
of the pendulum's path.

Fig. 1.8
Fig. 1.9
In physical situations in which the forces acting on an object are not in the same, opposite or
perpendicular directions, it is customary to resolve one or more of the forces into components.

20
This was the practice used in the analysis of sign hanging problems and inclined plane problems.
Typically one or more of the forces are resolved into perpendicular components that lie along
coordinate axes that are directed in the direction of the acceleration or perpendicular to it. So in
the case of a pendulum, it is the gravity force which gets resolved since the tension force is
already directed perpendicular to the motion. The diagram at the right shows the pendulum bob
at a position to the right of its equilibrium position and midway to the point of maximum
displacement. A coordinate axis system is sketched on the diagram and the force of gravity is
resolved into two components that lie along these axes. One of the components is directed
tangent to the circular arc along which the pendulum bob moves; this component is labeled
Fgrav-tangent. The other component is directed perpendicular to the arc; it is labeled Fgrav-perp.
You will notice that the perpendicular component of gravity is in the opposite direction of the
tension force. You might also notice that the tension force is slightly larger than this component
of gravity. The fact that the tension force (Ftens) is greater than the perpendicular component of
gravity (Fgrav-perp) means there will be a net force which is perpendicular to the arc of the bob's
motion. This must be the case since we expect that objects that move along circular paths will
experience an inward or centripetal force. The tangential component of gravity (Fgrav-tangent) is
unbalanced by any other force. So there is a net force directed along the other coordinate axes. It
is this tangential component of gravity which acts as the restoring force. As the pendulum bob
moves to the right of the equilibrium position, this force component is directed opposite its
motion back towards the equilibrium position.

The above analysis applies for a single location along the pendulum's arc. At the other locations
along the arc, the strength of the tension force will vary. Yet the process of resolving gravity into
two components along axes that are perpendicular and tangent to the arc remains the same. The
diagram below shows the results of the force analysis for several other positions.

Fig. 1.10

There are a couple comments to be made. First, observe the diagram for when the bob is
displaced to its maximum displacement to the right of the equilibrium position. This is the
position in which the pendulum bob momentarily has a velocity of 0 m/s and is changing its
direction. The tension force (Ftens) and the perpendicular component of gravity (Fgrav-perp)
balance each other. At this instant in time, there is no net force directed along the axis that is
perpendicular to the motion. Since the motion of the object is momentarily paused, there is no
need for a centripetal force.

21
Second, observe the diagram for when the bob is at the equilibrium position (the string is
completely vertical). When at this position, there is no component of force along the tangent
direction. When moving through the equilibrium position, the restoring force is momentarily
absent. Having been restored to the equilibrium position, there is no restoring force. The
restoring force is only needed when the pendulum bob has been displaced away from the
equilibrium position. You might also notice that the tension force (Ftens) is greater than the
perpendicular component of gravity (Fgrav-perp) when the bob moves through this equilibrium
position. Since the bob is in motion along a circular arc, there must be a net centripetal force at
this position.

The Sinusoidal Nature of Pendulum Motion


In the previous sub-section, we investigated the
sinusoidal nature of the motion of a mass on a spring.
We will conduct a similar investigation here for the
motion of a pendulum bob. Let's suppose that we could
measure the amount that the pendulum bob is
displaced to the left or to the right of its equilibrium or
rest position over the course of time. A displacement
to the right of the equilibrium position would be
regarded as a positive displacement; and a Fig. 1.11
displacement to the left would be regarded as a
negative displacement. Using this reference frame, the equilibrium position would be regarded as
the zero position. And suppose that we constructed a plot showing the variation in position with
respect to time. The resulting position vs. time plot is shown below. Similar to what was
observed for the mass on a spring, the position of the pendulum bob (measured along the arc
relative to its rest position) is a function of the sine of the time.

Fig. 1.12

Now suppose that we use our motion detector to investigate the how the velocity of the
pendulum changes with respect to the time. As the pendulum bob does the back and forth, the
velocity is continuously changing. There will be times at which the velocity is a negative value
(for moving leftward) and other times at which it will be a positive value (for moving rightward).
And of course there will be moments in time at which the velocity is 0 m/s. If the variations in
velocity over the course of time were plotted, the resulting graph would resemble the one shown
below.

22
Fig. 1.13

Now let's try to understand the relationship between the position of the bob along the arc of its
motion and the velocity with which it moves. Suppose we identify several locations along the arc
and then relate these positions to the velocity of the pendulum bob. The graphic below shows an
effort to make such a connection between position and velocity.

Fig. 1.14
As is often said, a picture is worth
a thousand words. Now here come the words. The plot above is based upon the equilibrium
position (D) being designated as the zero position. A displacement to the left of the equilibrium
position is regarded as a negative position. A displacement to the right is regarded as a positive
position. An analysis of the plots shows that the velocity is least when the displacement is
greatest. And the velocity is greatest when the displacement of the bob is least. The further the
bob has moved away from the equilibrium position, the slower it moves; and the closer the bob is
to the equilibrium position, the faster it moves. This can be explained by the fact that as the bob
moves away from the equilibrium position, there is a restoring force that opposes its motion.
This force slows the bob down. So as the bob moves leftward from position D to E to F to G, the
force and acceleration is directed rightward and the velocity decreases as it moves along the arc
from D to G. At G - the maximum displacement to the left - the pendulum bob has a velocity of 0
m/s. You might think of the bob as being momentarily paused and ready to change its direction.
Next the bob moves rightward along the arc from G to F to E to D. As it does, the restoring force
is directed to the right in the same direction as the bob is
moving. This force will accelerate the bob, giving it a
maximum speed at position D - the equilibrium position.
As the bob moves past position D, it is moving
rightward along the arc towards C, then B and then A.
As it does, there is a leftward restoring force opposing
its motion and causing it to slow down. So as the Fig. 1.15

23
displacement increases from D to A, the speed decreases due to the opposing force. Once the bob
reaches position A - the maximum displacement to the right - it has attained a velocity of 0 m/s.
Once again, the bob's velocity is least when the displacement is greatest. The bob completes its
cycle, moving leftward from A to B to C to D. Along this arc from A to D, the restoring force is
in the direction of the motion, thus speeding the bob up. So it would be logical to conclude that
as the position decreases (along the arc from A to D), the velocity increases. Once at position D,
the bob will have a zero displacement and a maximum velocity. The velocity is greatest when the
displacement is least.

The acceleration vector that is shown combines both the perpendicular and the tangential
accelerations into a single vector. You will notice that this vector is entirely tangent to the arc
when at maximum displacement; this is consistent with the force analysis discussed above. And
the vector is vertical (towards the center of the arc) when at the equilibrium position. This also is
consistent with the force analysis discussed above.

Energy Analysis
In a previous chapter of This Book, the energy possessed by a pendulum bob was discussed. We
will expand on that discussion here as we make an effort to associate the motion characteristics
described above with the concepts of kinetic energy, potential energy and total mechanical
energy.

The kinetic energy possessed by an object is the energy it possesses due to its motion. It is a
quantity that depends upon both mass and speed. The equation that relates kinetic energy (KE) to
mass (m) and speed (v) is

KE = ½•m•v2

The faster an object moves, the more kinetic energy that it will possess. We can combine this
concept with the discussion above about how speed changes during the course of motion. This
blending of concepts would lead us to conclude that the kinetic energy of the pendulum bob
increases as the bob approaches the equilibrium position. And the kinetic energy decreases as the
bob moves further away from the equilibrium position.

24
Fig. 1.16

The potential energy possessed by an object is the stored energy of position. Two types of
potential energy are discussed in The Book - gravitational potential energy and elastic potential
energy. Elastic potential energy is only present when a spring (or other elastic medium) is
compressed or stretched. A simple pendulum does not consist of a spring. The form of potential
energy possessed by a pendulum bob is gravitational potential energy. The amount of
gravitational potential energy is dependent upon the mass (m) of the object and the height (h) of
the object. The equation for gravitational potential energy (PE) is

PE = m•g•h

where g represents the gravitational field strength (sometimes referred to as the acceleration
caused by gravity) and has the value of 9.8 N/kg.

The height of an object is expressed relative to some arbitrarily assigned zero level. In other
words, the height must be measured as a vertical distance above some reference position. For a
pendulum bob, it is customary to call the lowest position the reference position or the zero level.
So when the bob is at the equilibrium position (the lowest position), its height is zero and its
potential energy is 0 J. As the pendulum bob does the back and forth, there are times during
which the bob is moving away from the equilibrium position. As it does, its height is increasing
as it moves further and further away. It reaches a maximum height as it reaches the position of
maximum displacement from the equilibrium position. As the bob moves towards its equilibrium
position, it decreases its height and decreases its potential energy.

Fig. 1.17

Now let's put these two concepts of kinetic energy and potential energy together as we consider
the motion of a pendulum bob moving along the arc shown in the diagram at the right. We will
use an energy bar chart to represent the changes in the two forms of energy. The amount of each
form of energy is represented by a bar. The height of the bar is proportional to the amount of that
form of energy. In addition to the potential energy (PE) bar and kinetic energy (KE) bar, there is
a third bar labeled TME. The TME bar represents the total amount of mechanical energy

25
possessed by the pendulum bob. The total mechanical energy is simply the sum of the two forms
of energy – kinetic plus potential energy. Take some time to inspect the bar charts shown below
for positions A, B, D, F and G. What do you notice?

Fig. 1.18

When you inspect the bar charts, it is evident that as the bob moves from A to D, the kinetic
energy is increasing and the potential energy is decreasing. However, the total amount of these
two forms of energy is remaining constant. Whatever potential energy is lost in going from
position A to position D appears as kinetic energy. There is a transformation of potential energy
into kinetic energy as the bob moves from position A to position D. Yet the total mechanical
energy remains constant. We would say that mechanical energy is conserved. As the bob moves
past position D towards position G, the opposite is observed. Kinetic energy decreases as the bob
moves rightward and (more importantly) upward toward position G. There is an increase in
potential energy to accompany this decrease in kinetic energy. Energy is being transformed from
kinetic form into potential form. Yet, as illustrated by the TME bar, the total amount of
mechanical energy is conserved. This very principle of energy conservation was explained in the
Energy chapter.

The Period of a Pendulum


Our final discussion will pertain to the period of the pendulum. As discussed previously in this
Section, the period is the time it takes for a vibrating object to complete its cycle. In the case of
pendulum, it is the time for the pendulum to start at one extreme, travel to the opposite extreme,
and then return to the original location. Here we will be interested in the question What variables
affect the period of a pendulum? We will concern ourselves with possible variables. The
variables are the mass of the pendulum bob, the length of the string on which it hangs, and the
angular displacement. The angular displacement or arc angle is the angle that the string makes
with the vertical when released from rest. These three variables and their effect on the period are
easily studied and are often the focus of a physics lab in an introductory physics class. The data
table below provides representative data for such a study.

26
Table 1.3

Trial Mass (kg) Length (m) Arc Angle (°) Period (s)
1 0.02- 0.40 15.0 1.25
2 0.050 0.40 15.0 1.29
3 0.100 0.40 15.0 1.28
4 0.200 0.40 15.0 1.24
5 0.500 0.40 15.0 1.26
6 0.200 0.60 15.0 1.56
7 0.200 0.80 15.0 1.79
8 0.200 1.00 15.0 2.01
9 0.200 1.20 15.0 2.19
10 0.200 0.40 10.0 1.27
11 0.200 0.40 20.0 1.29
12 0.200 0.40 25.0 1.25
13 0.200 0.40 30.0 1.26

In trials 1 through 5, the mass of the bob was systematically altered while keeping the other
quantities constant. By so doing, the experimenters were able to investigate the possible effect of
the mass upon the period. As can be seen in these five trials, alterations in mass have little effect
upon the period of the pendulum.

In trials 4 and 6-9, the mass is held constant at 0.200 kg and the arc angle is held constant at 15°.
However, the length of the pendulum is varied. By so doing, the experimenters were able to
investigate the possible effect of the length of the string upon the period. As can be seen in these
five trials, alterations in length definitely have an effect upon the period of the pendulum. As the
string is lengthened, the period of the pendulum is increased. There is a direct relationship
between the period and the length.

Finally, the experimenters investigated the possible effect of the arc angle upon the period in
trials 4 and 10-13. The mass is held constant at 0.200 kg and the string length is held constant at
0.400 m. As can be seen from these five trials, alterations in the arc angle have little to no effect
upon the period of the pendulum.

So the conclusion from such an experiment is that the one variable that effects the period of the
pendulum is the length of the string. Increases in the length lead to increases in the period. But
the investigation doesn't have to stop there. The quantitative equation relating these variables can
be determined if the data is plotted and linear regression analysis is performed. The two plots

27
below represent such an analysis. In each plot, values of period (the dependent variable) are
placed on the vertical axis. In the plot on the left, the length of the pendulum is placed on the
horizontal axis. The shape of the curve indicates some sort of power relationship between period
and length. In the plot on the right, the square root of the length of the pendulum (length to the ½
power) is plotted. The results of the regression analysis are shown.

Slope: 1.7536 Slope: 2.0045


Fig. 1.19 Y-intercept: 0.2616 Y-intercept: 0.0077 Fig. 1.20
COR: 0.9183 COR: 0.9999

The analysis shows that there is a better fit of the data and the regression line for the graph on the
right. As such, the plot on the right is the basis for the equation relating the period and the length.
For this data, the equation is

Period = 2.0045•Length0.5 + 0.0077

Using T as the symbol for period and L as the symbol for length, the equation can be rewritten as

T = 2.0045•L0.5 + 0.0077

The commonly reported equation based on theoretical development is

T = 2•π•(L/g)0.5

where g is a constant known as the gravitational field strength or the acceleration of gravity (9.8
N/kg). The value of 2.0045 from the experimental investigation agrees well with what would be
expected from this theoretically reported equation. Substituting the value of g into this equation,
yields a proportionality constant of 2π/g0.5 , which is 2.0071, very similar to the 2.0045
proportionality constant developed in the experiment.

28
Exercise 1.2
1. A pendulum bob is pulled back to position A and released from rest. The bob swings through
its usual circular arc and is caught at position C. Determine the position (A, B, C or all the same)
where the …

a. … force of gravity is the greatest?


b. … restoring force is the greatest?
c. … speed is the greatest?
d. … potential energy is the greatest?
e. … kinetic energy is the greatest
f. … total mechanical energy is the greatest?

2. Use energy conservation to fill in the blanks in the following diagram.

Fig. Q1.2.2

3. A pair of trapeze performers at the circus is swinging from ropes attached to a large elevated
platform. Suppose that the performers can be treated as a simple pendulum with a length of 16
m. Determine the period for one complete back and forth cycle.

4. Which would have the highest frequency of vibration?

Pendulum A: A 200-g mass attached to a 1.0-m length string


Pendulum B: A 400-g mass attached to a 0.5-m length string

5. Anna Litical wishes to make a simple pendulum that serves as a timing device. She plans to
make it such that its period is 1.00 second. What length must the pendulum have?

29
1.1.4 Motion of a Mass on a Spring
In a previous sub-section, the motion of a mass attached to a spring was described as an example
of a vibrating system. The mass on a spring motion was discussed in more detail as we sought to
understand the mathematical properties of objects that are in periodic motion. Now we will
investigate the motion of a mass on a spring in even greater detail as we focus on how a variety
of quantities change over the course of time. Such quantities will include forces, position,
velocity and energy - both kinetic and potential energy.

Hooke's Law

We will begin our discussion with an investigation of the forces exerted


by a spring on a hanging mass. Consider the system shown at the right
with a spring attached to a support. The spring hangs in a relaxed,
unstretched position. If you were to hold the bottom of the spring and
pull downward, the spring would stretch. If you were to pull with just a
little force, the spring would stretch just a little bit. And if you were to Fig. 1.21
pull with a much greater force, the spring would stretch a much greater
extent. Exactly what is the quantitative relationship between the amount
of pulling force and the amount of stretch?

To determine this quantitative relationship between the amount of force and the amount of
stretch, objects of known mass could be attached to the spring. For each object which is added,
the amount of stretch could be measured. The force which is applied in each instance would be
the weight of the object. A regression analysis of the force-stretch data could be performed in
order to determine the quantitative relationship between the force and the amount of stretch. The
data table below shows some representative data for such an experiment.

Table 1.4

Mass (kg) Force on Spring (N) Amount of Stretch (m)


0.000 0.000 0.0000
0.050 0.490 0.0021
0.100 0.980 0.0040
0.150 1.470 0.0063
0.200 1.960 0.0081
0.250 2.450 0.0099
0.300 2.940 0.0123
0.400 3.920 0.0160
0.500 4.900 0.0199

30
By plotting the force-stretch data and performing a linear regression analysis, the quantitative
relationship or equation can be determined. The plot is shown below.

Fig. 1.22

A linear regression analysis yields the following statistics:

slope = 0.00406 m/N


y-intercept = 3.43 x10-5 (pert near close to 0.000)
regression constant = 0.999

The equation for this line is

Stretch = 0.00406•Force + 3.43x10-5

The fact that the regression constant is very close to 1.000 indicates that there is
a strong fit between the equation and the data points. This strong fit lends
credibility to the results of the experiment.

This relationship between the force applied to a spring and the amount of
stretch was first discovered in 1678 by English scientist Robert Hooke. As Hooke put it: Ut
tensio, sic vis. Translated from Latin, this means "As the extension, so the force." In other words,
the amount that the spring extends is proportional to the amount of force with which it pulls. If
we had completed this study about 350 years ago (and if we knew some Latin), we would be
famous! Today this quantitative relationship between force and stretch is referred to as Hooke's
law and is often reported in textbooks as

Fspring = -k•x

where Fspring is the force exerted upon the spring, x is the amount that the spring stretches
relative to its relaxed position, and k is the proportionality constant, often referred to as the
spring constant. The spring constant is a positive constant whose value is dependent upon the
spring which is being studied. A stiff spring would have a high spring constant. This is to say
that it would take a relatively large amount of force to cause a little displacement. The units on

31
the spring constant are Newton/meter (N/m). The negative sign in the above equation is an
indication that the direction that the spring stretches is opposite the direction of the force which
the spring exerts. For instance, when the spring was stretched below its relaxed position, x is
downward. The spring responds to this stretching by exerting an upward force. The x and the F
are in opposite directions. A final comment regarding this equation is that it works for a spring
which is stretched vertically and for a spring is stretched horizontally (such as the one to be
discussed below).

Force Analysis of a Mass on a Spring

Earlier in this Section we learned that an object that is vibrating is acted upon by a restoring
force. The restoring force causes the vibrating object to slow down as it moves away from the
equilibrium position and to speed up as it approaches the equilibrium position. It is this restoring
force which is responsible for the vibration. So what is the restoring force for a mass on a spring?

We will begin our discussion of this question by considering the system in the diagram below.

Fig. 1.23
The diagram shows an air track and a glider. The glider is attached by a spring to a vertical
support. There is a negligible amount of friction between the glider and the air track. As such,
there are three dominant forces acting upon the glider. These three forces are shown in the free-
body diagram at the right. The force of gravity (Fgrav) is a rather predictable force - both in
terms of its magnitude and its direction. The force of gravity always acts
downward; its magnitude can be found as the product of mass and the
acceleration of gravity (m•9.8 N/kg). The support force (Fsupport)
balances the force of gravity. It is supplied by the air from the air track,
causing the glider to levitate about the track's surface. The final force is
the spring force (Fspring). As discussed above, the spring force varies in
magnitude and in direction. Its magnitude can be found using Hooke's Fig. 1.24
law. Its direction is always opposite the direction of stretch and towards
the equilibrium position. As the air track glider does the back and forth, the spring force
(Fspring) acts as the restoring force. It acts leftward on the glider when it is positioned to the
right of the equilibrium position; and it acts rightward on the glider when it is positioned to the
left of the equilibrium position.

Let's suppose that the glider is pulled to the right of the equilibrium position and released from
rest. The diagram below shows the direction of the spring force at five different positions over
the course of the glider's path. As the glider moves from position A (the release point) to position

32
B and then to position C, the spring force acts leftward upon the leftward moving glider. As the
glider approaches position C, the amount of stretch of the spring decreases and the spring force
decreases, consistent with Hooke's Law. Despite this decrease in the spring force, there is still an
acceleration caused by the restoring force for the entire span from position A to position C. At
position C, the glider has reached its maximum speed. Once the glider passes to the left of
position C, the spring force acts rightward. During this phase of the glider's cycle, the spring is
being compressed. The further past position C that the glider moves, the greater the amount of
compression and the greater the spring force. This spring force acts as a restoring force, slowing
the glider down as it moves from position C to position D to position E. By the time the glider
has reached position E, it has slowed down to a momentary rest position before changing its
direction and heading back towards the equilibrium position. During the glider's motion from
position E to position C, the amount that the spring is compressed decreases and the spring force
decreases. There is still an acceleration for the entire distance from position E to position C. At
position C, the glider has reached its maximum speed. Now the glider begins to move to the right
of point C. As it does, the spring force acts leftward upon the rightward moving glider. This
restoring force causes the glider to slow down during the entire path from position C to position
D to position E.

Fig. 1.25

33
Sinusoidal Nature of the Motion of a Mass on a Spring
Previously in this Section, the variations in the position of a mass on a spring with respect to
time were discussed. At that time, it was shown that the position of a mass on a spring varies
with the sine of the time. The discussion pertained to a mass that was vibrating up and down
while suspended from the spring. The discussion would be just as applicable to our glider
moving along the air track. If a motion detector were placed at the right end of the air track to
collect data for a position vs. time plot, the plot would look like the plot below. Position A is the
right-most position on the air track when the glider is closest to the detector.

Fig. 1.26

The labeled positions in the diagram above are the same positions used in the discussion of
restoring force above. You might recall from that discussion that positions A and E were
positions at which the mass had a zero velocity. Position C was the equilibrium position and was
the position of maximum speed. If the same motion detector that collected position-time data
were used to collect velocity-time data, then the plotted data would look like the graph below.

Fig. 1.27

Observe that the velocity-time plot for the mass on a spring is also a sinusoidal shaped plot. The
only difference between the position-time and the velocity-time plots is that one is shifted one-
fourth of a vibrational cycle away from the other. Also observe in the plots that the absolute
value of the velocity is greatest at position C (corresponding to the equilibrium position). The
velocity of any moving object, whether vibrating or not, is the speed with a direction. The
magnitude of the velocity is the speed. The direction is often expressed as a positive or a
negative sign. In some instances, the velocity has a negative direction (the glider is moving
leftward) and its velocity is plotted below the time axis. In other cases, the velocity has a positive
direction (the glider is moving rightward) and its velocity is plotted above the time axis. You will
also notice that the velocity is zero whenever the position is at an extreme. This occurs at
positions A and E when the glider is beginning to change direction. So just as in the case of
pendulum motion, the speed is greatest when the displacement of the mass relative to its

34
equilibrium position is the least. And the speed is least when the displacement of the mass
relative to its equilibrium position is the greatest.

Energy Analysis of a Mass on a Spring

On the previous sub-section, an energy analysis for the vibration of a pendulum was discussed.
Here we will conduct a similar analysis for the motion of a mass on a spring. In our discussion,
we will refer to the motion of the frictionless glider on the air track that was introduced above.
The glider will be pulled to the right of its equilibrium position and be released from rest
(position A). As mentioned, the glider then accelerates towards position C (the equilibrium
position). Once the glider passes the equilibrium position, it begins to slow down as the spring
force pulls it backwards against its motion. By the time it has reached position E, the glider has
slowed down to a momentary pause before changing directions and accelerating back towards
position C. Once again, after the glider passes position C, it begins to slow down as it approaches
position A. Once at position A, the cycle begins all over again ... and again ... and again.

The kinetic energy possessed by an object is the energy it possesses due to its motion. It is a
quantity that depends upon both mass and speed. The equation that relates kinetic energy (KE) to
mass (m) and speed (v) is

KE = ½•m•v2

The faster an object moves, the more kinetic energy that it will possess. We can combine this
concept with the discussion above about how speed changes during the course of motion. This
blending of the concepts would lead us to conclude that the kinetic energy of the mass on the
spring increases as it approaches the equilibrium position; and it decreases as it moves away
from the equilibrium position.

This information is summarized in the table below:

Table 1.5

Stage of Cycle Change in Speed Change in Kinetic Energy


A to B to C Increasing Increasing
C to D to E Decreasing Decreasing
E to D to C Increasing Increasing
C to B to A Decreasing Decreasing

35
Kinetic energy is only one form of mechanical energy. The other form is potential energy.
Potential energy is the stored energy of position possessed by an object. The potential energy
could be gravitational potential energy, in which case the position refers to the height above the
ground. Or the potential energy could be elastic potential energy, in which case the position
refers to the position of the mass on the spring relative to the equilibrium position. For our
vibrating air track glider, there is no change in height. So the gravitational potential energy does
not change. This form of potential energy is not of much interest in our analysis of the energy
changes. There is however a change in the position of the mass relative to its equilibrium
position. Every time the spring is compressed or stretched relative to its relaxed position, there is
an increase in the elastic potential energy. The amount of elastic potential energy depends on the
amount of stretch or compression of the spring. The equation that relates the amount of elastic
potential energy (PEspring) to the amount of compression or stretch (x) is

PEspring = ½ • k•x2

where k is the spring constant (in N/m) and x is the distance that the spring is stretched or
compressed relative to the relaxed, unstretched position.

When the air track glider is at its equilibrium position (position C), it is moving it's fastest (as
discussed above). At this position, the value of x is 0 meter. So the amount of elastic potential
energy (PEspring) is 0 Joules. This is the position where the potential energy is the least. When
the glider is at position A, the spring is stretched the greatest distance and the elastic potential
energy is a maximum. A similar statement can be made for position E. At position E, the spring
is compressed the most and the elastic potential energy at this location is also a maximum. Since
the spring stretches as much as compresses, the elastic potential energy at position A (the
stretched position) is the same as at position E (the compressed position). At these two positions
- A and E - the velocity is 0 m/s and the kinetic energy is 0 J. So just like the case of a vibrating
pendulum, a vibrating mass on a spring has the greatest potential energy when it has the smallest
kinetic energy. And it also has the smallest potential energy (position C) when it has the greatest
kinetic energy. These principles are shown in the animation below.

Fig. 1.28

When conducting an energy analysis, a common representation is an energy bar chart. An energy
bar chart uses a bar graph to represent the relative amount and form of energy possessed by an
object as it is moving. It is a useful conceptual tool for showing what form of energy is present
and how it changes over the course of time. The diagram below is an energy bar chart for the air
track glider and spring system.

36
Fig. 1.29

The bar chart reveals that as the mass on the spring moves from A to B to C, the kinetic energy
increases and the elastic potential energy decreases. Yet the total amount of these two forms of
mechanical energy remains constant. Mechanical energy is being transformed from potential
form to kinetic form; yet the total amount is being conserved. A similar conservation of energy
phenomenon occurs as the mass moves from C to D to E. As the spring becomes compressed and
the mass slows down, its kinetic energy is transformed into elastic potential energy. As this
transformation occurs, the total amount of mechanical energy is conserved. This very principle
of energy conservation was explained in a previous book 3 - the Energy chapter.

Period of a Mass on a Spring

As is likely obvious, not all springs are created equal. And not all spring-mass systems are
created equal. One measurable quantity that can be used to distinguish one spring-mass system
from another is the period. As discussed earlier in this lesson, the period is the time for a
vibrating object to make one complete cycle of vibration. The variables that effect the period of a
spring-mass system are the mass and the spring constant. The equation that relates these
variables resembles the equation for the period of a pendulum. The equation is

T = 2•π•(m/k).5

where T is the period, m is the mass of the object attached to the spring, and k is the spring
constant of the spring. The equation can be interpreted to mean that more massive objects will
vibrate with a longer period. Their greater inertia means that it takes more time to complete a
cycle. And springs with a greater spring constant (stiffer springs) have a smaller period; masses
attached to these springs take less time to complete a cycle. Their greater spring constant means
they exert stronger restoring forces upon the attached mass. This greater force reduces the length
of time to complete one cycle of vibration.

37
Looking Forward to Section 2

As we have seen in this section vibrating objects are wiggling in place. They oscillate back and
forth about a fixed position. A simple pendulum and a mass on a spring are classic examples of
such vibrating motion. Though not evident by simple observation, the use of motion detectors
reveals that the vibrations of these objects have a sinusoidal nature. There is a subtle wave-like
behavior associated with the manner in which the position and the velocity vary with respect to
time. In the next lesson, we will investigate waves. As we will soon find out, if a mass on a
spring is a wiggle in time, then a wave is a collection of wigglers spread through space. As we
begin our study of waves in Section 1.2, concepts of frequency, wavelength and amplitude will
remain important.

Exercise 1.3
1. A force of 16 N is required to stretch a spring a distance of 40 cm from its rest position. What
force (in Newtons) is required to stretch the same spring …

a. … twice the distance?


b. … three times the distance?
c. … one-half the distance?

2. Perpetually disturbed by the habit of the backyard squirrels to raid his bird feeders, Mr. H
decides to use a little physics for better living. His current plot involves equipping his bird feeder
with a spring system that stretches and oscillates when the mass of a squirrel lands on the feeder.
He wishes to have the highest amplitude of vibration that is possible. Should he use a spring with
a large spring constant or a small spring constant?

3. Referring to the previous question. If Mr. H wishes to have his bird feeder (and attached
squirrel) vibrate with the highest possible frequency, should he use a spring with a large spring
constant or a small spring constant?

4. Use energy conservation to fill in the blanks in the following diagram.

38
Fig. Q1.3.4

5. Which of the following mass-spring systems will have the highest frequency of vibration?

Case A: A spring with a k=300 N/m and a mass of 200 g suspended from it.
Case B: A spring with a k=400 N/m and a mass of 200 g suspended from it.

6. Which of the following mass-spring systems will have the highest frequency of vibration?

Case A: A spring with a k=300 N/m and a mass of 200 g suspended from it.
Case B: A spring with a k=300 N/m and a mass of 100 g suspended from it.

1.2 The Nature of a Wave


Waves and Wavelike Motion | What is a Wave? | Categories of Waves

1.2.1 Waves and Wavelike Motion


Waves are everywhere. Whether we recognize it or not, we encounter
waves on a daily basis. Sound waves, visible light waves, radio waves,
microwaves, water waves, sine waves, cosine waves, stadium waves,
earthquake waves, waves on a string, and slinky waves and are just a few
of the examples of our daily encounters with waves. In addition to waves,
there are a variety of phenomena in our physical world that resemble
waves so closely that we can describe such phenomenon as being
wavelike. The motion of a pendulum, the motion of a mass suspended by a
spring, the motion of a child on a swing, and the "Hello, Good Morning!"

39
wave of the hand can be thought of as wavelike phenomena. Waves (and wavelike phenomena)
are everywhere!

We study the physics of waves because it provides a rich glimpse into the physical world that we
seek to understand and describe as students of physics. Before beginning a formal discussion of
the nature of waves, it is often useful to ponder the various encounters and exposures that we
have of waves. Where do we see waves or examples of wavelike motion? What experiences do
we already have that will help us in understanding the physics of waves?

For many people, the first thought concerning waves conjures up a picture of a wave moving
across the surface of an ocean, lake, pond or other body of water. The waves are created by some
form of a disturbance, such as a rock thrown into the water, a duck shaking its tail in the water or
a boat moving through the water. The water wave has a crest and a trough and travels from one
location to another. One crest is often followed by a second crest
that is often followed by a third crest. Every crest is separated by a
trough to create an alternating pattern of crests and troughs. A duck
or gull at rest on the surface of the water is observed to bob up-and-
down at rather regular time intervals as the wave passes by. The
waves may appear to be plane waves that travel together as a front
in a straight-line direction, perhaps towards a sandy shore. Or the
waves may be circular waves that originate from the point where
the disturbances occur; such circular waves travel across the surface
of the water in all directions. These mental pictures of water waves
are useful for understanding the nature of a wave and will be
revisited later when we begin our formal discussion of the topic.
Fig. 1.30
The thought of waves often brings to mind a recent encounter at the baseball or football stadium
when the crowd enthusiastically engaged in doing the wave. When performed with reasonably
good timing, a noticeable ripple is produced that travels around the circular stadium or back and
forth across a section of bleachers. The observable ripple results when a group of enthusiastic
fans rise up from their seats, swing their arms up high, and then sit back down. Beginning in
Section 1, the first row of fans abruptly rise up to begin the wave; as they sit back down, row 2
begins its motion; as row 2 sits back down, row 3 begins its motion. The process continues, as
each consecutive row becomes involved by a momentary standing up and sitting back down. The
wave is passed from row to row as each individual member of the row becomes temporarily
displaced out of his or her seat, only to return to it as the wave passes by. This mental picture of a
stadium wave will also provide a useful context for
the discussion of the physics of wave motion.

Another picture of waves involves the movement of a


slinky or similar set of coils. If a slinky is stretched
out from end to end, a wave can be introduced into
the slinky by either vibrating the first coil up and
down vertically or back and forth horizontally. A
wave will subsequently be seen traveling from one
end of the slinky to the other. As the wave moves

40 Fig. 1.31
along the slinky, each individual coil is seen to move out of place and then return to its original
position. The coils always move in the same direction that the first coil was vibrated. A
continued vibration of the first coil results in a continued back and forth motion of the other
coils. If looked at closely, one notices that the wave does not stop when it reaches the end of the
slinky; rather it seems to bounce off the end and head back from where it started. A slinky wave
provides an excellent mental picture of a wave and will be used in discussions and
demonstrations throughout this unit.

We likely have memories from childhood of holding a long jump rope with a friend and
vibrating an end up and down. The up and down vibration of the end of the rope created a
disturbance of the rope that subsequently moved towards the other end. Upon reaching the
opposite end, the disturbance often bounced back to return to the end we were holding. A single
disturbance could be created by the single vibration of one end of the rope. On the other hand, a
repeated disturbance would result in a repeated and regular vibration of the rope. The shape of
the pattern formed in the rope was influenced by the frequency at which we vibrated it. If we
vibrated the rope rapidly, then a short wave was created. And if we vibrated the rope less
frequently (not as often), a long wave was created. While we were likely unaware of it as
children, we were entering the world of the physics of waves as we contentedly played with the
rope.

Then there is the "Hello, Good Morning!" wave. Whether encountered in the driveway as you
begin your trip to school, on the street on the way to school, in the parking lot upon arrival to
school, or in the hallway on the way to your first class, the "Hello, Good Morning!" wave
provides a simple (yet excellent) example of physics in action. The simple back and forth motion
of the hand is called a wave. When Mom commands us to "wave to Mr. Smith," she is telling us
to raise our hand and to temporarily or even repeatedly vibrate it back and forth. The hand is
raised, moved to the left, and then back to the far right and finally returns to its original position.
Energy is put into the hand and the hand begins its back-and-forth vibrational motion. And we
call the process of doing it "waving." Soon we will see how this simple act is representative of
the nature of a physical wave.

We also encountered waves in Math class in the form of the sine and cosine function. We often
plotted y = B•sine(A•x) on our calculator or by hand and observed that its graphical shape
resembled the characteristic shape of a wave. There was a crest and a trough and a repeating
pattern. If we changed the constant A in the equation, we noticed that we could change the length
of the repeating pattern. And if we changed B in the equation, we noticed that we changed the
height of the pattern. In math class, we encountered the underlying mathematical functions that
describe the physical nature of waves.

Finally, we are familiar with microwaves and visible light waves. While we have never seen
them, we believe that they exist because we have witnessed how they carry energy from one
location to another. And similarly, we are familiar with radio waves and sound waves. Like
microwaves, we have never seen them. Yet we believe they exist because we have witnessed the
signals that they carry from one location to another and we have even learned how to tune into
those signals through use of our ears or a tuner on a television or radio. Waves, as we will learn,
carry energy from one location to another. And if the frequency of those waves can be changed,

41
then we can also carry a complex signal that is capable of transmitting an idea or thought from
one location to another. Perhaps this is one of the most important aspects of waves and will
become a focus of our study in later units.

Waves are everywhere in nature. Our understanding of the physical world is not complete until
we understand the nature, properties and behaviors of waves. The goal of this Section is to
develop mental models of waves and ultimately apply those models to an understanding of the
two most common types of waves - sound waves and light waves.

1.2.2 What is a Wave?


So waves are everywhere. But what makes a wave a wave? What characteristics, properties, or
behaviors are shared by the phenomena that we typically characterize as being a wave? How can
waves be described in a manner that allows us to understand their basic nature and qualities?

A wave can be described as a disturbance that travels through a medium from one location to
another location. Consider a slinky wave as an example of a wave. When the slinky is stretched
from end to end and is held at rest, it assumes a natural position known as the equilibrium or
rest position. The coils of the slinky naturally assume this position, spaced equally far apart. To
introduce a wave into the slinky, the first particle is displaced or moved from its equilibrium or
rest position. The particle might be moved upwards or downwards, forwards or backwards; but
once moved, it is returned to its original equilibrium or rest position. The act of moving the first
coil of the slinky in a given direction and then returning it to its equilibrium position creates a
disturbance in the slinky. We can then observe this disturbance moving through the slinky from
one end to the other. If the first coil of the slinky is given a single back-and-forth vibration, then
we call the observed motion of the disturbance through the slinky a slinky pulse. A pulse is a
single disturbance moving through a medium from one location to another location. However, if
the first coil of the slinky is continuously and periodically vibrated in a back-and-forth manner,
we would observe a repeating disturbance moving within the slinky that endures over some
prolonged period of time. The repeating and periodic disturbance that moves through a medium
from one location to another is referred to as a wave.

42
Fig. 1.32

What is a Medium?

But what is meant by the word medium? A medium is a substance or material that carries the
wave. You have perhaps heard of the phrase news media. The news media refers to the various
institutions (newspaper offices, television stations, radio stations, etc.) within our society that
carry the news from one location to another. The news moves through the media. The media
doesn't make the news and the media isn't the same as the news. The news media is merely the
thing that carries the news from its source to various locations. In a similar manner, a wave
medium is the substance that carries a wave (or disturbance) from one location to another. The
wave medium is not the wave and it doesn't make the wave; it merely carries or transports the
wave from its source to other locations. In the case of our slinky wave, the medium through that
the wave travels is the slinky coils. In the case of a water wave in the ocean, the medium through
which the wave travels is the ocean water. In the case of a sound wave moving from the church
choir to the pews, the medium through which the sound wave travels is the air in the room. And
in the case of the stadium wave, the medium through which the stadium wave travels is the fans
that are in the stadium.

Particle-to-Particle Interaction

To fully understand the nature of a wave, it is important to consider the medium as a collection
of interacting particles. In other words, the medium is composed of parts that are capable of
interacting with each other. The interactions of one particle of the medium with the next adjacent
particle allow the disturbance to travel through the medium. In the case of the slinky wave, the
particles or interacting parts of the medium are the individual coils of the slinky. In the case of a
sound wave in air, the particles or interacting parts of the medium are the individual molecules
of air. And in the case of a stadium wave, the particles or interacting parts of the medium are the
fans in the stadium.

43
Consider the presence of a wave in a slinky. The first coil becomes disturbed and begins to push
or pull on the second coil; this push or pull on the second coil will displace the second coil from
its equilibrium position. As the second coil becomes displaced, it begins to push or pull on the
third coil; the push or pull on the third coil displaces it from its equilibrium position. As the third
coil becomes displaced, it begins to push or pull on the fourth coil. This process continues in
consecutive fashion, with each individual particle acting to
displace the adjacent particle. Subsequently, the disturbance
travels through the medium. The medium can be pictured as
a series of particles connected by springs. As one particle
moves, the spring connecting it to the next particle begins to
stretch and apply a force to its adjacent neighbor. As this
neighbor begins to move, the spring attaching this neighbor
to its neighbor begins to stretch and apply a force on its
adjacent neighbor.
Fig. 1.33

A Wave Transports Energy and Not Matter

When a wave is present in a medium (that is, when there is a disturbance moving through a
medium), the individual particles of the medium are only temporarily displaced from their rest
position. There is always a force acting upon the particles that restores them to their original
position. In a slinky wave, each coil of the slinky ultimately returns to its original position. In a
water wave, each molecule of the water ultimately returns to its original position. And in a
stadium wave, each fan in the bleacher ultimately returns to its original position. It is for this
reason, that a wave is said to involve the movement of a disturbance without the movement of
matter. The particles of the medium (water molecules, slinky coils, stadium fans) simply vibrate
about a fixed position as the pattern of the disturbance moves from one location to another
location.

Waves are said to be an energy transport phenomenon. As a disturbance moves through a


medium from one particle to its adjacent particle, energy is being transported from one end of the
medium to the other. In a slinky wave, a person imparts energy to the first coil by doing work
upon it. The first coil receives a large amount of energy that it subsequently transfers to the
second coil. When the first coil returns to its original position, it possesses the same amount of
energy as it had before it was displaced. The first coil transferred its energy to the second coil.
The second coil then has a large amount of energy that it subsequently transfers to the third coil.
When the second coil returns to its original position, it possesses the same amount of energy as it
had before it was displaced. The third coil has received the energy of the second coil. This
process of energy transfer continues as each coil interacts with its neighbor. In this manner,
energy is transported from one end of the slinky to the other, from its source to another location.

This characteristic of a wave as an energy transport phenomenon distinguishes waves from other
types of phenomenon. Consider a common phenomenon observed at a softball game - the
collision of a bat with a ball. A batter is able to transport energy from her to the softball by
means of a bat. The batter applies a force to the bat, thus imparting energy to the bat in the form

44
of kinetic energy. The bat then carries this energy to the softball and transports the energy to the
softball upon collision. In this example, a bat is used to transport energy from the player to the
softball. However, unlike wave phenomena, this phenomenon involves the transport of matter.
The bat must move from its starting location to the contact location in order to transport energy.
In a wave phenomenon, energy can move from one location to another, yet the particles of matter
in the medium return to their fixed position. A wave transports its energy without transporting
matter.

Waves are seen to move through an ocean or lake; yet the water always returns to its rest
position. Energy is transported through the medium, yet the water molecules are not transported.
Proof of this is the fact that there is still water in the middle of the ocean. The water has not
moved from the middle of the ocean to the shore. If we were to observe a gull or duck at rest on
the water, it would merely bob up-and-down in a somewhat circular fashion as the disturbance
moves through the water. The gull or duck always returns to its original position. The gull or
duck is not transported to the shore because the water on which it rests is not transported to the
shore. In a water wave, energy is transported without the transport of water.

The same thing can be said about a stadium wave. In a stadium wave, the fans do not get out of
their seats and walk around the stadium. We all recognize that it would be silly (and
embarrassing) for any fan to even contemplate such a thought. In a stadium wave, each fan rises
up and returns to the original seat. The disturbance moves through the stadium, yet the fans are
not transported. Waves involve the transport of energy without the transport of matter.

In conclusion, a wave can be described as a disturbance that travels through a medium,


transporting energy from one location (its source) to another location without transporting
matter. Each individual particle of the medium is temporarily displaced and then returns to its
original equilibrium positioned.

Exercise 1.4
1. TRUE or FALSE:

In order for John to hear Jill, air molecules must move from the lips of Jill to the ears of John.

2. Curly and Moe are conducting a wave experiment using a slinky. Curly introduces a
disturbance into the slinky by giving it a quick back and forth jerk. Moe places his cheek (facial)
at the opposite end of the slinky. Using the terminology of this unit, describe what Moe
experiences as the pulse reaches the other end of the slinky.

45
3. Mac and Tosh are experimenting with pulses on a rope. They vibrate an end up and down to
create the pulse and observe it moving from end to end. How does the position of a point on the
rope, before the pulse comes, compare to the position after the pulse has passed?

4. Minute after minute, hour after hour, day after day, ocean waves continue to splash onto the
shore. Explain why the beach is not completely submerged and why the middle of the ocean has
not yet been depleted of its water supply.

5. A medium is able to transport a wave from one location to another because the particles of the
medium are ____.

a. frictionless

b. isolated from one another

c. able to interact

d. very light

1.2.3 Categories of Waves


Waves come in many shapes and forms. While all waves share some basic characteristic
properties and behaviors, some waves can be distinguished from others based on some
observable (and some non-observable) characteristics. It is common to categorize waves based
on these distinguishing characteristics.

Longitudinal versus Transverse Waves versus Surface Waves

One way to categorize waves is on the basis of the direction of movement of the individual
particles of the medium relative to the direction that the waves travel. Categorizing waves on this
basis leads to three notable categories: transverse waves, longitudinal waves, and surface waves.

A transverse wave is a wave in which particles of the medium move in a direction perpendicular
to the direction that the wave moves. Suppose that a slinky is stretched out in a horizontal
direction across the classroom and that a pulse is introduced into the slinky on the left end by
vibrating the first coil up and down. Energy will begin to be transported through the slinky from
left to right. As the energy is transported from left to right, the individual coils of the medium
will be displaced upwards and downwards. In this case, the particles of the medium move
perpendicular to the direction that the pulse moves. This type of wave is a transverse wave.
Transverse waves are always characterized by particle motion being perpendicular to wave
motion.

46
Fig. 1.34

A longitudinal wave is a wave in which particles of the medium move in a direction parallel to
the direction that the wave moves. Suppose that a slinky is stretched out in a horizontal direction
across the classroom and that a pulse is introduced into the slinky on the left end by vibrating the
first coil left and right. Energy will begin to be transported through the slinky from left to right.
As the energy is transported from left to right, the individual coils of the medium will be
displaced leftwards and rightwards. In this case, the particles of the medium move parallel to the
direction that the pulse moves. This type of wave is a longitudinal wave. Longitudinal waves are
always characterized by particle motion being parallel to wave motion.

Fig. 1.35

A sound wave traveling through air is a classic example of a longitudinal wave. As a sound wave
moves from the lips of a speaker to the ear of a listener, particles of air vibrate back and forth in
the same direction and the opposite direction of energy transport. Each individual particle pushes
on its neighboring particle so as to push it forward. The collision of particle #1 with its neighbor
serves to restore particle #1 to its original position and displace particle #2 in a forward direction.
This back and forth motion of particles in the direction of energy transport creates regions within
the medium where the particles are pressed together and other regions where the particles are
spread apart. Longitudinal waves can always be quickly identified by the presence of such
regions. This process continues along the chain of particles until the sound wave reaches the ear
of the listener. A detailed discussion of sound is presented in the next chapter of this Book.

Fig. 1.36

47
Waves traveling through a solid medium can be either transverse waves or longitudinal waves.
Yet waves traveling through the bulk of a fluid (such as a liquid or a gas) are always longitudinal
waves. Transverse waves require a relatively rigid medium in order to transmit their energy. As
one particle begins to move it must be able to exert a pull on its nearest neighbor. If the medium
is not rigid as is the case with fluids, the particles will slide past each other. This sliding action
that is characteristic of liquids and gases prevents one particle from displacing its neighbor in a
direction perpendicular to the energy transport. It is for this reason that only longitudinal waves
are observed moving through the bulk of liquids such as our oceans. Earthquakes are capable of
producing both transverse and longitudinal waves that travel through the solid structures of the
Earth. When seismologists began to study earthquake waves they noticed that only longitudinal
waves were capable of traveling through the core of the Earth. For this reason, geologists believe
that the Earth's core consists of a liquid - most likely molten iron.

While waves that travel within the depths of the ocean are longitudinal waves, the waves that
travel along the surface of the oceans are referred to as surface waves. A surface wave is a wave
in which particles of the medium undergo a circular motion. Surface waves are neither
longitudinal nor transverse. In longitudinal and transverse waves, all the particles in the entire
bulk of the medium move in a parallel and a perpendicular direction (respectively) relative to the
direction of energy transport. In a surface wave, it is only the particles at the surface of the
medium that undergo the circular motion. The motion of particles tends to decrease as one
proceeds further from the surface.

Fig. 1.37

Any wave moving through a medium has a source. Somewhere along the medium, there was an
initial displacement of one of the particles. For a slinky wave, it is usually the first coil that
becomes displaced by the hand of a person. For a sound wave, it is usually the vibration of the
vocal chords or a guitar string that sets the first particle of air in vibrational motion. At the
location where the wave is introduced into the medium, the particles that are displaced from their
equilibrium position always moves in the same direction as the source of the vibration. So if you
wish to create a transverse wave in a slinky, then the first coil of the slinky must be displaced in
a direction perpendicular to the entire slinky. Similarly, if you wish to create a longitudinal wave
in a slinky, then the first coil of the slinky must be displaced in a direction parallel to the entire
slinky.

48
Fig. 1.38

Electromagnetic versus Mechanical Waves

Another way to categorize waves is on the basis of their ability or inability to transmit energy
through a vacuum (i.e., empty space). Categorizing waves on this basis leads to two notable
categories: electromagnetic waves and mechanical waves.

An electromagnetic wave is a wave that is capable of transmitting its energy through a vacuum
(i.e., empty space). Electromagnetic waves are produced by the vibration of charged particles.
Electromagnetic waves that are produced on the sun subsequently travel to Earth through the
vacuum of outer space. Were it not for the ability of electromagnetic waves to travel to through a
vacuum, there would undoubtedly be no life on Earth. All light waves are examples of
electromagnetic waves. Light waves are the topic of another chapter of this Book. While the
basic properties and behaviors of light will be discussed, the detailed nature of an
electromagnetic wave is quite complicated and beyond the scope of this Book.

A mechanical wave is a wave that is not capable of transmitting its energy through a vacuum.
Mechanical waves require a medium in order to transport their energy from one location to
another. A sound wave is an example of a mechanical wave. Sound waves are incapable of
traveling through a vacuum. Slinky waves, water waves, stadium waves, and jump rope waves
are other examples of mechanical waves; each requires some medium in order to exist. A slinky
wave requires the coils of the slinky; a water wave requires water; a stadium wave requires fans
in a stadium; and a jump rope wave requires a jump rope.

The above categories represent just a few of the ways in which physicists categorize waves in
order to compare and contrast their behaviors and characteristic properties. This listing of
categories is not exhaustive; there are other categories as well. The five categories of waves
listed here will be used periodically throughout this unit on waves as well as the units on sound
and light.

Exercise 1.5
49
1. A transverse wave is transporting energy from east to west. The particles of the medium will
move_____.

a. east to west only

b. both eastward and westward

c. north to south only

d. both northward and southward

2.A wave is transporting energy from left to right. The particles of the medium are moving back
and forth in a leftward and rightward direction. This type of wave is known as a ____.

a. mechanical b. electromagnetic
c. transverse d. longitudinal

3. Describe how the fans in a stadium must move in order to produce a longitudinal stadium
wave.

4. A sound wave is a mechanical wave, not an electromagnetic wave. This means that

a. particles of the medium move perpendicular to the direction of energy transport.

b. a sound wave transports its energy through a vacuum.

c. particles of the medium regularly and repeatedly oscillate about their rest position.

d. a medium is required in order for sound waves to transport energy.

5. A science fiction film depicts inhabitants of one spaceship (in outer space) hearing the sound
of a nearby spaceship as it zooms past at high speeds. Critique the physics of this film.

6. If you strike a horizontal rod vertically from above, what can be said about the waves created
in the rod?

a. The particles vibrate horizontally along the direction of the rod.

b. The particles vibrate vertically, perpendicular to the direction of the rod.

c. The particles vibrate in circles, perpendicular to the direction of the rod.

d. The particles travel along the rod from the point of impact to its end.

50
7. Which of the following is not a characteristic of mechanical waves?

a. They consist of disturbances or oscillations of a medium.

b. They transport energy.

c. They travel in a direction that is at right angles to the direction of the particles of the medium.

d. T0hey are created by a vibrating source.

8. The sonar device on a fishing boat uses underwater sound to locate fish. Would you expect
sonar to be a longitudinal or a transverse wave?

1.3 Properties of Waves


The Anatomy of a Wave | Frequency and Period of a Wave
Energy Transport and the Amplitude of a Wave | The Speed of a Wave | The Wave Equation

1.3.1 The Anatomy of a Wave


A transverse wave is a wave in which the particles of the medium are displaced in a direction
perpendicular to the direction of energy transport. A transverse wave can be created in a rope if
the rope is stretched out horizontally and the end is vibrated back-and-forth in a vertical
direction. If a snapshot of such a transverse wave could be taken so as to freeze the shape of the
rope in time, then it would look like the following diagram.

51
Fig. 1.39

The dashed line drawn through the center of the diagram represents the equilibrium or rest
position of the string. This is the position that the string would assume if there were no
disturbance moving through it. Once a disturbance is introduced into the string, the particles of
the string begin to vibrate upwards and downwards. At any given moment in time, a particle on
the medium could be above or below the rest position. Points A, E and H on the diagram
represent the crests of this wave. The crest of a wave is the point on the medium that exhibits the
maximum amount of positive or upward displacement from the rest position. Points C and J on
the diagram represent the troughs of this wave. The trough of a wave is the point on the medium
that exhibits the maximum amount of negative or downward displacement from the rest position.

The wave shown above can be described by a variety of properties. One such property is
amplitude. The amplitude of a wave refers to the maximum amount of displacement of a particle
on the medium from its rest position. In a sense, the amplitude is the distance from rest to crest.
Similarly, the amplitude can be measured from the rest position to the trough position. In the
diagram above, the amplitude could be measured as the distance of a line segment that is
perpendicular to the rest position and extends vertically upward from the rest position to point A.

The wavelength is another property of a wave that is portrayed in the diagram above. The
wavelength of a wave is simply the length of one complete wave cycle. If you were to trace your
finger across the wave in the diagram above, you would notice that your finger repeats its path.
A wave is a repeating pattern. It repeats itself in a periodic and regular fashion over both time
and space. And the length of one such spatial repetition (known as a wave cycle) is the
wavelength. The wavelength can be measured as the distance from crest to crest or from trough
to trough. In fact, the wavelength of a wave can be measured as the distance from a point on a
wave to the corresponding point on the next cycle of the wave. In the diagram above, the
wavelength is the horizontal distance from A to E, or the horizontal distance from B to F, or the
horizontal distance from D to G, or the horizontal distance from E to H. Any one of these
distance measurements would suffice in determining the wavelength of this wave.

A longitudinal wave is a wave in which the particles of the medium are displaced in a direction
parallel to the direction of energy transport. A longitudinal wave can be created in a slinky if the
slinky is stretched out horizontally and the end coil is vibrated back-and-forth in a horizontal
direction. If a snapshot of such a longitudinal wave could be taken so as to freeze the shape of the
slinky in time, then it would look like the following diagram.

52
Fig. 1.40

Because the coils of the slinky are vibrating longitudinally, there are regions where they become
pressed together and other regions where they are spread apart. A region where the coils are
pressed together in a small amount of space is known as a compression. A compression is a
point on a medium through which a longitudinal wave is traveling that has the maximum density.
A region where the coils are spread apart, thus maximizing the distance between coils, is known
as a rarefaction. A rarefaction is a point on a medium through which a longitudinal wave is
traveling that has the minimum density. Points A, C and E on the diagram above represent
compressions and points B, D, and F represent rarefactions. While a transverse wave has an
alternating pattern of crests and troughs, a longitudinal wave has an alternating pattern of
compressions and rarefactions.

As discussed above, the wavelength of a wave is the length of one complete cycle of a wave. For
a transverse wave, the wavelength is determined by measuring from crest to crest. A longitudinal
wave does not have crest; so how can its wavelength be determined? The wavelength can always
be determined by measuring the distance between any two corresponding points on adjacent
waves. In the case of a longitudinal wave, a wavelength measurement is made by measuring the
distance from a compression to the next compression or from a rarefaction to the next
rarefaction. On the diagram above, the distance from point A to point C or from point B to point
D would be representative of the wavelength.

Exercise 1.6
Consider the diagram below in order to answer questions #1-2.

Fig. Q1.6.1

1. The wavelength of the wave in the diagram above is given by letter ______.

2. The amplitude of the wave in the diagram above is given by letter _____.

53
3. Indicate the interval that represents one full wavelength.

Fig. Q1.6.3

a. A to C

b. B to D

c. A to G

d. C to G

1.3.2 Frequency and Period of a Wave


The nature of a wave was discussed in Section 1.1 of this book . In that lesson, it was mentioned
that a wave is created in a slinky by the periodic and repeating vibration of the first coil of the
slinky. This vibration creates a disturbance that moves through the slinky and transports energy
from the first coil to the last coil. A single back-and-forth vibration of the first coil of a slinky
introduces a pulse into the slinky. But the act of continually vibrating the first coil with a back-
and-forth motion in periodic fashion introduces a wave into the slinky.

Suppose that a hand holding the first coil of a slinky is moved back-and-forth two complete
cycles in one second. The rate of the hand's motion would be 2 cycles/second. The first coil,
being attached to the hand, in turn would vibrate at a rate of 2
cycles/second. The second coil, being attached to the first coil, would
vibrate at a rate of 2 cycles/second. The third coil, being attached to
the second coil, would vibrate at a rate of 2 cycles/second. In fact,
every coil of the slinky would vibrate at this rate of 2 cycles/second.
This rate of 2 cycles/second is referred to as the frequency of the
wave. The frequency of a wave refers to how often the particles of the medium vibrate when a
wave passes through the medium. Frequency is a part of our common, everyday language. For
example, it is not uncommon to hear a question like "How frequently do you mow the lawn
during the summer months?" Of course the question is an inquiry about how often the lawn is
mowed and the answer is usually given in the form of "1 time per week." In mathematical terms,
the frequency is the number of complete vibrational cycles of a medium per a given amount of
time. Given this definition, it is reasonable that the quantity frequency would have units of

54
cycles/second, waves/second, vibrations/second, or something/second. Another unit for
frequency is the Hertz (abbreviated Hz) where 1 Hz is equivalent to 1 cycle/second. If a coil of
slinky makes 2 vibrational cycles in one second, then the frequency is 2 Hz. If a coil of slinky
makes 3 vibrational cycles in one second, then the frequency is 3 Hz. And if a coil makes 8
vibrational cycles in 4 seconds, then the frequency is 2 Hz (8 cycles/4 s = 2 cycles/s).

The quantity frequency is often confused with the quantity period. Period refers to the time that it
takes to do something. When an event occurs repeatedly, then we say that the event is periodic
and refer to the time for the event to repeat itself as the period. The period of a wave is the time
for a particle on a medium to make one complete vibrational cycle. Period, being a time, is
measured in units of time such as seconds, hours, days or years. The period of orbit for the Earth
around the Sun is approximately 365 days; it takes 365 days for the Earth to complete a cycle.
The period of a typical class at a high school might be 55 minutes; every 55 minutes a class cycle
begins (50 minutes for class and 5 minutes for passing time means that a class begins every 55
minutes). The period for the minute hand on a clock is 3600 seconds (60 minutes); it takes the
minute hand 3600 seconds to complete one cycle around the clock.

Frequency and period are distinctly different, yet related, quantities. Frequency refers to how
often something happens. Period refers to the time it takes something to happen. Frequency is a
rate quantity. Period is a time quantity. Frequency is the cycles/second. Period is the
seconds/cycle. As an example of the distinction and the relatedness of frequency and period,
consider a woodpecker that drums upon a tree at a periodic rate. If the woodpecker drums upon a
tree 2 times in one second, then the frequency is 2 Hz. Each drum must endure for one-half a
second, so the period is 0.5 s. If the woodpecker drums upon a tree 4 times in one second, then
the frequency is 4 Hz; each drum must endure for one-fourth a second, so the period is 0.25 s. If
the woodpecker drums upon a tree 5 times in one second, then the frequency is 5 Hz; each drum
must endure for one-fifth a second, so the period is 0.2 s. Do you observe the relationship?
Mathematically, the period is the reciprocal of the frequency and vice versa. In equation form,
this is expressed as follows.

Since the symbol f is used for frequency and the symbol T is used for period, these equations are
also expressed as:

The quantity frequency is also confused with the quantity speed. The
speed of an object refers to how fast an object is moving and is usually
expressed as the distance traveled per time of travel. For a wave, the
speed is the distance traveled by a given point on the wave (such as a
crest) in a given period of time. So while wave frequency refers to the
number of cycles occurring per second, wave speed refers to the meters traveled per second. A
wave can vibrate back and forth very frequently, yet have a small speed; and a wave can vibrate

55
back and forth with a low frequency, yet have a high speed. Frequency and speed are distinctly
different quantities. Wave speed will be discussed in more detail later in this Section.

Exercise 1.7
Throughout this unit, internalize the meaning of terms such as period, frequency, and
wavelength. Utilize the meaning of these terms to answer conceptual questions; avoid a formula
fixation.

1. A wave is introduced into a thin wire held tight at each end. It has an amplitude of 3.8 cm, a
frequency of 51.2 Hz and a distance from a crest to the neighboring trough of 12.8 cm.
Determine the period of such a wave.

2. Frieda the fly flaps its wings back and forth 121 times each second. The period of the wing
flapping is ____ sec.

3. A tennis coach paces back and forth along the sideline 10 times in 2 minutes. The frequency of
her pacing is ________ Hz.

a. 5.0 b. 0.20 c. 0.12 d. 0.083

4. Non-digital clocks (which are becoming more rare) have a second hand that rotates around in
a regular and repeating fashion. The frequency of rotation of a second hand on a clock is
_______ Hz.

a. 1/60 b. 1/12 c. ½
d. 1 e. 60

5. Olive Udadi accompanies her father to the park for an afternoon of fun. While there, she hops
on the swing and begins a motion characterized by a complete back-and-forth cycle every 2
seconds. The frequency of swing is _________.

a. 0.5 Hz b. 1 Hz c. 2 Hz

6. In problem #5, the period of swing is __________.

a. 0.5 second b. 1 second c. 2 second

7. A period of 5.0 seconds corresponds to a frequency of ________ Hertz.

a. 0.2 b. 0.5 c. 0.02


d. 0.05 e. 0.002

56
8. A common physics lab involves the study of the oscillations of a pendulum. If a pendulum
makes 33 complete back-and-forth cycles of vibration in 11 seconds, then its period is ______.

9. A child in a swing makes one complete back and forth motion in 3.2 seconds. This statement
provides information about the child's

a. speed

b. frequency

c. period

10. The period of the sound wave produced by a 440 Hertz tuning fork is ___________.

11. As the frequency of a wave increases, the period of the wave ___________.

a. decreases

b. increases

c. remains the same

1.3.3 Energy Transport and the Amplitude of a Wave


As mentioned earlier, a wave is an energy transport phenomenon that transports energy along a
medium without transporting matter. A pulse or a wave is introduced into a slinky when a person
holds the first coil and gives it a back-and-forth motion. This creates a disturbance within the
medium; this disturbance subsequently travels from coil to coil, transporting energy as it moves.
The energy is imparted to the medium by the person as he/she does work upon the first coil to
give it kinetic energy. This energy is transferred from coil to coil until it arrives at the end of the
slinky. If you were holding the opposite end of the slinky, then you would feel the energy as it
reaches your end. In fact, a high energy pulse would likely do some rather noticeable work upon
your hand upon reaching the end of the medium; the last coil of the medium would displace you
hand in the same direction of motion of the coil. For the same reasons, a high energy ocean wave
can do considerable damage to the rocks and piers along the shoreline when it crashes upon it.

The amount of energy carried by a wave is related to the amplitude of the wave. A high energy
wave is characterized by a high amplitude; a low energy wave is characterized by a low
amplitude. As discussed earlier in Section 1.2, the amplitude of a wave refers to the maximum
amount of displacement of a particle on the medium from its rest position. The logic underlying

57
the energy-amplitude relationship is as follows: If a slinky is stretched out in a horizontal
direction and a transverse pulse is introduced into the slinky, the first coil is given an initial
amount of displacement. The displacement is due to the force applied by the person upon the coil
to displace it a given amount from rest. The more energy that the person puts into the pulse, the
more work that he/she will do upon the first coil. The more work that is done upon the first coil,
the more displacement that is given to it. The more displacement that is given to the first coil, the
more amplitude that it will have. So in the end, the amplitude of a transverse pulse is related to
the energy which that pulse transports through the medium. Putting a lot of energy into a
transverse pulse will not affect the wavelength, the frequency or the speed of the pulse. The
energy imparted to a pulse will only affect the amplitude of that pulse.

Fig. 1.41

Consider two identical slinkies into which a pulse is introduced. If the same amount of energy is
introduced into each slinky, then each pulse will have the same amplitude. But what if the
slinkies are different? What if one is made of zinc and the other is made of copper? Will the
amplitudes now be the same or different? If a pulse is introduced into two different slinkies by
imparting the same amount of energy, then the amplitudes of the pulses will not necessarily be
the same. In a situation such as this, the actual amplitude assumed by the pulse is dependent
upon two types of factors: an inertial factor and an elastic factor. Two different materials have
different mass densities. The imparting of energy to the first coil of a slinky is done by the
application of a force to this coil. More massive slinkies have a greater inertia and thus tend to
resist the force; this increased resistance by the greater mass tends to cause a reduction in the
amplitude of the pulse. Different materials also have differing degrees of springiness or
elasticity. A more elastic medium will tend to offer less resistance to the force and allow a
greater amplitude pulse to travel through it; being less rigid (and therefore more elastic), the
same force causes a greater amplitude.

The energy transported by a wave is directly proportional to the square of the amplitude of the
wave. This energy-amplitude relationship is sometimes expressed in the following manner.

Table 1.6

This means that a doubling of the amplitude of a wave is indicative of a


quadrupling of the energy transported by the wave. A tripling of the
amplitude of a wave is indicative of a nine-fold increase in the amount of
energy transported by the wave. And a quadrupling of the amplitude of a
wave is indicative of a 16-fold increase in the amount of energy transported

58
by the wave. The table at the right further expresses this energy-amplitude relationship. Observe
that whenever the amplitude increased by a given factor, the energy value is increased by the
same factor squared. For example, changing the amplitude from 1 unit to 2 units represents a 2-
fold increase in the amplitude and is accompanied by a 4-fold (22 ) increase in the energy; thus 2
units of energy becomes 4 times bigger - 8 units. As another example, changing the amplitude
from 1 unit to 4 units represents a 4-fold increase in the amplitude and is accompanied by a 16-
fold (42 ) increase in the energy; thus 2 units of energy becomes 16 times bigger - 32 units.

Exercise 1.8
1. Mac and Tosh stand 8 meters apart and demonstrate the motion of a transverse wave on a
snakey. The wave e can be described as having a vertical distance of 32 cm from a trough to a
crest, a frequency of 2.4 Hz, and a horizontal distance of 48 cm from a crest to the nearest
trough. Determine the amplitude, period, and wavelength of such a wave.

2. An ocean wave has an amplitude of 2.5 m. Weather conditions suddenly change such that the
wave has an amplitude of 5.0 m. The amount of energy transported by the wave is __________.

a. halved

b. doubled

c. quadrupled

d. remains the same

3. Two waves are traveling through a container of an inert gas. Wave A has an amplitude of .1
cm. Wave B has an amplitude of .2 cm. The energy transported by wave B must be __________
the energy transported by wave A.

a. one-fourth

b. one-half

c. two times larger than

d. four times larger than

59
1.3.4 The Speed of a Wave
A wave is a disturbance that moves along a medium from one end to the other. If one watches an
ocean wave moving along the medium (the ocean water), one can observe that the crest of the
wave is moving from one location to another over a given interval of time. The crest is observed
to cover distance. The speed of an object refers to how fast an object is moving and is usually
expressed as the distance traveled per time of travel. In the case of a wave, the speed is the
distance traveled by a given point on the wave (such as a crest) in a given interval of time. In
equation form,

If the crest of an ocean wave moves a distance of 20 meters in 10 seconds, then the speed of the
ocean wave is 2 m/s. On the other hand, if the crest of an ocean wave moves a distance of 25
meters in 10 seconds (the same amount of time), then the speed of this ocean wave is 2.5 m/s.
The faster wave travels a greater distance in the same amount of time.

Sometimes a wave encounters the end of a medium and the presence of a different medium. For
example, a wave introduced by a person into one end of a slinky will travel through the slinky
and eventually reach the end of the slinky and the presence of the hand of a second person. One
behavior that waves undergo at the end of a medium is reflection. The wave will reflect or
bounce off the person's hand. When a wave undergoes reflection, it remains within the medium
and merely reverses its direction of travel. In the case of a slinky wave, the disturbance can be
seen traveling back to the original end. A slinky wave that travels to the end of a slinky and back
has doubled its distance. That is, by reflecting back to the original location, the wave has
traveled a distance that is equal to twice the length of the slinky.

Reflection phenomena are commonly observed with sound waves.


When you let out a holler within a canyon, you often hear the echo
of the holler. The sound wave travels through the medium (air in this
case), reflects off the canyon wall and returns to its origin (you). The
result is that you hear the echo (the reflected sound wave) of your
holler. A classic physics problem goes like this:

Noah stands 170 meters away from a steep canyon wall. He shouts
and hears the echo of his voice one second later. What is the speed of
the wave? Fig. 1.42

In this instance, the sound wave travels 340 meters in 1 second, so the speed of the wave is 340
m/s. Remember, when there is a reflection, the wave doubles its distance. In other words, the
distance traveled by the sound wave in 1 second is equivalent to the 170 meters down to the
canyon wall plus the 170 meters back from the canyon wall.

60
Variables Affecting Wave Speed

What variables affect the speed at which a wave travels through a medium?
Does the frequency or wavelength of the wave affect its speed? Does the
amplitude of the wave affect its speed? Or are other variables such as the
mass density of the medium or the elasticity of the medium responsible for
affecting the speed of the wave? These questions are often investigated in
the form of a lab in a physics classroom.

Suppose a wave generator is used to produce several waves within a rope of


a measurable tension. The wavelength, frequency and speed are determined. Then the frequency
of vibration of the generator is changed to investigate the affect of frequency upon wave speed.
Finally, the tension of the rope is altered to investigate the affect of tension upon wave speed.
Sample data for the experiment are shown below.

Table 1.7 Speed of a Wave Lab - Sample Data

Tension Frequency Wavelength Speed

Trial (N) (Hz) (m) (m/s)


1 2.0 4.05 4.00 16.2
2 2.0 8.03 2.00 16.1
3 2.0 12.30 1.33 16.4
4 2.0 16.2 1.00 16.2
5 2.0 20.2 0.800 16.2
6 5.0 12.8 2.00 25.6
7 5.0 19.3 1.33 25.7
8 5.0 25.5 1.00 25.5

In the first five trials, the tension of the rope was held constant and the frequency was
systematically changed. The data in rows 1-5 of the table above demonstrate that a change in the
frequency of a wave does not affect the speed of the wave. The speed remained a near constant

61
value of approximately 16.2 m/s. The small variations in the values for the speed were the result
of experimental error, rather than a demonstration of some physical law. The data convincingly
show that wave frequency does not affect wave speed. An increase in wave frequency caused a
decrease in wavelength while the wave speed remained constant.

The last three trials involved the same procedure with a different rope tension. Observe that the
speed of the waves in rows 6-8 is distinctly different than the speed of the wave in rows 1-5. The
obvious cause of this difference is the alteration of the tension of the rope. The speed of the
waves was significantly higher at higher tensions. Waves travel through tighter ropes at higher
speeds. So while the frequency did not affect the speed of the wave, the tension in the medium
(the rope) did. In fact, the speed of a wave is not dependent upon (causally affected by)
properties of the wave itself. Rather, the speed of the wave is dependent upon the properties of
the medium such as the tension of the rope.

One theme of this unit has been that "a wave is a disturbance moving through a medium." There
are two distinct objects in this phrase - the "wave" and the "medium." The medium could be
water, air, or a slinky. These media are distinguished by their properties - the material they are
made of and the physical properties of that material such as the density, the temperature, the
elasticity, etc. Such physical properties describe the material itself, not the wave. On the other
hand, waves are distinguished from each other by their properties - amplitude, wavelength,
frequency, etc. These properties describe the wave, not the material through which the wave is
moving. The lesson of the lab activity described above is that wave speed depends upon the
medium through which the wave is moving. Only an alteration in the properties of the medium
will cause a change in the speed.

Exercise 1.9
1. A teacher attaches a slinky to the wall and begins introducing pulses with different amplitudes.
Which of the two pulses (A or B) below will travel from the hand to the wall in the least amount
of time? Justify your answer.

Fig. Q1.9.1

2. The teacher then begins introducing pulses with a different wavelength. Which of the two
pulses (C or D) will travel from the hand to the wall in the least amount of time ? Justify your
answer.

62
Fig. Q1.9.2

3. The time required for the sound waves (v = 340 m/s) to travel from the tuning fork to
point A is ____ .

Fig. Q1.9.3

a. 0.020 second b. 0.059 second


c. 0.59 second d. 2.9 second

4. Two waves are traveling through the same container of nitrogen gas. Wave A has a
wavelength of 1.5 m. Wave B has a wavelength of 4.5 m. The speed of wave B must be
________ the speed of wave A.

a. one-ninth b. one-third
c. the same as d. three times larger than

5. An automatic focus camera is able to focus on objects by use of an ultrasonic sound wave. The
camera sends out sound waves that reflect off distant objects and return to the camera. A sensor
detects the time it takes for the waves to return and then determines the distance an object is from
the camera. The camera lens then focuses at that distance. Now that's a smart camera! In a
subsequent life, you might have to be a camera; so try this problem for practice:

If a sound wave (speed = 340 m/s) returns to the camera 0.150 seconds after leaving the camera,
then how far away is the object?

6. TRUE or FALSE:

Doubling the frequency of a wave source doubles the speed of the waves.

7. While hiking through a canyon, Noah Formula lets out a scream. An echo (reflection of the
scream off a nearby canyon wall) is heard 0.82 seconds after the scream. The speed of the sound
wave in air is 342 m/s. Calculate the distance from Noah to the nearby canyon wall.

63
8. Mac and Tosh are resting on top of the water near the end of the pool when Mac creates a
surface wave. The wave travels the length of the pool and back in 25 seconds. The pool is 25
meters long. Determine the speed of the wave.

9. The water waves below are traveling along the surface of the ocean at a speed of 2.5 m/s and
splashing periodically against Wilbert's perch. Each adjacent crest is 5 meters apart. The crests
splash Wilbert's feet upon reaching his perch. How much time passes between each successive
drenching? Answer and explain using complete sentences.

Fig. Q1.9.9

1.3.5 The Wave Equation


As was discussed in Section 10.1, a wave is produced when a vibrating source periodically
disturbs the first particle of a medium. This creates a wave pattern that begins to travel along the
medium from particle to particle. The frequency at which each individual particle vibrates is
equal to the frequency at which the source vibrates. Similarly, the period of vibration of each
individual particle in the medium is equal to the period of vibration of the source. In one period,
the source is able to displace the first particle upwards from rest, back to rest, downwards from
rest, and finally back to rest. This complete back-and-forth movement constitutes one complete
wave cycle.

The diagrams at the right show several "snapshots" of the production of a


wave within a rope. The motion of the disturbance along the medium after
every one-fourth of a period is depicted. Observe that in the time it takes
from the first to the last snapshot, the hand has made one complete back-
and-forth motion. A period has elapsed. Observe that during this same
amount of time, the leading edge of the disturbance has moved a distance
equal to one complete wavelength. So in a time of one period, the wave
has moved a distance of one wavelength. Combining this information with

64
the equation for speed (speed = distance/time), it can be said that the speed of a wave is also the
wavelength/period.

Since the period is the reciprocal of the frequency, the expression 1/f can be substituted into the
above equation for period. Rearranging the equation yields a new equation of the form:

Speed = Wavelength • Frequency Fig. 1.43

The above equation is known as the wave equation. It states the mathematical relationship
between the speed (v) of a wave and its wavelength ( ) and frequency (f). Using the symbols v,
, and f, the equation can be rewritten as

v=f•
As a test of your understanding of the wave equation and its mathematical use in analyzing
wave motion, consider the following three-part question:

QUIZ 1.2

1. Stan and Anna are conducting a slinky experiment. They are studying the possible
affect of several variables upon the speed of a wave in a slinky. Their data table is
shown below. Fill in the blanks in the table, analyze the data, and answer the
following questions.

Table Q.1.7: Table for Quiz 1.2

Medium Wavelength Frequency Speed


Zinc,
1.75 m 2.0 Hz ______
1-in. dia. Coils
Zinc,
0.90 m 3.9 Hz ______
1-in. dia. coils
Copper, 1.19 m 2.1 Hz ______

65
1-in. dia. coils
Copper,
0.60 m 4.2 Hz ______
1-in. dia. coils
Zinc,
0.95 m 2.2 Hz ______
3-in. dia. coils
Zinc,
1.82 m 1.2 Hz ______
3-in. dia. coils

2. As the wavelength of a wave in a uniform medium increases, its speed will _____.

a. decrease b. increase c. remain the same

3. As the wavelength of a wave in a uniform medium increases, its frequency will ____

a. decrease b. increase c. remain the same

4. The speed of a wave depends upon (i.e., is causally affected by) ...

a. the properties of the medium through which the wave travels

b. the wavelength of the wave.

c. the frequency of the wave.

d. both the wavelength and the frequency of the wave.

The above example illustrates how to use the wave equation to solve mathematical problems. It
also illustrates the principle that wave speed is dependent upon medium properties and
independent of wave properties. Even though the wave speed is calculated by multiplying
wavelength by frequency, an alteration in wavelength does not affect wave speed. Rather, an
alteration in wavelength affects the frequency in an inverse manner. A doubling of the
wavelength results in a halving of the frequency; yet the wave speed is not changed.

66
Exercise 1.10
1. Two waves on identical strings have frequencies in a ratio of 2 to 1. If their wave speeds are
the same, then how do their wavelengths compare?

a. 2:1 b. 1:2 c. 4:1 d. 1:4

2. Mac and Tosh stand 8 meters apart and demonstrate the motion of a transverse wave on a
snakey. The wave e can be described as having a vertical distance of 32 cm from a trough to a
crest, a frequency of 2.4 Hz, and a horizontal distance of 48 cm from a crest to the nearest
trough. Determine the amplitude, period, and wavelength and speed of such a wave.

3. Dawn and Aram have stretched a slinky between them and begin experimenting with waves.
As the frequency of the waves is doubled,

a. the wavelength is halved and the speed remains constant

b. the wavelength remains constant and the speed is doubled

c. both the wavelength and the speed are halved.

d. both the wavelength and the speed remain constant.

4. A ruby-throated hummingbird beats its wings at a rate of about 70 wing beats per second.

a. What is the frequency in Hertz of the sound wave?

b. Assuming the sound wave moves with a velocity of 350 m/s, what is the wavelength of the
wave?

5. Ocean waves are observed to travel along the water surface during a developing storm. A
Coast Guard weather station observes that there is a vertical distance from high point to low
point of 4.6 meters and a horizontal distance of 8.6 meters between adjacent crests. The waves
splash into the station once every 6.2 seconds. Determine the frequency and the speed of these
waves.

67
6. Two boats are anchored 4 meters apart. They bob up
and down, returning to the same up position every 3
seconds. When one is up the other is down. There are
never any wave crests between the boats. Calculate the
speed of the waves.
Fig. Q1.10.6

1.4 Behavior of Waves


Boundary Behavior | Reflection, Refraction, and Diffraction | Interference of Waves
The Doppler Effect

1.4.1 Boundary Behavior


As a wave travels through a medium, it will often reach the end of the medium and encounter an
obstacle or perhaps another medium through which it could travel. One example of this has
already been mentioned in Lesson 2. A sound wave is known to reflect off canyon walls and
other obstacles to produce an echo. A sound wave traveling through air within a canyon reflects
off the canyon wall and returns to its original source. What affect does reflection have upon a
wave? Does reflection of a wave affect the speed of the wave? Does reflection of a wave affect
the wavelength and frequency of the wave? Does reflection of a wave affect the amplitude of the
wave? Or does reflection affect other properties and characteristics of a wave's motion? The
behavior of a wave (or pulse) upon reaching the end of a medium is referred to as boundary
behavior. When one medium ends, another medium begins; the interface of the two media is
referred to as the boundary and the behavior of a wave at that boundary is described as its
boundary behavior. The questions that are listed above are the types of questions we seek to
answer when we investigate the boundary behavior of waves.

Fixed End Reflection

First consider an elastic rope stretched from end to end. One end will be securely attached to a
pole on a lab bench while the other end will be held in the
hand in order to introduce pulses into the medium. Because
the right end of the rope is attached to a pole (which is
attached to a lab bench) (which is attached to the floor that is
attached to the building that is attached to the Earth), the last
particle of the rope will be unable to move when a
Fig. 1.44
68
disturbance reaches it. This end of the rope is referred to as a fixed end.

If a pulse is introduced at the left end of the rope, it will travel through the rope towards the right
end of the medium. This pulse is called the incident pulse since it is incident towards (i.e.,
approaching) the boundary with the pole. When the incident pulse reaches the boundary, two
things occur:

• A portion of the energy carried by the pulse is reflected and returns towards the left end
of the rope. The disturbance that returns to the left after bouncing off the pole is known as
the reflected pulse.
• A portion of the energy carried by the pulse is transmitted to the pole, causing the pole
to vibrate.

Because the vibrations of the pole are not visibly obvious, the energy transmitted to it is not
typically discussed. The focus of the discussion will be on the reflected pulse. What
characteristics and properties could describe its motion?

When one observes the reflected pulse off the fixed end, there are several notable observations.
First the reflected pulse is inverted. That is, if an upward displaced pulse is incident towards a
fixed end boundary, it will reflect and return as a downward displaced pulse. Similarly, if a
downward displaced pulse is incident towards a fixed end boundary, it will reflect and return as
an upward displaced pulse.

Fig. 1.45

The inversion of the reflected pulse can be explained by returning to our conceptions of the
nature of a mechanical wave. When a crest reaches the end of a medium ("medium A"), the last
particle of the medium A receives an upward displacement. This particle is attached to the first
particle of the other medium ("medium B") on the other side of the boundary. As the last particle
of medium A pulls upwards on the first particle of medium B, the first particle of medium B
pulls downwards on the last particle of medium A. This is merely Newton's third law of action-
reaction. For every action, there is an equal and opposite reaction. The upward pull on the first
particle of medium B has little affect upon this particle due to the large mass of the pole and the
lab bench to which it is attached. The affect of the downward pull on the last particle of medium
A (a pull that is in turn transmitted to the other particles) results in causing the upward
displacement to become a downward displacement. The upward displaced incident pulse thus
returns as a downward displaced reflected pulse. It is important to note that it is the heaviness of
the pole and the lab bench relative to the rope that causes the rope to become inverted upon
interacting with the wall. When two media interact by exerting pushes and pulls upon each other,

69
the most massive medium wins the interaction. Just like in arm wrestling, the medium that loses
receives a change in its state of motion.

Other notable characteristics of the reflected pulse include:

• The speed of the reflected pulse is the same as the speed of the incident pulse.
• The wavelength of the reflected pulse is the same as the wavelength of the incident pulse.
• The amplitude of the reflected pulse is less than the amplitude of the incident pulse.

Of course, it is not surprising that the speed of the incident and reflected pulse are identical since
the two pulses are traveling in the same medium. Since the speed of a wave (or pulse) is
dependent upon the medium through which it travels, two pulses in the same medium will have
the same speed. A similar line of reasoning explains why the incident and reflected pulses have
the same wavelength. Every particle within the rope will have the same frequency. Being
connected to one another, they must vibrate at the same frequency. Since the wavelength of a
wave depends upon the frequency and the speed, two waves having the same frequency and the
same speed must also have the same wavelength. Finally, the amplitude of the reflected pulse is
less than the amplitude of the incident pulse since some of the energy of the pulse was
transmitted into the pole at the boundary. The reflected pulse is carrying less energy away from
the boundary compared to the energy that the incident pulse carried towards the boundary. Since
the amplitude of a pulse is indicative of the energy carried by the pulse, the reflected pulse has a
smaller amplitude than the incident pulse.

Flickr Physics Photo


This sequence photography photo shows an upward displaced pulse traveling from the left end of
a wave machine towards the right end. The right end is held tightly; it is a fixed end. The wave
reflects off this fixed end and returns as a downward displaced pulse. Reflection off a fixed end
results in inversion.

Fig. 1.46 (cortesy: www.flickr.com)

70
Free End Reflection+

Now consider what would happen if the end of the rope


were free to move. Instead of being securely attached to a
lab pole, suppose it is attached to a ring that is loosely fit
around the pole. Because the right end of the rope is no
longer secured to the pole, the last particle of the rope
will be able to move when a disturbance reaches it. This
end of the rope is referred to as a free end.
Fig. 1.47
Once more if a pulse is introduced at the left end of the rope, it will travel through the rope
towards the right end of the medium. When the incident pulse reaches the end of the medium, the
last particle of the rope can no longer interact with the first particle of the pole. Since the rope
and pole are no longer attached and interconnected, they will slide past each other. So when a
crest reaches the end of the rope, the last particle of the rope receives the same upward
displacement; only now there is no adjoining particle to pull downward upon the last particle of
the rope to cause it to be inverted. The result is that the reflected pulse is not inverted. When an
upward displaced pulse is incident upon a free end, it returns as an upward displaced pulse after
reflection. And when a downward displaced pulse is incident upon a free end, it returns as a
downward displaced pulse after reflection. Inversion is not observed in free end reflection.

Fig. 1.48

Watch It!
A pulse is introduced into the left end of a wave machine. The incident pulse is displaced
upward. When it reaches the right end, it reflects back. The reflected pulse is not inverted. It is
also displaced upward.

The above discussion of free end and fixed end reflection focuses upon the reflected pulse. As
was mentioned, the transmitted portion of the pulse is difficult to observe when it is transmitted
into a pole. But what if the original medium were attached to another rope with different
properties? How could the reflected pulse and transmitted pulse be described in situations in
which an incident pulse reflects off and transmits into a second medium?

Transmission of a Pulse Across a Boundary from Less to More Dense

71
Let's consider a thin rope attached to a thick rope, with each rope held at opposite ends by
people. And suppose that a pulse is introduced by the person holding the end of the thin rope. If
this is the case, there will be an incident pulse traveling in the less dense medium (the thin rope)
towards the boundary with a more dense medium (the thick rope).

Fig. 1.49

Upon reaching the boundary, the usual two behaviors will occur.

• A portion of the energy carried by the incident pulse is reflected and returns towards the
left end of the thin rope. The disturbance that returns to the left after bouncing off the
boundary is known as the reflected pulse.
• A portion of the energy carried by the incident pulse is transmitted into the thick rope.
The disturbance that continues moving to the right is known as the transmitted pulse.

The reflected pulse will be found to be inverted in situations such as this. During the interaction
between the two media at the boundary, the first particle of the more dense medium overpowers
the smaller mass of the last particle of the less dense medium. This causes an upward displaced
pulse to become a downward displaced pulse. The more dense medium on the other hand was at
rest prior to the interaction. The first particle of this medium receives an upward pull when the
incident pulse reaches the boundary. Since the more dense medium was originally at rest, an
upward pull can do nothing but cause an upward displacement. For this reason, the transmitted
pulse is not inverted. In fact, transmitted pulses can never be inverted. Since the particles in this
medium are originally at rest, any change in their state of motion would be in the same direction
as the displacement of the particles of the incident pulse.

The Before and After snapshots of the two media are shown in the diagram below.

Fig. 1.50

72
Comparisons can also be made between the characteristics of the transmitted pulse and those of
the reflected pulse. Once more there are several noteworthy characteristics.

• The transmitted pulse (in the more dense medium) is traveling slower than the reflected
pulse (in the less dense medium).
• The transmitted pulse (in the more dense medium) has a smaller wavelength than the
reflected pulse (in the less dense medium).
• The speed and the wavelength of the reflected pulse are the same as the speed and the
wavelength of the incident pulse.

One goal of physics is to use physical models and ideas to explain the observations made of the
physical world. So how can these three characteristics be explained? First recall from Ssection
10.2 that the speed of a wave is dependent upon the properties of the medium. In this case, the
transmitted and reflected pulses are traveling in two distinctly different media. Waves always
travel fastest in the least dense medium. Thus, the reflected pulse will be traveling faster than the
transmitted pulse. Second, particles in the more dense medium will be vibrating with the same
frequency as particles in the less dense medium. Since the transmitted pulse was introduced into
the more dense medium by the vibrations of particles in the less dense medium, they must be
vibrating at the same frequency. So the reflected and transmitted pulses have the different speeds
but the same frequency. Since the wavelength of a wave depends upon the frequency and the
speed, the wave with the greatest speed must also have the greatest wavelength. Finally, the
incident and the reflected pulse share the same medium. Since the two pulses are in the same
medium, they will have the same speed. Since the reflected pulse was created by the vibrations
of the incident pulse, they will have the same frequency. And two waves with the same speed
and the same frequency must also have the same wavelength.

Flickr Physics Photo


A wave machine is used to demonstrate the behavior of a wave at a boundary.
TOP: An incident pulse is introduced into the right end of the wave machine. It travels through
the less dense medium until it reaches the boundary with a more dense medium.
MIDDLE: At the boundary, both reflection and transmission occur.
BOTTOM: The reflected pulse is inverted and of about the same length (though a smaller
amplitude) as the incident pulse. The transmitted pulse is shorter and slower than the incident
and transmitted pulse.

73
Fig. 1.51 (courtesy: www.flickr.com)

Transmission of a Pulse Across a Boundary from More to Less Dense

Finally, let's consider a thick rope attached to a thin rope, with the incident pulse originating in
the thick rope. If this is the case, there will be an incident pulse traveling in the more dense
medium (thick rope) towards the boundary with a less dense medium (thin rope). Once again
there will be partial reflection and partial transmission at the boundary. The reflected pulse in
this situation will not be inverted. Similarly, the transmitted pulse is not inverted (as is always
the case). Since the incident pulse is in a heavier medium, when it reaches the boundary, the first
particle of the less dense medium does not have sufficient mass to overpower the last particle of
the more dense medium. The result is that an upward displaced pulse incident towards the
boundary will reflect as an upward displaced pulse. For the same reasons, a downward displaced
pulse incident towards the boundary will reflect as a downward displaced pulse.

The Before and After snapshots of the two media are shown in the diagram below.

74
Fig. 1.52

Comparisons between the characteristics of the transmitted pulse and the reflected pulse lead to
the following observations.

• The transmitted pulse (in the less dense medium) is traveling faster than the reflected
pulse (in the more dense medium).
• The transmitted pulse (in the less dense medium) has a larger wavelength than the
reflected pulse (in the more dense medium).
• The speed and the wavelength of the reflected pulse are the same as the speed and the
wavelength of the incident pulse.

These three observations are explained using the same logic as used above.

Flickr Physics Photo


A wave machine is used to demonstrate the behavior of a wave at a boundary.
TOP: An incident pulse is introduced into the left end of the wave machine. It travels through the
more dense medium until it reaches the boundary with a less dense medium.
MIDDLE: At the boundary, both reflection and transmission occur.
BOTTOM: The reflected pulse is NOT inverted and of about the same length (though a smaller
amplitude) as the incident pulse. The transmitted pulse is longer and faster than the incident and
transmitted pulse.

75
Fig. 1.53(courtesy: www.flickr.com)

The boundary behavior of waves in ropes can be summarized by the following principles:

• The wave speed is always greatest in the least dense rope.


• The wavelength is always greatest in the least dense rope.
• The frequency of a wave is not altered by crossing a boundary.
• The reflected pulse becomes inverted when a wave in a less dense rope is heading
towards a boundary with a more dense rope.
• The amplitude of the incident pulse is always greater than the amplitude of the reflected
pulse.

All the observations discussed here can be explained by the simple application of these
principles. Take a few moments to use these principles to answer the following questions.

76
Exercise 1.11
Case 1: A pulse in a more dense medium is traveling towards the boundary with a less dense
medium.

Fig. Q1.11.1

1. The reflected pulse in medium 1 ________ (will, will not) be inverted because _______.

2. The speed of the transmitted pulse will be ___________ (greater than, less than, the same as)
the speed of the incident pulse.

3. The speed of the reflected pulse will be ______________ (greater than, less than, the same as)
the speed of the incident pulse.

4. The wavelength of the transmitted pulse will be ___________ (greater than, less than, the
same as) the wavelength of the incident pulse.

5. The frequency of the transmitted pulse will be ___________ (greater than, less than, the same
as) the frequency of the incident pulse.

Case 2: A pulse in a less dense medium is traveling towards the boundary with a more dense
medium.

Fig. Q1.11.5

6. The reflected pulse in medium 1 ________ (will, will not) be inverted because
_____________.

7. The speed of the transmitted pulse will be ___________ (greater than, less than, the same as)
the speed of the incident pulse.

8. The speed of the reflected pulse will be ______________ (greater than, less than, the same as)
the speed of the incident pulse.

77
9. The wavelength of the transmitted pulse will be ___________ (greater than, less than, the
same as) the wavelength of the incident pulse.

10. The frequency of the transmitted pulse will be ___________ (greater than, less than, the
same as) the frequency of the incident pulse.

1.4.2 Reflection, Refraction, and Diffraction


Previously in Section 1.3, the behavior of waves traveling along a rope from a more dense
medium to a less dense medium (and vice versa) was discussed. The wave doesn't just stop when
it reaches the end of the medium. Rather, a wave will undergo certain behaviors when it
encounters the end of the medium. Specifically, there will be some reflection off the boundary
and some transmission into the new medium. But what if the wave is traveling in a two-
dimensional medium such as a water wave traveling through ocean water? Or what if the wave is
traveling in a three-dimensional medium such as a sound wave or a light wave traveling through
air? What types of behaviors can be expected of such two- and three-dimensional waves?

The study of waves in two dimensions is often done using a ripple


tank. A ripple tank is a large glass-bottomed tank of water that is
used to study the behavior of water waves. A light typically shines
upon the water from above and illuminates a white sheet of paper
placed directly below the tank. A portion of light is absorbed by the
water as it passes through the tank. A crest of water will absorb
more light than a trough. So the bright spots represent wave troughs
and the dark spots represent wave crests. As the water waves move
through the ripple tank, the dark and bright spots move as well. As
the waves encounter obstacles in their path, their behavior can be
observed by watching the movement of the dark and bright spots on
the sheet of paper. Ripple tank demonstrations are commonly done
in a Physics class in order to discuss the principles underlying the Fig. 1.54
reflection, refraction, and diffraction of waves.

If a linear object attached to an oscillator bobs back and forth within the water, it becomes a
source of straight waves. These straight waves have alternating crests and troughs. As viewed on
the sheet of paper below the tank, the crests are the dark lines stretching across
the paper and the troughs are the bright lines. These waves will travel through
the water until they encounter an obstacle - such as the wall of the tank or an
object placed within the water. The diagram at the right depicts a series of
straight waves approaching a long barrier extending at an angle across the tank
of water. The direction that these wave-fronts (straight-line crests) are traveling

78 Fig. 1.55
through the water is represented by the blue arrow. The blue arrow is called a ray and is drawn
perpendicular to the wave-fronts. Upon reaching the barrier placed within the water, these waves
bounce off the water and head in a different direction. The diagram below shows the reflected
wave-fronts and the reflected ray. Regardless of the angle at which the wave-fronts approach the
barrier, one general law of reflection holds true: the waves will always reflect in such a way that
the angle at which they approach the barrier equals the angle at which they reflect off the barrier.
This is known as the law of reflection. This law will be discussed in more detail in Book 6.

Fig. 1.56

The discussion above pertains to the reflection of waves off of straight


surfaces. But what if the surface is curved, perhaps in the shape of a parabola?
What generalizations can be made for the reflection of water waves off
parabolic surfaces? Suppose that a rubber tube having the shape of a parabola
is placed within the water. The diagram at the right depicts such a parabolic
barrier in the ripple tank. Several wavefronts are approaching the barrier; the
ray is drawn for these wavefronts. Upon reflection off the parabolic barrier, Fig. 1.57
the water waves will change direction and head towards a point. This is depicted in the diagram
below. It is as though all the energy being carried by the water waves is converged at a single
point - the point is known as the focal point. After passing through the focal point, the waves
spread out through the water. Reflection of waves off of curved surfaces will be discussed in
more detail in Book 6.

Fig. 1.58

Reflection involves a change in direction of waves when they bounce off a barrier. Refraction of
waves involves a change in the direction of waves as they pass from one medium to another.
Refraction, or the bending of the path of the waves, is accompanied by a change in speed and
wavelength of the waves. In Section 1.2, it was mentioned that the speed of a wave is dependent

79
upon the properties of the medium through which the waves travel. So if the medium (and its
properties) is changed, the speed of the waves is changed. The most significant property of water
that would affect the speed of waves traveling on its surface is the depth of the water. Water
waves travel fastest when the medium is the deepest. Thus, if water waves are passing from deep
water into shallow water, they will slow down. And as mentioned in the previous sub-section of
Section 1.3, this decrease in speed will also be accompanied by a decrease in wavelength. So as
water waves are transmitted from deep water into shallow water, the speed
decreases, the wavelength decreases, and the direction changes.

This boundary behavior of water waves can be observed in a ripple tank if the
tank is partitioned into a deep and a shallow section. If a pane of glass is
placed in the bottom of the tank, one part of the tank will be deep and the other
part of the tank will be shallow. Waves traveling from the deep end to the Fig. 1.59
shallow end can be seen to refract (i.e., bend), decrease wavelength (the wave-fronts get closer
together), and slow down (they take a longer time to travel the same distance). When traveling
from deep water to shallow water, the waves are seen to bend in such a manner that they seem to
be traveling more perpendicular to the surface. If traveling from shallow water to deep water, the
waves bend in the opposite direction. The refraction of light waves will be discussed in more
detail in Book 6.

Reflection involves a change in direction of waves when they bounce off a barrier; refraction of
waves involves a change in the direction of waves as they pass from one medium to another; and
diffraction involves a change in direction of waves as they pass through an
opening or around a barrier in their path. Water waves have the ability to
travel around corners, around obstacles and through openings. This ability is
most obvious for water waves with longer wavelengths. Diffraction can be
demonstrated by placing small barriers and obstacles in a ripple tank and
observing the path of the water waves as they encounter the obstacles. The
waves are seen to pass around the barrier into the regions behind it;
subsequently the water behind the barrier is disturbed. The amount of
diffraction (the sharpness of the bending) increases with increasing
wavelength and decreases with decreasing wavelength. In fact, when the
wavelength of the waves is smaller than the obstacle, no noticeable diffraction occurs. Fig. 1.60

Diffraction of water waves is observed in a harbor as waves bend around small boats and are
found to disturb the water behind them. The same waves however are unable to diffract around
larger boats since their wavelength is smaller than the boat. Diffraction of sound waves is
commonly observed; we notice sound diffracting around corners, allowing us to hear others who
are speaking to us from adjacent rooms. Many forest-dwelling birds take advantage of the
diffractive ability of long-wavelength sound waves. Owls for instance are able to communicate
across long distances due to the fact that their long-wavelength hoots are able to diffract around
forest trees and carry farther than the short-wavelength tweets of songbirds. Diffraction is
observed of light waves but only when the waves encounter obstacles with extremely small

80
wavelengths (such as particles suspended in our atmosphere). Diffraction of sound waves and of
light waves will be discussed in a later Section.

Reflection, refraction and diffraction are all boundary behaviors of waves associated with the
bending of the path of a wave. The bending of the path is an observable behavior when the
medium is a two- or three-dimensional medium. Reflection occurs when there is a bouncing off
of a barrier. Reflection of waves off straight barriers follows the law of reflection. Reflection of
waves off parabolic barriers results in the convergence of the waves at a focal point. Refraction
is the change in direction of waves that occurs when waves travel from one medium to another.
Refraction is always accompanied by a wavelength and speed change. Diffraction is the bending
of waves around obstacles and openings. The amount of diffraction increases with increasing
wavelength.

1.4.3 Interference of Waves


What happens when two waves meet while they travel through the same medium? What effect
will the meeting of the waves have upon the appearance of the medium? Will the two waves
bounce off each other upon meeting (much like two billiard balls would) or will the two waves
pass through each other? These questions involving the meeting of two or more waves along the
same medium pertain to the topic of wave interference.

Wave interference is the phenomenon that occurs when two waves meet while traveling along
the same medium. The interference of waves causes the medium to take on a shape that results
from the net effect of the two individual waves upon the particles of the medium. To begin our
exploration of wave interference, consider two pulses of the same amplitude traveling in
different directions along the same medium. Let's suppose that each displaced upward 1 unit at
its crest and has the shape of a sine wave. As the sine pulses move towards each other, there will
eventually be a moment in time when they are completely overlapped. At that moment, the
resulting shape of the medium would be an upward displaced sine pulse with an amplitude of 2
units. The diagrams below depict the before and during interference snapshots of the medium for
two such pulses. The individual sine pulses are drawn in red and blue and the resulting
displacement of the medium is drawn in green.

Fig. 1.61
This type of interference is sometimes called constructive interference. Constructive
interference is a type of interference that occurs at any location along the medium where the two

81
interfering waves have a displacement in the same direction. In this case, both waves have an
upward displacement; consequently, the medium has an upward displacement that is greater than
the displacement of the two interfering pulses. Constructive interference is observed at any
location where the two interfering waves are displaced upward. But it is also observed when both
interfering waves are displaced downward. This is shown in the diagram below for two
downward displaced pulses.

Fig. 1.62

In this case, a sine pulse with a maximum displacement of -1 unit (negative means a downward
displacement) interferes with a sine pulse with a maximum displacement of -1 unit. These two
pulses are drawn in red and blue. The resulting shape of the medium is a sine pulse with a
maximum displacement of -2 units.

Destructive interference is a type of interference that occurs at any location along the medium
where the two interfering waves have a displacement in the opposite direction. For instance,
when a sine pulse with a maximum displacement of +1 unit meets a sine pulse with a maximum
displacement of -1 unit, destructive interference occurs. This is depicted in the diagram below.

Fig. 1.63

In the diagram above, the interfering pulses have the same maximum displacement but in
opposite directions. The result is that the two pulses completely destroy each other when they are
completely overlapped. At the instant of complete overlap, there is no resulting displacement of
the particles of the medium. This "destruction" is not a permanent condition. In fact, to say that
the two waves destroy each other can be partially misleading. When it is said that the two pulses
destroy each other, what is meant is that when overlapped, the effect of one of the pulses on the
displacement of a given particle of the medium is destroyed or canceled by the effect of the other
pulse. Recall from Section 1.2 that waves transport energy through a medium by means of each
individual particle pulling upon its nearest neighbor. When two pulses with opposite
displacements (i.e., one pulse displaced up and the other down) meet at a given location, the
upward pull of one pulse is balanced (canceled or destroyed) by the downward pull of the other
pulse. Once the two pulses pass through each other, there is still an upward displaced pulse and a
downward displaced pulse heading in the same direction that they were heading before the
interference. Destructive interference leads to only a momentary condition in which the
medium's displacement is less than the displacement of the largest-amplitude wave.

The two interfering waves do not need to have equal amplitudes in opposite directions for
destructive interference to occur. For example, a pulse with a maximum displacement of +1 unit

82
could meet a pulse with a maximum displacement of -2 units. The resulting displacement of the
medium during complete overlap is -1 unit.

Fig. 1.64
This is still destructive interference since the two interfering pulses have opposite displacements.
In this case, the destructive nature of the interference does not lead to complete cancellation.

Interestingly, the meeting of two waves along a medium does not alter the individual waves or
even deviate them from their path. This only becomes an astounding behavior when it is
compared to what happens when two billiard balls meet or two football players meet. Billiard
balls might crash and bounce off each other and football players might crash and come to a stop.
Yet two waves will meet, produce a net resulting shape of the medium, and then continue on
doing what they were doing before the interference.

Fig. 1.65

The task of determining the shape of the resultant demands that the principle of superposition is
applied. The principle of superposition is sometimes stated as follows:

When two waves interfere, the resulting displacement of the medium at any location is
the algebraic sum of the displacements of the individual waves at that same location.

In the cases above, the summing of the individual displacements for locations of complete
overlap was made out to be an easy task - as easy as simple arithmetic:

Displacement of Pulse 1 Displacement of Pulse 2 = Resulting Displacement


+1 +1 = +2
-1 -1 = -2
+1 -1 = 0
+1 -2 = -1

In actuality, the task of determining the complete shape of the entire medium during interference
demands that the principle of superposition be applied for every point (or nearly every point)
83
along the medium. As an example of the complexity of this
task, consider the two interfering waves at the right. A
snapshot of the shape of each individual wave at a particular
instant in time is shown. To determine the precise shape of
the medium at this given instant in time, the principle of
superposition must be applied to several locations along the
medium. A short cut involves measuring the displacement
from equilibrium at a few strategic locations. Thus,
approximately 20 locations have been picked and labeled as
A, B, C, D, etc. The actual displacement of each individual
wave can be counted by measuring from the equilibrium
position up to the particular wave. At position A, there is no
displacement for either individual wave; thus, the resulting
displacement of the medium at position will be 0 units. At position B, the smaller wave has a
displacement of approximately 1.4 units (indicated by the red dot); the larger wave has a
displacement of approximately 2 units (indicated by the blue dot). Thus, the resulting
displacement of the medium will be approximately 3.4 units. At position C, the smaller wave has
a displacement of approximately 2 units; the larger wave has a displacement of approximately 4
units; thus, the resulting displacement of the medium will be approximately 6 units. At position
D, the smaller wave has a displacement of approximately 1.4 units; the larger wave has a
displacement of approximately 2 units; thus, the resulting displacement of the medium will be
approximately 3.4 units. This process can be repeated for every position. When finished, a dot
(done in green below) can be marked on the graph to note the displacement of the medium at
each given location. The actual shape of the medium can then be sketched by estimating the
position between the various marked points and sketching the wave. This is shown as the green
line in the diagram below.

Fig. 1.67

84
Exercise 1.12

1. Several positions along the medium are labeled with a


letter. Categorize each labeled position along the medium as
being a position where either constructive or destructive
interference occurs.

Fig.Q 1.12.1

2. Twin water bugs Jimminy and Johnny are both


creating a series of circular waves by jiggling their legs in
the water. The waves undergo interference and create the
pattern represented in the diagram at the right. The thick
lines in the diagram represent wave crests and the thin
lines represent wave troughs. Several of positions in the
water are labeled with a letter. Categorize each labeled
position as being a position where either constructive or
destructive interference occurs. Fig.Q 1.12.2

1.4.4 The Doppler Effect


Suppose that there is a happy bug in the center of a circular water puddle.
The bug is periodically shaking its legs in order to produce disturbances
that travel through the water. If these disturbances originate at a point,
then they would travel outward from that point in all directions. Since
each disturbance is traveling in the same medium, they would all travel in
every direction at the same speed. The pattern produced by the bug's
shaking would be a series of concentric circles as shown in the diagram at
the right. These circles would reach the edges of the water puddle at the
same frequency. An observer at point A (the left edge of the puddle)
would observe the disturbances to strike the puddle's edge at the same frequency that would be
observed by an observer at point B (at the right edge of the puddle). In fact,Fig.
the 1.68
frequency at
which disturbances reach the edge of the puddle would be the same as the frequency at which the

85
bug produces the disturbances. If the bug produces disturbances at a frequency of 2 per second,
then each observer would observe them approaching at a frequency of 2 per second.

Now suppose that our bug is moving to the right across the puddle of
water and producing disturbances at the same frequency of 2
disturbances per second. Since the bug is moving towards the right, each
consecutive disturbance originates from a position that is closer to Fig. 1.69
observer B and farther from observer A. Subsequently, each consecutive
disturbance has a shorter distance to travel before reaching observer B
and thus takes less time to reach observer B. Thus, observer B observes
that the frequency of arrival of the disturbances is higher than the
frequency at which disturbances are produced. On the other hand, each
consecutive disturbance has a further distance to travel before reaching observer A. For this
reason, observer A observes a frequency of arrival that is less than the frequency at which the
disturbances are produced. The net effect of the motion of the bug (the source of waves) is that
the observer towards whom the bug is moving observes a frequency that is higher than 2
disturbances/second; and the observer away from whom the bug is moving observes a frequency
that is less than 2 disturbances/second. This effect is known as the Doppler effect.

The Doppler effect is observed whenever the source of waves is moving with respect to an
observer. The Doppler effect can be described as the effect produced by a moving source of
waves in which there is an apparent upward shift in frequency for observers towards whom the
source is approaching and an apparent downward shift in frequency for observers from whom the
source is receding. It is important to note that the effect does not result because of an actual
change in the frequency of the source. Using the example above, the bug is still producing
disturbances at a rate of 2 disturbances per second; it just appears to the observer whom the bug
is approaching that the disturbances are being produced at a frequency greater than 2
disturbances/second. The effect is only observed because the distance between observer B and
the bug is decreasing and the distance between observer A and the bug is increasing.

The Doppler effect can be observed for any type of wave - water wave, sound wave, light wave,
etc. We are most familiar with the Doppler effect because of our experiences with sound waves.
Perhaps you recall an instance in which a police car or emergency vehicle was traveling towards
you on the highway. As the car approached with its siren blasting, the pitch of the siren sound (a
measure of the siren's frequency) was high; and then suddenly after the car passed by, the pitch
of the siren sound was low. That was the Doppler effect - an apparent shift in frequency for a
sound wave produced by a moving source.

86
Fig. 1.70

The Doppler effect is of intense interest to astronomers who use the information about the shift
in frequency of electromagnetic waves produced by moving stars in our galaxy and beyond in
order to derive information about those stars and galaxies. The belief that the universe is
expanding is based in part upon observations of electromagnetic waves emitted by stars in distant
galaxies. Furthermore, specific information about stars within galaxies can be determined by
application of the Doppler effect. Galaxies are clusters of stars that typically rotate about some
center of mass point. Electromagnetic radiation emitted by such stars in a distant galaxy would
appear to be shifted downward in frequency (a red shift) if the star is rotating in its cluster in a
direction that is away from the Earth. On the other hand, there is an upward shift in frequency (a
blue shift) of such observed radiation if the star is rotating in a direction that is towards the Earth.

1.5 Standing Waves


Traveling Waves vs. Standing Waves | Formation of Standing Waves
Nodes and Anti-nodes | Harmonics and Patterns | Mathematics of Standing Waves

1.5.1 Traveling Waves vs. Standing Waves


A mechanical wave is a disturbance that is created by a vibrating object and subsequently travels
through a medium from one location to another, transporting energy as it moves. The mechanism
by which a mechanical wave propagates itself through a medium involves
particle interaction; one particle applies a push or pull on its adjacent
neighbor, causing a displacement of that neighbor from the equilibrium or
rest position. As a wave is observed traveling through a medium, a crest is
seen moving along from particle to particle. This crest is followed by a

87
Fig. 1.71
trough that is in turn followed by the next crest. In fact, one would observe a distinct wave
pattern (in the form of a sine wave) traveling through the medium. This sine wave pattern
continues to move in uninterrupted fashion until it encounters another wave along the medium or
until it encounters a boundary with another medium. This type of wave pattern that is seen
traveling through a medium is sometimes referred to as a traveling wave.

Traveling waves are observed when a wave is not confined to a given space along the medium.
The most commonly observed traveling wave is an ocean wave. If a wave is introduced into an
elastic cord with its ends held 3 meters apart, it becomes confined in a small region. Such a wave
has only 3 meters along which to travel. The wave will quickly reach the end of the cord, reflect
and travel back in the opposite direction. Any reflected portion of the wave will then interfere
with the portion of the wave incident towards the fixed end. This interference produces a new
shape in the medium that seldom resembles the shape of a sine wave. Subsequently, a traveling
wave (a repeating pattern that is observed to move through a medium in uninterrupted fashion) is
not observed in the cord. Indeed there are traveling waves in the cord; it is just that they are not
easily detectable because of their interference with each other. In such instances, rather than
observing the pure shape of a sine wave pattern, a rather irregular and non-repeating pattern is
produced in the cord that tends to change appearance over time. This irregular looking shape is
the result of the interference of an incident sine wave pattern with a reflected sine wave pattern
in a rather non-sequenced and untimely manner. Both the incident and reflected wave patterns
continue their motion through the medium, meeting up with one another at different locations in
different ways. For example, the middle of the cord might experience a crest meeting a half
crest; then moments later, a crest meeting a quarter trough; then moments later, a three-quarters
crest meeting a one-fifth trough, etc. This interference leads to a very irregular and non-repeating
motion of the medium. The appearance of an actual wave pattern is difficult to detect amidst the
irregular motions of the individual particles.

It is however possible to have a wave confined to a given space in a medium and still produce a
regular wave pattern that is readily discernible amidst the motion of the medium. For instance, if
an elastic rope is held end-to-end and vibrated at just the right frequency, a wave pattern would
be produced that assumes the shape of a sine wave and is seen to change over time. The wave
pattern is only produced when one end of the rope is vibrated at just the right frequency. When
the proper frequency is used, the interference of the incident wave and the reflected wave occur
in such a manner that there are specific points along the medium that appear to be standing still.
Because the observed wave pattern is characterized by points that appear to be standing still, the
pattern is often called a standing wave pattern. There are other points along the medium whose
displacement changes over time, but in a regular manner. These points vibrate back and forth
from a positive displacement to a negative displacement; the vibrations occur at regular time
intervals such that the motion of the medium is regular and repeating. A pattern is readily
observable.

The diagram at the right depicts a standing wave pattern in a medium.


A snapshot of the medium over time is depicted using various colors.
Note that point A on the medium moves from a maximum positive to a
maximum negative displacement over time. The diagram only shows
one-half cycle of the motion of the standing wave pattern. The motion
Fig. 1.72
88
would continue and persist, with point A returning to the same maximum positive displacement
and then continuing its back-and-forth vibration between the up to the down position. Note that
point B on the medium is a point that never moves. Point B is a point of no displacement. Such
points are known as nodes and will be discussed in more detail later in this Section. The standing
wave pattern that is shown at the right is just one of many different patterns that could be
produced within the rope. Other patterns will be discussed later in the Section.

1.5.2 Formation of Standing Waves


A standing wave pattern is a vibrational pattern created within a medium when the vibrational
frequency of the source causes reflected waves from one end of the medium to interfere with
incident waves from the source. This interference occurs in such a manner that specific points
along the medium appear to be standing still. Because the observed wave pattern is characterized
by points that appear to be standing still, the pattern is often called a standing wave pattern. Such
patterns are only created within the medium at specific frequencies of vibration. These
frequencies are known as harmonic frequencies, or merely harmonics. At any frequency other
than a harmonic frequency, the interference of reflected and incident waves leads to a resulting
disturbance of the medium that is irregular and non-repeating.

But how are standing wave formations formed? And why are they only
formed when the medium is vibrated at specific frequencies? And what
makes these so-called harmonic frequencies so special and magical? To
answer these questions, let's consider a snakey stretched across the room,
approximately 4-meters from end to end. (A "snakey" is a slinky-like device
that consists of a large concentration of small-diameter metal coils.) If an
upward displaced pulse is introduced at the left end of the snakey, it will
travel rightward across the snakey until it reaches the fixed end on the right
side of the snakey. Upon reaching the fixed end, the single pulse will reflect
and undergo inversion. That is, the upward displaced pulse will become a
downward displaced pulse. Now suppose that a second upward displaced
pulse is introduced into the snakey at the precise moment that the first crest Fig. 1.73
undergoes its fixed end reflection. If this is done with perfect timing, a rightward moving,
upward displaced pulse will meet up with a leftward moving, downward displaced pulse in the
exact middle of the snakey. As the two pulses pass through each other, they will undergo
destructive interference. Thus, a point of no displacement in the exact middle of the snakey will
be produced. The animation below shows several snapshots of the meeting of the two pulses at
various stages in their interference. The individual pulses are drawn in blue and red; the resulting
shape of the medium (as found by the principle of superposition) is shown in green. Note that
there is a point on the diagram in the exact middle of the medium that never experiences any
displacement from the equilibrium position.

89
Fig. 1.74

An upward displaced pulse introduced at one end will destructively interfere in the exact middle
of the snakey with a second upward displaced pulse introduced from the same end if the
introduction of the second pulse is performed with perfect timing. The same rationale could be
applied to two downward displaced pulses introduced from the same end. If the second pulse is
introduced at precisely the moment that the first pulse is reflecting from the fixed end, then
destructive interference will occur in the exact middle of the snakey.

The above discussion only explains why two pulses might


interfere destructively to produce a point of no displacement in
the middle of the snakey. A wave is certainly different than a
pulse. What if there are two waves traveling in the medium?
Understanding why two waves introduced into a medium with
perfect timing might produce a point of displacement in the
middle of the medium is a mere extension of the above
discussion. While a pulse is a single disturbance that moves Fig. 1.75
through a medium, a wave is a repeating pattern of crests and troughs. Thus, a wave can be
thought of as an upward displaced pulse (crest) followed by a downward displaced pulse (trough)
followed by an upward displaced pulse (crest) followed by a downward displaced pulse (trough)
followed by... . Since the introduction of a crest is followed by the introduction of a trough, every
crest and trough will destructively interfere in such a way that the middle of the medium is a
point of no displacement.

Of course, this all demands that the timing is perfect. In the above discussion, perfect timing was
achieved if every wave crest was introduced into the snakey at the precise time that the previous
wave crest began its reflection at the fixed end. In this situation, there will be one complete
wavelength within the snakey moving to the right at every instant in time; this incident wave will
meet up with one complete wavelength moving to the left at every instant in time. Under these
conditions, destructive interference always occurs in the middle of the snakey. Either a full crest
meets a full trough or a half-crest meets a half-trough or a quarter-crest meets a quarter-trough
at this point. The animation below represents several snapshots of two waves traveling in
opposite directions along the same medium. The waves are interfering in such a manner that
there are points of no displacement produced at the same positions along the medium. These
points along the medium are known as nodes and are labeled with an N. There are also points
along the medium that vibrate back and forth between points of large positive displacement and
points of large negative displacement. These points are known as antinodes and are labeled with
an AN. The two individual waves are drawn in blue and green and the resulting shape of the
medium is drawn in black.

90
Fig. 1.76
There are other ways to achieve this perfect timing. The main idea behind the timing is to
introduce a crest at the instant that another crest is either at the halfway point across the medium
or at the end of the medium. Regardless of the number of crests and troughs that are in between,
if a crest is introduced at the instant another crest is undergoing its fixed end reflection, a node
(point of no displacement) will be formed in the middle of the medium. The number of other
nodes that will be present along the medium is dependent upon the number of crests that were
present in between the two timed crests. If a crest is introduced at the instant another crest is at
the halfway point across the medium, then an antinode (point of maximum displacement) will be
formed in the middle of the medium by means of constructive interference. In such an instance,
there might also be nodes and antinodes located elsewhere along the medium.

A standing wave pattern is an interference phenomenon. It is formed as the result of the perfectly
timed interference of two waves passing through the same medium. A standing wave pattern is
not actually a wave; rather it is the pattern resulting from the presence of two waves (sometimes
more) of the same frequency with different directions of travel within the same medium. The
physics of musical instruments has a basis in the conceptual and mathematical aspects of
standing waves. For this reason, the topic will be revisited in the Sound and Music chapter.

1.5.3 Nodes and Anti-nodes


As mentioned earlier in Section1.5, a standing wave pattern is an interference phenomenon. It is
formed as the result of the perfectly timed interference of two waves passing through the same
medium. A standing wave pattern is not actually a wave; rather it is the pattern resulting from the
presence of two waves of the same frequency with different directions of travel within the same
medium.

One characteristic of every standing wave pattern is that there are


points along the medium that appear to be standing still. These
points, sometimes described as points of no displacement, are
referred to as nodes. There are other points along the medium that
undergo vibrations between a large positive and large negative

91
displacement. These are the points that undergo the maximum displacement during each
vibrational cycle of the standing wave. In a sense, these points are the opposite of nodes, and so
they are called antinodes. A standing wave pattern always consists of an alternating pattern of
nodes and antinodes. The animation shown below depicts a rope vibrating with a standing wave
pattern. The nodes and antinodes are labeled on the diagram. When a standing wave pattern is
established in a medium, the nodes and the antinodes are always located at the same position
along the medium; they are standing still. It is this characteristic that has earned the pattern the
name standing wave.

Fig. 1.77

Flickr Physics Photo


A standing wave is established upon a vibrating string using a harmonic oscillator and a
frequency generator. A strobe is used to illuminate the string several times during each cycle.
The finger is pointing at a nodal position.

Fig. 1.78

92
The positioning of the nodes and antinodes in a standing wave pattern can be explained by
focusing on the interference of the two waves. The nodes are produced at locations where
destructive interference occurs. For instance, nodes form at locations where a crest of one wave
meets a trough of a second wave; or a half-crest of one wave meets a half-trough of a second
wave; or a quarter-crest of one wave meets a quarter-trough of a second wave; etc. Antinodes,
on the other hand, are produced at locations where constructive interference occurs. For instance,
if a crest of one wave meets a crest of a second wave, a point of large positive displacement
results. Similarly, if a trough of one wave meets a trough of a second wave, a point of large
negative displacement results. Antinodes are always vibrating back and forth between these
points of large positive and large negative displacement; this is because during a complete cycle
of vibration, a crest will meet a crest; and then one-half cycle later, a trough will meet a trough.
Because antinodes are vibrating back and forth between a large positive and large negative
displacement, a diagram of a standing wave is sometimes depicted by drawing the shape of the
medium at an instant in time and at an instant one-half vibrational cycle later. This is done in the
diagram below.

Fig. 1.79

Nodes and antinodes should not be confused with crests and troughs. When the motion of a
traveling wave is discussed, it is customary to refer to a point of large maximum displacement as
a crest and a point of large negative displacement as a trough. These represent points of the
disturbance that travel from one location to another through the medium. An antinode on the
other hand is a point on the medium that is staying in the same location. Furthermore, an
antinode vibrates back and forth between a large upward and a large downward displacement.
And finally, nodes and antinodes are not actually part of a wave. Recall that a standing wave is
not actually a wave but rather a pattern that results from the interference of two or more waves.
Since a standing wave is not technically a wave, an antinode is not technically a point on a wave.
The nodes and antinodes are merely unique points on the medium that make up the wave pattern.

Watch It!

A physics instructor demonstrates and explains the formation of a longitudinal standing wave in
a spring.

93
Exercise 1.13
1. Suppose that there was a ride at an amusement park that was titled The Standing Wave. Which
location - node or antinode - on the ride would give the greatest thrill?

2. A standing wave is formed when ____.

a. a wave refracts due to changes in the properties of the medium.

b. a wave reflects off a canyon wall and is heard shortly after it is formed.

c. red, orange, and yellow wavelengths bend around suspended atmospheric particles.

d. two identical waves moving different directions along the same medium interfere.

3. The number of nodes in the standing wave shown in the


diagram at the right is ____.

a. 6 b. 7
c. 8 d. 14 Fig.Q 1.13.3

4. The number of antinodes in the standing wave shown in the diagram above right is ____.

a. 6 b. 7 c. 8 d. 14

Consider the standing wave pattern at the right in answering these next two questions.

5. The number of nodes in the entire pattern is ___.

a. 7 b. 8
c. 9 d. 16

Fig.Q. 1.13.5
6. Of all the labeled points, destructive interference occurs at point(s) ____.

94
a. B, C, and D b. A, E, and F c. A only
d. C only e. all points

1.5.4 Harmonics and Patterns


As mentioned earlier in Section 1.5, standing wave patterns are wave patterns produced in a
medium when two waves of identical frequencies interfere in such a manner to produce points
along the medium that always appear to be standing still. These points that have the appearance
of standing still are referred to as nodes. Standing waves are often demonstrated in a Physics
class using a snakey that is vibrated by the teacher at one end and held fixed at the other end by a
student. The waves reflect off the fixed end and interfere with the waves introduced by the
teacher to produce this regular and repeating pattern known as a standing wave pattern. A variety
of actual wave patterns could be produced, with each pattern characterized by a distinctly
different number of nodes. Such standing wave patterns can only be produced within the medium
when it is vibrated at certain frequencies. There are several frequencies with which the snakey
can be vibrated to produce the patterns. Each frequency is associated with a different standing
wave pattern. These frequencies and their associated wave patterns are referred to as harmonics.

As discussed earlier in Section 1.5, the production of standing wave patterns demand that the
introduction of crests and troughs into the medium be precisely timed. If the timing is not
precise, then a regular and repeating wave pattern will not be discerned within the medium - a
harmonic does not exist at such a frequency. With precise timing, reflected vibrations from the
opposite end of the medium will interfere with vibrations introduced into the medium in such a
manner that there are points that always appear to be standing still. These points of no
displacement are referred to as nodes. Positioned in between every node is a point that undergoes
maximum displacement from a positive position to a negative position. These points of
maximum displacement are referred to as antinodes.

The simplest standing wave pattern that could be produced within a snakey is one that has points
of no displacement (nodes) at the two ends of the snakey and one point of maximum
displacement (antinode) in the middle. The animation below depicts the vibrational pattern
observed when the medium is seen vibrating in this manner.

First Harmonic Standing Wave Pattern

95
Fig. 1.80
The above standing wave pattern is known as the first harmonic. It is the simplest wave pattern
produced within the snakey and is obtained when the teacher introduced vibrations into the end
of the medium at low frequencies.

Other wave patterns can be observed within the snakey when it is vibrated at greater frequencies.
For instance, if the teacher vibrates the end with twice the frequency as that associated with the
first harmonic, then a second standing wave pattern can be achieved. This standing wave pattern
is characterized by nodes on the two ends of the snakey and an additional node in the exact
center of the snakey. As in all standing wave patterns, every node is separated by an antinode.
This pattern with three nodes and two antinodes is referred to as the second harmonic and is
depicted in the animation shown below.

Second Harmonic Standing Wave Pattern

Fig. 1.81
If the frequency at which the teacher vibrates the snakey is increased even more, then the third
harmonic wave pattern can be produced within the snakey. The standing wave pattern for the
third harmonic has an additional node and antinode between the ends of the snakey. The pattern
is depicted in the animation shown below.

Third Harmonic Standing Wave Pattern

96
Fig. 1.82
Observe that each consecutive harmonic is characterized by having one additional node and
antinode compared to the previous one. The table below summarizes the features of the standing
wave patterns for the first several harmonics.

Table 1.8
Harmonic # of Nodes # of Antinodes Pattern

1st 2 1

2nd 3 2

3rd 4 3

4th 5 4

5th 6 5

6th 7 6

Nth n+1 N --

As one studies harmonics and their standing wave patterns, it becomes evident that there is a
predictability about them. Not surprisingly, this predictability expresses itself in a series of
mathematical relationships that relate the wavelength of the wave pattern to the length of the
medium. Additionally, the frequency of each harmonic is mathematically related to the
frequency of the first harmonic. The next sub-section of Section 1.5 will explore these
mathematical relationships.

Flickr Physics Photo

97
A home-made wave machine was made using string, PVC pipe and connections, a battery, two
motors and some wire. The wave machine does a great job producing the second and third
harmonic standing wave patterns. The third harmonic is shown here. Observe the two nodes and
the three antinodes positioned between the ends of the string.

Fig. 1.83

1.5.5 Mathematics of Standing Waves


As discussed in Section 1.5, standing wave patterns are wave patterns produced in a medium
when two waves of identical frequencies interfere in such a manner to produce points along the
medium that always appear to be standing still. Such standing wave patterns are produced within
the medium when it is vibrated at certain frequencies. Each frequency is associated with a
different standing wave pattern. These frequencies and their associated wave patterns are
referred to as harmonics. A careful study of the standing wave patterns reveal a clear
mathematical relationship between the wavelength of the wave that produces the pattern and the
length of the medium in which the pattern is displayed. Furthermore, there is a predictability
about this mathematical relationship that allows one to generalize and deduce a statement
concerning this relationship. To illustrate, consider the first harmonic standing wave pattern for a

vibrating rope as shown below.


Fig. 1.84

The pattern for the first harmonic reveals a single antinode in the middle of the rope. This
antinode position along the rope vibrates up and down from a maximum upward displacement
from rest to a maximum downward displacement as shown. The vibration of the rope in this
manner creates the appearance of a loop within the string. A complete wave in a pattern could be
described as starting at the rest position, rising upward to a peak displacement, returning back

98
down to a rest position, then descending to a peak downward displacement and finally returning
back to the rest position. The animation below depicts this familiar pattern. As shown in the
animation, one complete wave in a standing wave pattern consists of two loops. Thus, one loop
is equivalent to one-half of a wavelength.

Fig. 1.85
In comparing the standing wave pattern for the first harmonic with its single loop to the diagram
of a complete wave, it is evident that there is only one-half of a wave stretching across the length
of the string. That is, the length of the string is equal to one-half the length of a wave. Put in the
form of an equation:

Now consider the string being vibrated with a frequency that establishes the standing wave
pattern for the second harmonic.

Fig. 1.86
The second harmonic pattern consists of two anti-nodes. Thus, there are two loops within the
length of the string. Since each loop is equivalent to one-half a wavelength, the length of the
string is equal to two-halves of a wavelength. Put in the form of an equation:

The same reasoning pattern can be applied to the case of the string being vibrated with a
frequency that establishes the standing wave pattern for the third harmonic.

Fig. 1.87
The third harmonic pattern consists of three anti-nodes. Thus, there are three loops within the
length of the string. Since each loop is equivalent to one-half a wavelength, the length of the
string is equal to three-halves of a wavelength. Put in the form of an equation:

When inspecting the standing wave patterns and the length-wavelength relationships for the first
three harmonics, a clear pattern emerges. The number of antinodes in the pattern is equal to the

99
harmonic number of that pattern. The first harmonic has one antinode; the second harmonic has
two antinodes; and the third harmonic has three antinodes. Thus, it can be generalized that the
nth harmonic has n antinodes where n is an integer representing the harmonic number.
Furthermore, one notices that there are n halves wavelengths present within the length of the
string. Put in the form of an equation:

This information is summarized in the table below.

Table 1.9
Length-Wavelength
Harmonic Pattern # of Loops
Relationship

1st 1 L=1/2•

2nd 2 L=2/2•

3rd 3 L=3/2•

4th 4 L=4/2•

5th 5 L=5/2•

6th 6 L=6/2•

-- n L=n/2•
nth

For standing wave patterns, there is a clear mathematical relationship between the length of a
string and the wavelength of the wave that creates the pattern. The mathematical relationship
simply emerges from the inspection of the pattern and the understanding that each loop in the
pattern is equivalent to one-half of a wavelength. The general equation that describes this length-
wavelength relationship for any harmonic is:

100
Exercise 1.14
1. Suppose that a string is 1.2 meters long and vibrates in the first, second and third harmonic
standing wave patterns. Determine the wavelength of the waves for each of the three patterns.

2. The string at the right is 1.5 meters long and is vibrating as the first
harmonic. The string vibrates up and down with 33 complete
vibrational cycles in 10 seconds. Determine the frequency, period,
wavelength and speed for this wave.
Fig.Q. 1.14.2

3. The string at the right is 6.0 meters long and is vibrating as the third harmonic.
The string vibrates up and down with 45 complete vibrational cycles in 10 seconds.
Determine the frequency, period, wavelength and speed for this wave.

Fig.Q. 1.14.3

4. The string at the right is 5.0 meters long and is vibrating as the fourth
harmonic. The string vibrates up and down with 48 complete vibrational cycles
in 20 seconds. Determine the frequency, period, wavelength and speed for this
wave.
Fig.Q. 1.14.4

5. The string at the right is 8.2 meters long and is vibrating as the fifth
harmonic. The string vibrates up and down with 21 complete vibrational cycles
in 5 seconds. Determine the frequency, period, wavelength and speed for this
wave.

Fig.Q. 1.14.5

101
CHAPTER TWO
Sound Waves and Music

2.1 The Nature of a Sound Wave


Sound is a Mechanical Wave | Sound is a Longitudinal Wave | Sound is a Pressure Wave

2.1.1 Sound is a Mechanical Wave


Sound and music are parts of our everyday sensory experience. Just as humans have eyes for the
detection of light and color, so we are equipped with ears for the detection of sound. We seldom
take the time to ponder the characteristics and behaviors of sound and the mechanisms by which
sounds are produced, propagated, and detected. The basis for an understanding of sound, music
and hearing is the physics of waves. Sound is a wave that is created by vibrating objects and
propagated through a medium from one location to another. In this unit, we will investigate the
nature, properties and behaviors of sound waves and apply basic wave principles towards an
understanding of music.

As discussed in Chapter 1, a wave can be described as a disturbance that travels through a


medium, transporting energy from one location to another location. The medium is simply the
material through which the disturbance is moving; it can be thought of as a series of interacting
particles. The example of a slinky wave is often used to illustrate the nature of a wave. A
disturbance is typically created within the slinky by the back and forth movement of the first coil
of the slinky. The first coil becomes disturbed and begins to push or pull on the second coil. This
push or pull on the second coil will displace the second coil from its equilibrium position. As the
second coil becomes displaced, it begins to push or pull on the third coil; the push or pull on the
third coil displaces it from its equilibrium position. As the third coil becomes displaced, it begins
to push or pull on the fourth coil. This process continues in consecutive fashion, with each
individual particle acting to displace the adjacent particle. Subsequently the disturbance travels
through the slinky. As the disturbance moves from coil to coil, the energy that was originally
introduced into the first coil is transported along the medium from one location to another.

A sound wave is similar in nature to a slinky wave for a variety of reasons. First, there is a
medium that carries the disturbance from one location to another. Typically, this medium is air,
though it could be any material such as water or steel. The medium is simply a series of
interconnected and interacting particles. Second, there is an original source of the wave, some
vibrating object capable of disturbing the first particle of the medium. The disturbance could be

102
created by the vibrating vocal cords of a person, the vibrating string and soundboard of a guitar
or violin, the vibrating tines of a tuning fork, or the vibrating diaphragm of a radio speaker.
Third, the sound wave is transported from one location to another by means of particle-to-
particle interaction. If the sound wave is moving through air, then as one air particle is displaced
from its equilibrium position, it exerts a push or pull on its nearest neighbors, causing them to be
displaced from their equilibrium position. This particle interaction continues throughout the
entire medium, with each particle interacting and causing a disturbance of its nearest neighbors.
Since a sound wave is a disturbance that is transported through a medium via the mechanism of
particle-to-particle interaction, a sound wave is characterized as a mechanical wave.

The creation and propagation of sound waves are often demonstrated in class through the use of
a tuning fork. A tuning fork is a metal object consisting of two tines capable of vibrating if struck
by a rubber hammer or mallet. As the tines of the tuning forks vibrate back and forth, they begin
to disturb surrounding air molecules. These disturbances are passed on to adjacent air molecules
by the mechanism of particle interaction. The motion of the disturbance, originating at the tines
of the tuning fork and traveling through the medium (in this case, air) is what is referred to as a
sound wave. The generation and propagation of a sound wave is demonstrated in the animation
below.

Fig. 2.1

Many Physics demonstration tuning forks are mounted on a sound


box. In such instances, the vibrating tuning fork, being connected
to the sound box, sets the sound box into vibrational motion. In
turn, the sound box, being connected to the air inside of it, sets the
air inside of the sound box into vibrational motion. As the tines of
the tuning fork, the structure of the sound box, and the air inside
of the sound box begin vibrating at the same frequency, a louder
sound is produced. In fact, the more particles that can be made to
vibrate, the louder or more amplified the sound. This concept is Fig. 2.2
often demonstrated by the placement of a vibrating tuning fork against the glass panel of an
overhead projector or on the wooden door of a cabinet. The vibrating tuning fork sets the glass
panel or wood door into vibrational motion and results in an amplified sound.

We know that a tuning fork is vibrating because we hear the sound that is produced by its
vibration. Nonetheless, we do not actually visibly detect any vibrations
of the tines. This is because the tines are vibrating at a very high
frequency. If the tuning fork that is being used corresponds to middle C
on the piano keyboard, then the tines are vibrating at a frequency of
256 Hertz; that is, 256 vibrations per second. We are unable to visibly
detect vibrations of such high frequency. A common physics
demonstration involves slowing down the vibrations by through the use

103
of a strobe light. If the strobe light puts out a flash of light at a frequency of 512 Hz (two times
the frequency of the tuning fork), then the tuning fork can be observed to be moving in a back
and forth motion. With the room darkened, the strobe would allow us to view the position of the
tines two times during their vibrational cycle. Thus we would see the tines when they are
displaced far to the left and again when they are displaced far to the right. This would be
convincing proof that the tines of the tuning fork are indeed vibrating to produce sound.

In a previous chapter, a distinction was made between two categories of waves: mechanical
waves and electromagnetic waves. Electromagnetic waves are
waves that have an electric and magnetic nature and are capable
of traveling through a vacuum. Electromagnetic waves do not
require a medium in order to transport their energy. Mechanical
waves are waves that require a medium in order to transport
their energy from one location to another. Because mechanical
waves rely on particle interaction in order to transport their
energy, they cannot travel through regions of space that are void
of particles. That is, mechanical waves cannot travel through a
vacuum. This feature of mechanical waves is often
demonstrated in a Physics class. A ringing bell is placed in a jar and air inside the jar is
evacuated. Once air is removed from the jar, the sound of the ringing bell can no longer be heard.
The clapper is seen striking the bell; but the sound that it produces cannot be heard because there
are no particles inside of the jar to transport the disturbance through the vacuum. Sound is a
mechanical wave and cannot travel through a vacuum.

Exercise 2.1
1. A sound wave is different than a light wave in that a sound wave is

a. produced by an oscillating object and a light wave is not.

b. not capable of traveling through a vacuum.

c. not capable of diffracting and a light wave is.

d. capable of existing with a variety of frequencies and a light wave has a single frequency.

104
2.1.2 Sound as a Longitudinal Wave
In the first part of Section 11.1, it was mentioned that sound is a mechanical wave that is created
by a vibrating object. The vibrations of the object set particles in the surrounding medium in
vibrational motion, thus transporting energy through the medium. For a sound wave traveling
through air, the vibrations of the particles are best described as longitudinal. Longitudinal waves
are waves in which the motion of the individual particles of the medium is in a direction that is
parallel to the direction of energy transport. A longitudinal wave can be created in a slinky if the
slinky is stretched out in a horizontal direction and the first coils of the slinky are vibrated
horizontally. In such a case, each individual coil of the medium is set into vibrational motion in
directions parallel to the direction that the energy is transported.

Fig. 2.3

Sound waves in air (and any fluid medium) are longitudinal waves because particles of the
medium through which the sound is transported vibrate parallel to the direction that the sound
wave moves. A vibrating string can create longitudinal waves as depicted in the animation
below. As the vibrating string moves in the forward direction, it begins to push upon surrounding
air molecules, moving them to the right towards their nearest neighbor. This causes the air
molecules to the right of the string to be compressed into a small region of space. As the
vibrating string moves in the reverse direction (leftward), it lowers the pressure of the air
immediately to its right, thus causing air molecules to move back leftward. The lower pressure to
the right of the string causes air molecules in that region immediately to the right of the string to
expand into a large region of space. The back and forth vibration of the string causes individual
air molecules (or a layer of air molecules) in the region immediately to the right of the string to
continually vibrate back and forth horizontally. The molecules move rightward as the string
moves rightward and then leftward as the string moves leftward. These back and forth vibrations
are imparted to adjacent neighbors by particle-to-particle interaction. Other surrounding particles
begin to move rightward and leftward, thus sending a wave to the right. Since air molecules (the
particles of the medium) are moving in a direction that is parallel to the direction that the wave
moves, the sound wave is referred to as a longitudinal wave. The result of such longitudinal
vibrations is the creation of compressions and rarefactions within the air.

105
Fig. 2.4

Regardless of the source of the sound wave - whether it is a vibrating string or the vibrating tines
of a tuning fork - sound waves traveling through air are longitudinal waves. And the essential
characteristic of a longitudinal wave that distinguishes it from other types of waves is that the
particles of the medium move in a direction parallel to the direction of energy transport.

2.1.3 Sound is a Pressure Wave


Sound is a mechanical wave that results from the back and forth vibration of the particles of the
medium through which the sound wave is moving. If a sound wave is moving from left to right
through air, then particles of air will be displaced both rightward and leftward as the energy of
the sound wave passes through it. The motion of the particles is parallel (and anti-parallel) to the
direction of the energy transport. This is what characterizes sound waves in air as longitudinal
waves.

A vibrating tuning fork is capable of creating such a longitudinal wave. As the tines of the fork
vibrate back and forth, they push on neighboring air particles. The forward motion of a tine
pushes air molecules horizontally to the right and the backward retraction of the tine creates a
low-pressure area allowing the air particles to move back to the left.

Fig. 2.5

Because of the longitudinal motion of the air particles, there are regions in the air where the air
particles are compressed together and other regions where the air particles are spread apart.
These regions are known as compressions and rarefactions respectively. The compressions are
regions of high air pressure while the rarefactions are regions of low air pressure. The diagram
below depicts a sound wave created by a tuning fork and propagated through the air in an open
tube. The compressions and rarefactions are labeled.
106
Fig. 2.6

The wavelength of a wave is merely the distance that a disturbance travels along the medium in
one complete wave cycle. Since a wave repeats its pattern once every wave cycle, the
wavelength is sometimes referred to as the length of the repeating patterns - the length of one
complete wave. For a transverse wave, this length is commonly measured from one wave crest to
the next adjacent wave crest or from one wave trough to the next adjacent wave trough. Since a
longitudinal wave does not contain crests and troughs, its wavelength must be measured
differently. A longitudinal wave consists of a repeating pattern of compressions and rarefactions.
Thus, the wavelength is commonly measured as the distance from one compression to the next
adjacent compression or the distance from one rarefaction to the next adjacent rarefaction.

Since a sound wave consists of a repeating pattern of high-pressure and low-pressure regions
moving through a medium, it is sometimes referred to as a pressure wave. If a detector, whether
it is the human ear or a man-made instrument, were used to detect a sound wave, it would detect
fluctuations in pressure as the sound wave impinges upon the detecting device. At one instant in
time, the detector would detect a high pressure; this would correspond to the arrival of a
compression at the detector site. At the next instant in time, the detector might detect normal
pressure. And then finally a low pressure would be detected, corresponding to the arrival of a
rarefaction at the detector site. The fluctuations in pressure as detected by the detector occur at
periodic and regular time intervals. In fact, a plot of pressure versus time would appear as a sine
curve. The peak points of the sine curve correspond to compressions; the low points correspond
to rarefactions; and the "zero points" correspond to the pressure that the air would have if there
were no disturbance moving through it. The diagram below depicts the correspondence between
the longitudinal nature of a sound wave in air and the pressure-time fluctuations that it creates at
a fixed detector location.

Fig. 2.7

The above diagram can be somewhat misleading if you are not careful. The representation of
sound by a sine wave is merely an attempt to illustrate the sinusoidal nature of the pressure-time

107
fluctuations. Do not conclude that sound is a transverse wave that has crests and troughs. Sound
waves traveling through air are indeed longitudinal waves with compressions and rarefactions.
As sound passes through air (or any fluid medium), the particles of the air do not vibrate in a
transverse manner. Do not be misled - sound waves traveling through air are longitudinal waves.

Exercise 2.2
1. A sound wave is a pressure wave; regions of high (compressions) and low pressure
(rarefactions) are established as the result of the vibrations of the sound source. These
compressions and rarefactions result because sound

a. is more dense than air and thus has more inertia, causeng the bunching up of sound.

b. waves have a speed that is dependent only upon the properties of the medium.

c. is like all waves; it is able to bend into the regions of space behind obstacles.

d. is able to reflect off fixed ends and interfere with incident waves

e. vibrates longitudinally; the longitudinal movement of air produces pressure fluctuations.

108
2.2 Sound Properties and Their Perception
Pitch and Frequency | Intensity and the Decibel Scale | The Speed of Sound | The Human Ear

2.2.1 Pitch and Frequency


A sound wave, like any other wave, is introduced into a medium by a vibrating object. The
vibrating object is the source of the disturbance that moves through the
medium. The vibrating object that creates the disturbance could be the
vocal cords of a person, the vibrating string and sound board of a guitar
or violin, the vibrating tines of a tuning fork, or the vibrating
diaphragm of a radio speaker. Regardless of what vibrating object is
creating the sound wave, the particles of the medium through which the sound moves is vibrating
in a back and forth motion at a given frequency. The frequency of a wave refers to how often the
particles of the medium vibrate when a wave passes through the medium. The frequency of a
wave is measured as the number of complete back-and-forth vibrations of a particle of the
medium per unit of time. If a particle of air undergoes 1000 longitudinal vibrations in 2 seconds,
then the frequency of the wave would be 500 vibrations per second. A commonly used unit for
frequency is the Hertz (abbreviated Hz), where

1 Hertz = 1 vibration/second

As a sound wave moves through a medium, each particle of the medium vibrates at the same
frequency. This is sensible since each particle vibrates due to the motion of its nearest neighbor.
The first particle of the medium begins vibrating, at say 500 Hz, and begins to set the second
particle into vibrational motion at the same frequency of 500 Hz. The second particle begins
vibrating at 500 Hz and thus sets the third particle of the medium into vibrational motion at 500
Hz. The process continues throughout the medium; each particle vibrates at the same frequency.
And of course the frequency at which each particle vibrates is the same as the frequency of the
original source of the sound wave. Subsequently, a guitar string vibrating at 500 Hz will set the
air particles in the room vibrating at the same frequency of 500 Hz, which carries a sound signal
to the ear of a listener, which is detected as a 500 Hz sound wave.

The back-and-forth vibrational motion of the particles of the medium would not be the only
observable phenomenon occurring at a given frequency. Since a sound wave is a pressure wave,
a detector could be used to detect oscillations in pressure from a high pressure to a low pressure
and back to a high pressure. As the compressions (high pressure) and rarefactions (low pressure)
move through the medium, they would reach the detector at a given frequency. For example, a
compression would reach the detector 500 times per second if the frequency of the wave were
500 Hz. Similarly, a rarefaction would reach the detector 500 times per second if the frequency
of the wave were 500 Hz. The frequency of a sound wave not only refers to the number of back-
and-forth vibrations of the particles per unit of time, but also refers to the number of
compressions or rarefactions that pass a given point per unit of time. A detector could be used to

109
detect the frequency of these pressure oscillations over a given period of time. The typical output
provided by such a detector is a pressure-time plot as shown below.

Fig. 2.8

Since a pressure-time plot shows the fluctuations in pressure over time, the period of the sound
wave can be found by measuring the time between successive high pressure points
(corresponding to the compressions) or the time between successive low pressure points
(corresponding to the rarefactions). As discussed in an earlier chapter, the frequency is simply
the reciprocal of the period. For this reason, a sound wave with a high frequency would
correspond to a pressure time plot with a small period - that is, a plot corresponding to a small
amount of time between successive high pressure points. Conversely, a sound wave with a low
frequency would correspond to a pressure time plot with a large period - that is, a plot
corresponding to a large amount of time between successive high pressure points. The diagram
below shows two pressure-time plots, one corresponding to a high frequency and the other to a
low frequency.

Fig. 2.9

The ears of a human (and other animals) are sensitive detectors capable of detecting the
fluctuations in air pressure that impinge upon the eardrum. The mechanics of the ear's detection
ability will be discussed later in this chapter. For now, it is sufficient to say that the human ear is
capable of detecting sound waves with a wide range of frequencies, ranging between
approximately 20 Hz to 20 000 Hz. Any sound with a frequency below the audible range of
hearing (i.e., less than 20 Hz) is known as an infrasound and any sound with a frequency above
the audible range of hearing (i.e., more than 20 000 Hz) is known as an ultrasound. Humans are
not alone in their ability to detect a wide range of frequencies. Dogs can detect frequencies as

110
low as approximately 50 Hz and as high as 45 000 Hz. Cats can detect frequencies as low as
approximately 45 Hz and as high as 85 000 Hz. Bats, being nocturnal creature, must rely on
sound echolocation for navigation and hunting. Bats can detect frequencies as high as 120 000
Hz. Dolphins can detect frequencies as high as 200 000 Hz. While dogs, cats, bats, and dolphins
have an unusual ability to detect ultrasound, an elephant possesses the unusual ability to detect
infrasound, having an audible range from approximately 5 Hz to approximately 10 000 Hz.

The sensation of a frequency is commonly referred to as the pitch of a sound. A high pitch sound
corresponds to a high frequency sound wave and a low pitch sound corresponds to a low
frequency sound wave. Amazingly, many people, especially those who have been musically
trained, are capable of detecting a difference in frequency between two separate sounds that is as
little as 2 Hz. When two sounds with a frequency difference of greater than 7 Hz are played
simultaneously, most people are capable of detecting the presence of a complex wave pattern
resulting from the interference and superposition of the two sound waves. Certain sound waves
when played (and heard) simultaneously will produce a particularly pleasant sensation when
heard, are said to be consonant. Such sound waves form the basis of intervals in music. For
example, any two sounds whose frequencies make a 2:1 ratio are said to be separated by an
octave and result in a particularly pleasing sensation when heard. That is, two sound waves
sound good when played together if one sound has twice the frequency of the other. Similarly
two sounds with a frequency ratio of 5:4 are said to be separated by an interval of a third; such
sound waves also sound good when played together. Examples of other sound wave intervals and
their respective frequency ratios are listed in the table below.

Table. 2.1
Interval Frequency Ratio Examples
Octave 2:1 512 Hz and 256 Hz
Third 5:4 320 Hz and 256 Hz
Fourth 4:3 342 Hz and 256 Hz
Fifth 3:2 384 Hz and 256 Hz

The ability of humans to perceive pitch is associated with the frequency of the sound wave that
impinges upon the ear. Because sound waves traveling through air are longitudinal waves that
produce high- and low-pressure disturbances of the particles of the air at a given frequency, the
ear has an ability to detect such frequencies and associate them with the pitch of the sound. But
pitch is not the only property of a sound wave detectable by the human ear. In the next sub-
section of Section 2.2, we will investigate the ability of the ear to perceive the intensity of a
sound wave.

111
Exercise 2.3
1. Two musical notes that have a frequency ratio of 2:1 are said to be separated by an octave. A
musical note that is separated by an octave from middle C (256 Hz) has a frequency of _____.

a. 128 Hz b. 254 Hz c. 258 Hz


d. 345 Hz e. none of these

2.2.2 Intensity and the Decibel Scale


Sound waves are introduced into a medium by the vibration of an object. For example, a
vibrating guitar string forces surrounding air molecules to be compressed and expanded, creating
a pressure disturbance consisting of an alternating pattern of
compressions and rarefactions. The disturbance then travels
from particle to particle through the medium, transporting
energy as it moves. The energy that is carried by the
disturbance was originally imparted to the medium by the
vibrating string. The amount of energy that is transferred to
the medium is dependent upon the amplitude of vibrations of
the guitar string. If more energy is put into the plucking of
the string (that is, more work is done to displace the string a Fig. 2.10
greater amount from its rest position), then the string vibrates with a greater amplitude. The
greater amplitude of vibration of the guitar string thus imparts more energy to the medium,
causing air particles to be displaced a greater distance from their rest position. Subsequently, the
amplitude of vibration of the particles of the medium is increased, corresponding to an increased
amount of energy being carried by the particles. This relationship between energy and amplitude
was discussed in more detail in a previous Chapter.

The amount of energy that is transported past a given area of the medium per unit of time is
known as the intensity of the sound wave. The greater the amplitude of vibrations of the
particles of the medium, the greater the rate at which energy is transported through it, and the
more intense that the sound wave is. Intensity is the energy/time/area; and since the energy/time
ratio is equivalent to the quantity power, intensity is simply the power/area.

112
Typical units for expressing the intensity of a sound wave are Watts/meter2 .

As a sound wave carries its energy through a two-dimensional or three-dimensional medium, the
intensity of the sound wave decreases with increasing distance from the
source. The decrease in intensity with increasing distance is explained by
the fact that the wave is spreading out over a circular (2 dimensions) or
spherical (3 dimensions) surface and thus the energy of the sound wave is
being distributed over a greater surface area. The diagram at the right
shows that the sound wave in a 2-dimensional medium is spreading out
in space over a circular pattern. Since energy is conserved and the area
through which this energy is transported is increasing, the power (being a
quantity that is measured on a per area basis) must decrease. The
mathematical relationship between intensity and distance is sometimes
referred to as an inverse square relationship. The intensity varies Fig. 2.11
inversely with the square of the distance from the source. So if the distance from the source is
doubled (increased by a factor of 2), then the intensity is quartered (decreased by a factor of 4).
Similarly, if the distance from the source is quadrupled, then the intensity is decreased by a
factor of 16. Applied to the diagram at the right, the intensity at point B is one-fourth the
intensity as point A and the intensity at point C is one-sixteenth the intensity at point A. Since
the intensity-distance relationship is an inverse relationship, an increase in one quantity
corresponds to a decrease in the other quantity. And since the intensity-distance relationship is an
inverse square relationship, whatever factor by which the distance is increased, the intensity is
decreased by a factor equal to the square of the distance change factor. The sample data in the
table below illustrate the inverse square relationship between power and distance.

Table. 2.2
Distance Intensity
1m 160 units
2m 40 units
3m 17.8 units
4m 10 units

Humans are equipped with very sensitive ears capable of detecting sound waves of extremely
low intensity. The faintest sound that the typical human ear can detect has an intensity of 1*10-12
W/m2 . This intensity corresponds to a pressure wave in which a compression of the particles of
the medium increases the air pressure in that compressional region by a mere 0.3 billionth of an
atmosphere. A sound with an intensity of 1*10-12 W/m2 corresponds to a sound that will displace
particles of air by a mere one-billionth of a centimeter. The human ear can detect such a sound.
WOW! This faintest sound that a human ear can detect is known as the threshold of hearing.
The most intense sound that the ear can safely detect without suffering any physical damage is
more than one billion times more intense than the threshold of hearing.

113
Since the range of intensities that the human ear can detect is so large, the scale that is frequently
used by physicists to measure intensity is a scale based on multiples of 10. This type of scale is
sometimes referred to as a logarithmic scale. The scale for measuring intensity is the decibel
scale. The threshold of hearing is assigned a sound level of 0 decibels (abbreviated 0 dB); this
sound corresponds to an intensity of 1*10-12 W/m2 . A sound that is 10 times more intense ( 1*10-
11
W/m2 ) is assigned a sound level of 10 dB. A sound that is 10*10 or 100 times more intense
(1*10-10 W/m2 ) is assigned a sound level of 20 db. A sound that is 10*10*10 or 1000 times more
intense (1*10-9 W/m2 ) is assigned a sound level of 30 db. A sound that is 10*10*10*10 or 10000
times more intense (1*10-8 W/m2 ) is assigned a sound level of 40 db. Observe that this scale is
based on powers or multiples of 10. If one sound is 10x times more intense than another sound,
then it has a sound level that is 10*x more decibels than the less intense sound. The table below
lists some common sounds with an estimate of their intensity and decibel level.

Table. 2.3
Intensity # of Times
Source Intensity
Level Greater Than TOH
-12 2
Threshold of Hearing (TOH) 1*10 W/m 0 dB 100
Rustling Leaves 1*10-11 W/m2 10 dB 101
Whisper 1*10-10 W/m2 20 dB 102
Normal Conversation 1*10-6 W/m2 60 dB 106
Busy Street Traffic 1*10-5 W/m2 70 dB 107
Vacuum Cleaner 1*10-4 W/m2 80 dB 108
Large Orchestra 6.3*10-3 W/m2 98 dB 109.8
Walkman at Maximum Level 1*10-2 W/m2 100 dB 1010
Front Rows of Rock Concert 1*10-1 W/m2 110 dB 1011
Threshold of Pain 1*101 W/m2 130 dB 1013
Military Jet Takeoff 1*102 W/m2 140 dB 1014
Instant Perforation of Eardrum 1*104 W/m2 160 dB 1016

While the intensity of a sound is a very objective quantity that can be measured with sensitive
instrumentation, the loudness of a sound is more of a subjective response that will vary with a
number of factors. The same sound will not be perceived to have the same loudness to all
individuals. Age is one factor that affects the human ear's response to a sound. Quite obviously,
your grandparents do not hear like they used to. The same intensity sound would not be
perceived to have the same loudness to them as it would to you. Furthermore, two sounds with
the same intensity but different frequencies will not be perceived to have the same loudness.
Because of the human ear's tendency to amplify sounds having frequencies in the range from
1000 Hz to 5000 Hz, sounds with these intensities seem louder to the human ear. Despite the

114
distinction between intensity and loudness, it is safe to state that the more intense sounds will be
perceived to be the loudest sounds.

Exercise 2.4
1. A mosquito's buzz is often rated with a decibel rating of 40 dB. Normal conversation is often
rated at 60 dB. How many times more intense is normal conversation compared to a mosquito's
buzz?

a. 2 b. 20 c. 100 d. 200 e. 400

Table. Q.2.2
2. The table at the right represents the decibel level for several
sound sources. Use the table to make comparisons of the
intensities of the following sounds.

How many times more intense is the front row of a Smashin'


Pumpkins concert than ...

a. ... the 15th row of the same concert?

b. ... the average factory?

c. ... normal speech?

d. ... the library after school?

e. ... the sound that most humans can just barely hear?

3. On a good night, the front row of the Twisted Sister concert would surely result in a 120 dB
sound level. An IPod produces 100 dB. How many IPods would be needed to produce the same
intensity as the front row of the Twisted Sister concert?

115
2.2.3 The Speed of Sound
A sound wave is a pressure disturbance that travels through a medium by means of particle-to-
particle interaction. As one particle becomes disturbed, it exerts a force on the next adjacent
particle, thus disturbing that particle from rest and transporting the energy through the medium.
Like any wave, the speed of a sound wave refers to how fast the
disturbance is passed from particle to particle. While frequency refers
to the number of vibrations that an individual particle makes per unit
of time, speed refers to the distance that the disturbance travels per
unit of time. Always be cautious to distinguish between the two often-
confused quantities of speed (how fast...) and frequency (how often...).

Since the speed of a wave is defined as the distance that a point on a wave (such as a
compression or a rarefaction) travels per unit of time, it is often expressed in units of
meters/second (abbreviated m/s). In equation form, this is

speed = distance/time

The faster a sound wave travels, the more distance it will cover in the same period of time. If a
sound wave were observed to travel a distance of 700 meters in 2 seconds, then the speed of the
wave would be 350 m/s. A slower wave would cover less distance - perhaps 660 meters - in the
same time period of 2 seconds and thus have a speed of 330 m/s. Faster waves cover more
distance in the same period of time.

Factors Affecting Wave Speed

The speed of any wave depends upon the properties of the medium through which the wave is
traveling. Typically there are two essential types of properties that effect wave speed - inertial
properties and elastic properties. Elastic properties are those properties related to the tendency
of a material to maintain its shape and not deform whenever a force or stress is applied to it. A
material such as steel will experience a very small deformation of shape (and dimension) when a
stress is applied to it. Steel is a rigid material with a high elasticity. On the other hand, a material
such as a rubber band is highly flexible; when a force is applied to stretch the rubber band, it
deforms or changes its shape readily. A small stress on the rubber band causes a large
deformation. Steel is considered to be a stiff or rigid material, whereas a rubber band is
considered a flexible material. At the particle level, a stiff or rigid material is characterized by
atoms and/or molecules with strong attractions for each other. When a force is applied in an
attempt to stretch or deform the material, its strong particle interactions prevent this deformation
and help the material maintain its shape. Rigid materials such as steel are considered to have a
high elasticity. (Elastic modulus is the technical term). The phase of matter has a tremendous
impact upon the elastic properties of the medium. In general, solids have the strongest
interactions between particles, followed by liquids and then gases. For this reason, longitudinal
sound waves travel faster in solids than they do in liquids than they do in gases. Even though the

116
inertial factor may favor gases, the elastic factor has a greater influence on the speed (v) of a
wave, thus yielding this general pattern:

vsolids > vliquids > vgase s

Inertial properties are those properties related to the material's tendency to be sluggish to changes
in its state of motion. The density of a medium is an example of an inertial property. The
greater the inertia (i.e., mass density) of individual particles of the medium, the less responsive
they will be to the interactions between neighboring particles and the slower that the wave will
be. As stated above, sound waves travel faster in solids than they do in liquids than they do in
gases. However, within a single phase of matter, the inertial property of density tends to be the
property that has a greatest impact upon the speed of sound. A sound wave will travel faster in a
less dense material than a more dense material. Thus, a sound wave will travel nearly three times
faster in Helium than it will in air. This is mostly due to the lower mass of Helium particles as
compared to air particles.

The speed of a sound wave in air depends upon the properties of the air, mostly the temperature,
and to a lesser degree, the humidity. Humidity is the result of water vapor being present in air.
Like any liquid, water has a tendency to evaporate. As it does, particles of gaseous water become
mixed in the air. This additional matter will affect the mass density of the air (an inertial
property). The temperature will affect the strength of the particle interactions (an elastic
property). At normal atmospheric pressure, the temperature dependence of the speed of a sound
wave through dry air is approximated by the following equation:

v = 331 m/s + (0.6 m/s/C)•T

where T is the temperature of the air in degrees Celsius. Using this equation to determine the
speed of a sound wave in air at a temperature of 20 degrees Celsius yields the following solution.

v = 331 m/s + (0.6 m/s/C)•T

v = 331 m/s + (0.6 m/s/C)•(20 C)

v = 331 m/s + 12 m/s

v = 343 m/s

(The above equation relating the speed of a sound wave in air to the temperature provides
reasonably accurate speed values for temperatures between 0 and 100 Celsius. The equation
itself does not have any theoretical basis; it is simply the result of inspecting temperature-speed
data for this temperature range. Other equations do exist that are based upon theoretical
reasoning and provide accurate data for all temperatures. Nonetheless, the equation above will be
sufficient for our use as introductory Physics students.)

117
Using Wave Speed to Determine Distances

At normal atmospheric pressure and a temperature of 20 degrees Celsius, a sound wave will
travel at approximately 343 m/s; this is approximately equal to 750 miles/hour. While this speed
may seem fast by human standards (the fastest humans can sprint at approximately 11 m/s and
highway speeds are approximately 30 m/s), the speed of a sound wave is slow in comparison to
the speed of a light wave. Light travels through air at a speed of approximately 300 000 000 m/s;
this is nearly 900 000 times the speed of sound. For this reason, humans can observe a detectable
time delay between the thunder and the lightning during a storm. The arrival of the light wave
from the location of the lightning strike occurs in so little time that it is essentially negligible.
Yet the arrival of the sound wave from the location of the lightning strike occurs much later. The
time delay between the arrival of the light wave (lightning) and the arrival of the sound wave
(thunder) allows a person to approximate his/her distance from the storm location. For instance if
the thunder is heard 3 seconds after the lightning is seen, then sound (whose speed is
approximated as 345 m/s) has traveled a distance of

distance = v • t = 345 m/s • 3 s = 1035 m

If this value is converted to miles (divide by 1600 m/1 mi), then the storm is a distance of 0.65
miles away.

Another phenomenon related to the perception of time delays between two events is an echo. A
person can often perceive a time delay between the production of a sound and the arrival of a
reflection of that sound off a distant barrier. If you have ever made a holler within a canyon,
perhaps you have heard an echo of your holler off a distant canyon wall. The time delay between
the holler and the echo corresponds to the time for the holler to travel the round-trip distance to
the canyon wall and back. A measurement of this time would allow a person to estimate the one-
way distance to the canyon wall. For instance if an echo is heard 1.40 seconds after making the
holler, then the distance to the canyon wall can be found as follows:

distance = v • t = 345 m/s • 0.70 s = 242 m

The canyon wall is 242 meters away. You might have noticed that the time of 0.70 seconds is
used in the equation. Since the time delay corresponds to the time for the holler to travel the
round-trip distance to the canyon wall and back, the one-way distance to the canyon wall
corresponds to one-half the time delay.

While an echo is of relatively minimal importance to humans, echolocation is an essential trick


of the trade for bats. Being a nocturnal creature, bats must use sound waves to navigate and hunt.
They produce short bursts of ultrasonic sound waves that reflect off objects in their surroundings
and return. Their detection of the time delay between the sending and receiving of the pulses
allows a bat to approximate the distance to surrounding objects. Some bats, known as Doppler
bats, are capable of detecting the speed and direction of any moving objects by monitoring the
changes in frequency of the reflected pulses. These bats are utilizing the physics of the Doppler
effect discussed in chapter 1 (and also to be discussed later in Section 2.3). This method of
echolocation enables a bat to navigate and to hunt.

118
The Wave Equation Revisited

Like any wave, a sound wave has a speed that is mathematically related to the frequency and the
wavelength of the wave. As discussed in a previous chapter, the mathematical relationship
between speed, frequency and wavelength is given by the following equation.

Speed = Wavelength • Frequency

Using the symbols v, , and f, the equation can be rewritten as

v=f•
The above equation is useful for solving mathematical problems
related to the speed, frequency and wavelength relationship.
However, one important misconception could be conveyed by the
equation. Even though wave speed is calculated using the
frequency and the wavelength, the wave speed is not dependent
upon these quantities. An alteration in wavelength does not affect
(i.e., change) wave speed. Rather, an alteration in wavelength
affects the frequency in an inverse manner. A doubling of the wavelength results in a halving of
the frequency; yet the wave speed is not changed. The speed of a sound wave depends on the
properties of the medium through which it moves and the only way to change the speed is to
change the properties of the medium.

Exercuse 2.5
1. An automatic focus camera is able to focus on objects by use of an ultrasonic sound wave. The
camera sends out sound waves that reflect off distant objects and return to the camera. A sensor
detects the time it takes for the waves to return and then determines the distance an object is from
the camera. If a sound wave (speed = 340 m/s) returns to the camera 0.150 seconds after leaving
the camera, how far away is the object?

2. On a hot summer day, a pesky little mosquito produced its warning sound near your ear. The
sound is produced by the beating of its wings at a rate of about 600 wing beats per second.

a. What is the frequency in Hertz of the sound wave?

119
b. Assuming the sound wave moves with a velocity of 350 m/s, what is the wavelength of the
wave?

3. Doubling the frequency of a wave source doubles the speed of the waves.

a. True b. False

4. Playing middle C on the piano keyboard produces a sound with a frequency of 256 Hz.
Assuming the speed of sound in air is 345 m/s, determine the wavelength of the sound
corresponding to the note of middle C.

5. Most people can detect frequencies as high as 20 000 Hz. Assuming the speed of sound in air
is 345 m/s, determine the wavelength of the sound corresponding to this upper range of audible
hearing.

6. An elephant produces a 10 Hz sound wave. Assuming the speed of sound in air is 345 m/s,
determine the wavelength of this infrasonic sound wave.

7. Determine the speed of sound on a cold winter day (T=3 degrees C).

8. Miles Tugo is camping in Glacier National Park. In the midst of a glacier canyon, he makes a
loud holler. He hears an echo 1.22 seconds later. The air temperature is 20 degrees C. How far
away are the canyon walls?

9. Two sound waves are traveling through a container of unknown gas. Wave A has a
wavelength of 1.2 m. Wave B has a wavelength of 3.6 m. The velocity of wave B must be
__________ the velocity of wave A.

a. one-ninth b. one-third
c. the same as d. three times larger than

10. Two sound waves are traveling through a container of unknown gas. Wave A has a
wavelength of 1.2 m. Wave B has a wavelength of 3.6 m. The frequency of wave B must be
__________ the frequency of wave A.

a. one-ninth b. one-third
c. the same as d. three times larger than

120
2.2.4 The Human Ear
Understanding how humans hear is a complex subject involving the fields of physiology,
psychology and acoustics. In this sub-section, we will focus on the acoustics (the branch of
physics pertaining to sound) of hearing. We will attempt to understand how the human ear serves
as an astounding transducer, converting sound energy to mechanical energy to a nerve impulse
that is transmitted to the brain. The ear's ability to do this allows us to perceive the pitch of
sounds by detection of the wave's frequencies, the loudness of sound by detection of the wave's
amplitude and the timbre of the sound by the detection of the various frequencies that make up a
complex sound wave.

The ear consists of three basic parts - the outer ear, the middle ear, and the inner ear. Each part of
the ear serves a specific purpose in the task of detecting and interpreting sound. The outer ear
serves to collect and channel sound to the middle ear. The middle ear serves to transform the
energy of a sound wave into the internal vibrations of the bone structure of the middle ear and
ultimately transform these vibrations into a compressional wave in the inner ear. The inner ear
serves to transform the energy of a compressional wave within the inner ear fluid into nerve
impulses that can be transmitted to the brain. The three parts of the ear are shown below.

Fig. 2.12

The outer ear consists of an earflap and an approximately 2-cm long ear canal. The earflap
provides protection for the middle ear in order to prevent damage to the eardrum. The outer ear
also channels sound waves that reach the ear through the ear canal to the eardrum of the middle
ear. Because of the length of the ear canal, it is capable of amplifying sounds with frequencies of
approximately 3000 Hz. As sound travels through the outer ear, the sound is still in the form of a
pressure wave, with an alternating pattern of high and low pressure regions. It is not until the
sound reaches the eardrum at the interface of the outer and the middle ear that the energy of the
mechanical wave becomes converted into vibrations of the inner bone structure of the ear.

121
The middle ear is an air-filled cavity that consists of an eardrum and three tiny, interconnected
bones - the hammer, anvil, and stirrup. The eardrum is a very durable and tightly stretched
membrane that vibrates as the incoming pressure waves reach it. As shown below, a compression
forces the eardrum inward and a rarefaction forces the eardrum outward, thus vibrating the
eardrum at the same frequency of the sound wave.

Fig. 2.13

Being connected to the hammer, the movements of the eardrum will set the hammer, anvil, and
stirrup into motion at the same frequency of the sound wave. The stirrup is connected to the inner
ear; and thus the vibrations of the stirrup are transmitted to the fluid of the inner ear and create a
compression wave within the fluid. The three tiny bones of the middle ear act as levers to
amplify the vibrations of the sound wave. Due to a mechanical advantage, the displacements of
the stirrup are greater than that of the hammer. Furthermore, since the pressure wave striking the
large area of the eardrum is concentrated into the smaller area of the stirrup, the force of the
vibrating stirrup is nearly 15 times larger than that of the eardrum. This feature enhances our
ability of hear the faintest of sounds. The middle ear is an air-filled cavity that is connected by
the Eustachian tube to the mouth. This connection allows for the equalization of pressure within
the air-filled cavities of the ear. When this tube becomes clogged during a cold, the ear cavity is
unable to equalize its pressure; this will often lead to earaches and other pains.

The inner ear consists of a cochlea, the semicircular canals, and the auditory nerve. The cochlea
and the semicircular canals are filled with a water-like fluid. The fluid and nerve cells of the
semicircular canals provide no role in the task of hearing; they merely serve as accelerometers
for detecting accelerated movements and assisting in the task of maintaining balance. The
cochlea is a snail-shaped organ that would stretch to approximately 3 cm. In addition to being
filled with fluid, the inner surface of the cochlea is lined with over 20 000 hair-like nerve cells
that perform one of the most critical roles in our ability to hear. These nerve cells differ in length
by minuscule amounts; they also have different degrees of resiliency to the fluid that passes over
them. As a compressional wave moves from the interface between the hammer of the middle ear
and the oval window of the inner ear through the cochlea, the small hair-like nerve cells will be
set in motion. Each hair cell has a natural sensitivity to a particular frequency of vibration. When
the frequency of the compressional wave matches the natural frequency of the nerve cell, that
nerve cell will resonate with a larger amplitude of vibration. This increased vibrational amplitude
induces the cell to release an electrical impulse that passes along the auditory nerve towards the
brain. In a process that is not clearly understood, the brain is capable of interpreting the qualities
of the sound upon reception of these electric nerve impulses.

122
2.3 Behavior of Sound Waves
Interference and Beats | The Doppler Effect and Shock Waves
Boundary Behavior | Reflection, Refraction, and Diffraction

2.3.1 Interference and Beats


Wave interference is the phenomenon that occurs when two waves meet while traveling along
the same medium. The interference of waves causes the medium to take on a shape that results
from the net effect of the two individual waves upon the particles of the medium. As mentioned
in a previous chapter, if two upward displaced pulses having the same shape meet up with one
another while traveling in opposite directions along a medium, the medium will take on the
shape of an upward displaced pulse with twice the amplitude of the two interfering pulses. This
type of interference is known as constructive interference. If an upward displaced pulse and a
downward displaced pulse having the same shape meet up with one another while traveling in
opposite directions along a medium, the two pulses will cancel each other's effect upon the
displacement of the medium and the medium will assume the equilibrium position. This type of
interference is known as destructive interference. The diagrams below show two waves - one is
blue and the other is red - interfering in such a way to produce a resultant shape in a medium; the
resultant is shown in green. In two cases (on the left and in the middle), constructive interference
occurs and in the third case (on the far right, destructive interference occurs.

Fig. 2.14
But how can sound waves that do not possess upward and downward displacements interfere
constructively and destructively? Sound is a pressure wave that consists of compressions and
rarefactions. As a compression passes through a section of a medium, it tends to pull particles
together into a small region of space, thus creating a high-pressure region. And as a rarefaction
passes through a section of a medium, it tends to push particles apart, thus creating a low-
pressure region. The interference of sound waves causes the particles of the medium to behave in
a manner that reflects the net effect of the two individual waves upon the particles. For example,
if a compression (high pressure) of one wave meets up with a compression (high pressure) of a
second wave at the same location in the medium, then the net effect is that that particular
location will experience an even greater pressure. This is a form of constructive interference. If
two rarefactions (two low-pressure disturbances) from two different sound waves meet up at the
same location, then the net effect is that that particular location will experience an even lower
pressure. This is also an example of constructive interference. Now if a particular location along

123
the medium repeatedly experiences the interference of two compressions followed up by the
interference of two rarefactions, then the two sound waves will continually reinforce each other
and produce a very loud sound. The loudness of the sound is the result of the particles at that
location of the medium undergoing oscillations from very high to very low pressures. As
mentioned in a previous chapter, locations along the medium where constructive interference
continually occurs are known as anti-nodes. The animation below shows two sound waves
interfering constructively in order to produce very large oscillations in pressure at a variety of
anti-nodal locations. Note that compressions are labeled with a C and rarefactions are labeled
with an R.

Fig. 2.15

Now if two sound waves interfere at a given location in such a way that the compression of one
wave meets up with the rarefaction of a second wave, destructive interference results. The net
effect of a compression (which pushes particles together) and a rarefaction (which pulls particles
apart) upon the particles in a given region of the medium is to not even cause a displacement of
the particles. The tendency of the compression to push particles together is canceled by the
tendency of the rarefactions to pull particles apart; the particles would remain at their rest
position as though there wasn't even a disturbance passing through them. This is a form of
destructive interference. Now if a particular location along the medium repeatedly experiences
the interference of a compression and rarefaction followed up by the interference of a rarefaction
and a compression, then the two sound waves will continually cancel each other and no sound is
heard. The absence of sound is the result of the particles remaining at rest and behaving as
though there were no disturbance passing through it. Amazingly, in a situation such as this, two
sound waves would combine to produce no sound. As mentioned in a previous chapter, locations
along the medium where destructive interference continually occurs are known as nodes.

Two Source Sound Interference

A popular Physics demonstration involves the interference of two sound waves from two
speakers. The speakers are set approximately 1-meter apart and produced identical tones. The
two sound waves traveled through the air in front of the speakers, spreading our through the
room in spherical fashion. A snapshot in time of the appearance of these waves is shown in the
diagram below. In the diagram, the compressions of a wavefront are represented by a thick line
and the rarefactions are represented by thin lines. These two waves interfere in such a manner as
to produce locations of some loud sounds and other locations of no sound. Of course the loud
sounds are heard at locations where compressions meet compressions or rarefactions meet
rarefactions and the "no sound" locations appear wherever the compressions of one of the waves
meet the rarefactions of the other wave. If you were to plug one ear and turn the other ear
towards the place of the speakers and then slowly walk across the room parallel to the plane of
the speakers, then you would encounter an amazing phenomenon. You would alternatively hear
loud sounds as you approached anti-nodal locations and virtually no sound as you approached

124
nodal locations. (As would commonly be observed, the nodal locations are not true nodal
locations due to reflections of sound waves off the walls. These reflections tend to fill the entire
room with reflected sound. Even though the sound waves that reach the nodal locations directly
from the speakers destructively interfere, other waves reflecting off the walls tend to reach that
same location to produce a pressure disturbance.)

Fig. 2.16

Destructive interference of sound waves becomes an important issue in the design of concert
halls and auditoriums. The rooms must be designed in such as way as to reduce the amount of
destructive interference. Interference can occur as the result of sound from two speakers meeting
at the same location as well as the result of sound from a speaker meeting with sound reflected
off the walls and ceilings. If the sound arrives at a given location such that compressions meet
rarefactions, then destructive interference will occur resulting in a reduction in the loudness of
the sound at that location. One means of reducing the severity of destructive interference is by
the design of walls, ceilings, and baffles that serve to absorb sound rather than reflect it. This will
be discussed in more detail later in Section 2.3.

The destructive interference of sound waves can also be used advantageously in noise reduction
systems. Earphones have been produced that can be used by factory and construction workers to
reduce the noise levels on their jobs. Such earphones capture sound from the environment and
use computer technology to produce a second sound wave that one-half cycle out of phase. The
combination of these two sound waves within the headset will result in destructive interference
and thus reduce a worker's exposure to loud noise.

Musical Beats and Intervals

Interference of sound waves has widespread applications in the world of music. Music seldom
consists of sound waves of a single frequency played continuously. Few music enthusiasts would
be impressed by an orchestra that played music consisting of the note with a pure tone played by
all instruments in the orchestra. Hearing a sound wave of 256 Hz (middle C) would become
rather monotonous (both literally and figuratively). Rather, instruments are known to produce
overtones when played resulting in a sound that consists of a multiple of frequencies. Such
instruments are described as being rich in tone color. And even the best choirs will earn their
money when two singers sing two notes (i.e., produce two sound waves) that are an octave apart.
Music is a mixture of sound waves that typically have whole number ratios between the
125
frequencies associated with their notes. In fact, the major distinction between music and noise is
that noise consists of a mixture of frequencies whose mathematical relationship to one another is
not readily discernible. On the other hand, music consists of a mixture of frequencies that have a
clear mathematical relationship between them. While it may be true that "one person's music is
another person's noise" (e.g., your music might be thought of by your parents as being noise), a
physical analysis of musical sounds reveals a mixture of sound waves that are mathematically
related.

To demonstrate this nature of music, let's consider one of the simplest mixtures of two different
sound waves - two sound waves with a 2:1 frequency ratio. This combination of waves is known
as an octave. A simple sinusoidal plot of the wave pattern for two such waves is shown below.
Note that the red wave has two times the frequency of the blue wave. Also observe that the
interference of these two waves produces a resultant (in green) that has a periodic and repeating
pattern. One might say that two sound waves that have a clear whole number ratio between their
frequencies interfere to produce a wave with a regular and repeating pattern. The result is music.

Fig. 2.17
Another simple example of two sound waves with a clear mathematical relationship between
frequencies is shown below. Note that the red wave has three-halves the frequency of the blue
wave. In the music world, such waves are said to be a fifth apart and represent a popular musical
interval. Observe once more that the interference of these two waves produces a resultant (in
green) that has a periodic and repeating pattern. It should be said again: two sound waves that
have a clear whole number ratio between their frequencies interfere to produce a wave with a
regular and repeating pattern; the result is music.

Fig. 2.18

Finally, the diagram below illustrates the wave pattern produced by two dissonant or displeasing
sounds. The diagram shows two waves interfering, but this time there is no simple mathematical
relationship between their frequencies (in computer terms, one has a wavelength of 37 and the
other has a wavelength 20 pixels). Observe (look carefully) that the pattern of the resultant is
neither periodic nor repeating (at least not in the short sample of time that is shown). The
message is clear: if two sound waves that have no simple mathematical relationship between

126
their frequencies interfere to produce a wave, the result will be an irregular and non-repeating
pattern. This tends to be displeasing to the ear.

Fig. 2.19

A final application of physics to the world of music pertains to the topic of beats. Beats are the
periodic and repeating fluctuations heard in the intensity of a sound when two sound waves of
very similar frequencies interfere with one another. The diagram below illustrates the wave
interference pattern resulting from two waves (drawn in red and blue) with very similar
frequencies. A beat pattern is characterized by a wave whose amplitude is changing at a regular
rate. Observe that the beat pattern (drawn in green) repeatedly oscillates from zero amplitude to a
large amplitude, back to zero amplitude throughout the pattern. Points of constructive
interference (C.I.) and destructive interference (D.I.) are labeled on the diagram. When
constructive interference occurs between two crests or two troughs, a loud sound is heard. This
corresponds to a peak on the beat pattern (drawn in green). When destructive interference
between a crest and a trough occurs, no sound is heard; this corresponds to a point of no
displacement on the beat pattern. Since there is a clear relationship between the amplitude and
the loudness, this beat pattern would be consistent with a wave that varies in volume at a regular
rate.

Fig. 2.20

The beat frequency refers to the rate at which the volume is heard to be oscillating from high to
low volume. For example, if two complete cycles of high and low volumes are heard every
second, the beat frequency is 2 Hz. The beat frequency is always equal to the difference in

127
frequency of the two notes that interfere to produce the beats. So if two sound waves with
frequencies of 256 Hz and 254 Hz are played simultaneously, a beat frequency of 2 Hz will be
detected. A common physics demonstration involves producing beats using two tuning forks
with very similar frequencies. If a tine on one of two identical tuning forks is wrapped with a
rubber band, then that tuning forks frequency will be lowered. If both tuning forks are vibrated
together, then they produce sounds with slightly different frequencies. These sounds will
interfere to produce detectable beats. The human ear is capable of detecting beats with
frequencies of 7 Hz and below.

A piano tuner frequently utilizes the phenomenon of beats to tune a piano string. She will pluck
the string and tap a tuning fork at the same time. If the two sound sources - the piano string and
the tuning fork - produce detectable beats then their frequencies are not identical. She will then
adjust the tension of the piano string and repeat the process until the beats can no longer be
heard. As the piano string becomes more in tune with the tuning fork, the beat frequency will be
reduced and approach 0 Hz. When beats are no longer heard, the piano string is tuned to the
tuning fork; that is, they play the same frequency. The process allows a piano tuner to match the
strings' frequency to the frequency of a standardized set of tuning forks.

Important Note: Many of the diagrams on this page represent a sound wave by a sine wave.
Such a wave more closely resembles a transverse wave and may mislead people into thinking
that sound is a transverse wave. Sound is not a transverse wave, but rather a longitudinal wave.
Nonetheless, the variations in pressure with time take on the pattern of a sine wave and thus a
sine wave is often used to represent the pressure-time features of a sound wave.

Exercise 2.6
Two speakers are arranged so that sound waves with
the same frequency are produced and radiate through
the room. An interference pattern is created (as
represented in the diagram at the right). The thick
lines in the diagram represent wave crests and the thin
lines represent wave troughs. Use the diagram to
answer the next two questions.

1. At which of the labeled point(s) would constructive Fig.Q. 2.6.1


interference occur?

a. B only

128
b. A, B, and C

c. D, E, and F

d. A and B

2. How many of the six labeled points represent anti-nodes?

a. 1 b. 2 c. 3 d. 4 e. 6

3. A tuning fork with a frequency of 440 Hz is played simultaneously with a fork with a
frequency of 437 Hz. How many beats will be heard over a period of 10 seconds?

4. Why don't we hear beats when different keys on the piano are played at the same time?

2.3.2 The Doppler Effect and Shock Waves


The Doppler effect is a phenomenon observed whenever the source of waves is moving with
respect to an observer. The Doppler effect can be described as the effect produced by a moving
source of waves in which there is an apparent upward shift in frequency for the observer and the
source are approaching and an apparent downward shift in frequency when the observer and the
source is receding. The Doppler effect can be observed to occur with all types of waves - most
notably water waves, sound waves, and light waves. The application of this phenomenon to
water waves was discussed in detail in Chapter 10. In this unit, we will focus on the application
of the Doppler effect to sound.

We are most familiar with the Doppler effect because of our experiences with sound waves.
Perhaps you recall an instance in which a police car or emergency vehicle was traveling towards
you on the highway. As the car approached with its siren blasting, the pitch of the siren sound (a
measure of the siren's frequency) was high; and then suddenly after the car passed by, the pitch
of the siren sound was low. That was the Doppler effect - a shift in the apparent frequency for a
sound wave produced by a moving source.

129
Fig. 2.21
Another common experience is the shift in apparent frequency of the sound of a train horn. As
the train approaches, the sound of its horn is heard at a high pitch and as the train moved away,
the sound of its horn is heard at a low pitch. This is the Doppler effect.

A common Physics demonstration the use of a large Nerf ball equipped with a buzzer that
produces a sound with a constant frequency. The Nerf ball is then through around the room. As
the ball approaches you, you observe a higher pitch than when the ball is at rest. And when the
ball is thrown away from you, you observe a lower pitch than when the ball is at rest. This is the
Doppler effect.

Explaining the Doppler Effect

The Doppler effect is observed because the distance between the source of sound and the
observer is changing. If the source and the observer are approaching, then the distance is
decreasing and if the source and the observer are receding, then the distance is increasing. The
source of sound always emits the same frequency. Therefore, for the same period of time, the
same number of waves must fit between the source and the observer. if the distance is large, then
the waves can be spread apart; but if the distance is small, the waves must be compressed into
the smaller distance. For these reasons, if the source is moving towards the observer, the
observer perceives sound waves reaching him or her at a more frequent rate (high pitch). And if
the source is moving away from the observer, the observer perceives sound waves reaching him
or her at a less frequent rate (low pitch). It is important to note that the effect does not result
because of an actual change in the frequency of the source. The source puts out the same
frequency; the observer only perceives a different frequency because of the relative motion
between them. The Doppler effect is a shift in the apparent or observed frequency and not a shift
in the actual frequency at which the source vibrates.

130
Fig. 2.22

Shock Waves and Sonic Booms

The Doppler effect is observed whenever the speed of the source is


moving slower than the speed of the waves. But if the source actually
moves at the same speed as or faster than the wave itself can move, a
different phenomenon is observed. If a moving source of sound moves
at the same speed as sound, then the source will always be at the
leading edge of the waves that it produces. The diagram at the right
depicts snapshots in time of a variety of wavefronts produced by an
aircraft that is moving at the same speed as sound. The circular lines
represent compressional wavefronts of the sound waves. Notice that Fig. 2.23
these circles are bunched up at the front of the aircraft. This
phenomenon is known as a shock wave. Shock waves are also
produced if the aircraft moves faster than the speed of sound. If a
moving source of sound moves faster than sound, the source will
always be ahead of the waves that it produces. The diagram at the right
depicts snapshots in time of a variety of wavefronts produced by an
aircraft that is moving faster than sound. Note that the circular
compressional wavefronts fall behind the faster moving aircraft (in
actuality, these circles would be spheres). Fig. 2.24

If you are standing on the ground when a supersonic (faster than sound) aircraft passes overhead,
you might hear a sonic boom. A sonic boom occurs as the result of the piling up of
compressional wave-fronts along the conical edge of the wave pattern. These compressional
wave-fronts pile up and interfere to produce a very high-pressure zone. This is shown below.
Instead of these compressional regions (high-pressure regions) reaching you one at a time in
consecutive fashion, they all reach you at once. Since every compression is followed by a
rarefaction, the high-pressure zone will be immediately followed by a low-pressure zone. This
creates a very loud noise.

131
Fig. 2.25

If you are standing on the ground as the supersonic aircraft passes by, there will be a short time
delay and then you will hear the boom - the sonic boom. This boom is merely a loud noise
resulting from the high pressure sound followed by a low pressure sound. Do not be mistaken
into thinking that this boom only happens the instant that the aircraft surpasses the speed of
sound and that it is the signature that the aircraft just attained supersonic speed. Sonic booms are
observed when any aircraft that is traveling faster than the speed of sound passes overhead. It is
not a sign that the aircraft just overcame the sound barrier, but rather a sign that the aircraft is
traveling faster than sound.

Fig. 2.26

Exercise 2.7
1. Suppose you are standing on the passenger-loading platform of the commuter railway line. As
the commuter train approaches the station, it gradually slows down. During this process of
slowing down, the engineer sounds the horn at a constant frequency of 300 Hz. What pitch or
changes in pitch will you perceive as the train approaches you on the loading platform?

132
2.3.3 Boundary Behavior
As a sound wave travels through a medium, it will often reach the end of the medium and
encounter an obstacle or perhaps another medium through which it could travel. When one
medium ends, another medium begins; the interface of the two media is referred to as the
boundary and the behavior of a wave at that boundary is described as its boundary behavior.
The behavior of a wave (or pulse) upon reaching the end of a medium is referred to as boundary
behavior. There are essentially four possible behaviors that a wave could exhibit at a boundary:
reflection (the bouncing off of the boundary), diffraction (the bending around the obstacle
without crossing over the boundary), transmission (the crossing of the boundary into the new
material or obstacle), and refraction (occurs along with transmission and is characterized by the
subsequent change in speed and direction). In this sub-section, the focus will be upon the
reflection behavior of sound waves. Later in Section 2.3, diffraction, transmission, and refraction
will be discussed in more detail.

In Chapter 1, the boundary behavior of a pulse on a rope


was discussed. In that unit, it was mentioned that there are
two types of reflection for waves on ropes: fixed end
reflection and free end reflection. A pulse moving through
a rope will eventually reach its end. Upon reaching the end
of the medium, two things occur:
Fig. 2.27
• A portion of the energy carried by the pulse is reflected and returns towards the left end
of the rope. The disturbance that returns to the left is known as the reflected pulse.
• A portion of the energy carried by the pulse is transmitted into the new medium. If the
rope is attached to a pole (as shown at the right), the pole will receive some of the energy
and begin to vibrate. If the rope is not attached to a pole but rather resting on the ground,
then a portion of the energy is transmitted into the air (the new medium), causing slight
disturbances of the air particles.

The amount of energy that becomes reflected is dependent upon the dissimilarity of the two
media. The more similar that the two media on each side of the boundary are, the less reflection
that occurs and the more transmission that occurs. Conversely, the less similar that the two media
on each side of the boundary are, the more reflection that occurs and the less transmission that
occurs. So if a heavy rope is attached to a light rope (two very dissimilar media), little
transmission and mostly reflection occurs. And if a heavy rope is attached to another heavy rope
(two very similar media), little reflection and mostly transmission occurs.

133
The more similar the medium, the more transmission that occurs.

Fig. 2.28

These principles of reflection can be applied to sound waves. Though a sound wave does not
consist of crests and troughs, they do consist of compressions and rarefactions. If a sound wave
is traveling through a cylindrical tube, it will eventually come to the end of the tube. The end of
the tube represents a boundary between the enclosed air in the tube and the expanse of air outside
of the tube. Upon reaching the end of the tube, the sound wave will undergo partial reflection
and partial transmission. That is, a portion of the energy carried by the sound wave will pass
across the boundary and out of the tube (transmission) and a portion of the energy carried by the
sound wave will reflect off the boundary, remain in the tube and travel in the opposite direction
(reflection). The reflected pulse off the end of the tube can then interfere with any subsequent
incident pulses that are traveling in the opposite direction. If the disturbances within the tube are
the result of perpetual waves of a constant frequency, then interference between the incident and
reflected waves will occur along the length of the tube. The reflection behavior of sound waves
and the subsequent interference that occurs will become important in Section 2.5 during the
discussion of musical instruments. Many musical instruments operate as the result of sound
waves traveling back and forth inside of "tubes" or air columns.

The reflection of sound also becomes important to the design of concert halls and auditoriums.
The acoustics of sound must be considered in the design of such buildings. The most important
considerations include destructive interference and reverberations, both of which are the result of
reflections of sound off the walls and ceilings. Designers attempt to reduce the severity of these
problems by using building materials that reduce the amount of reflection and enhance the
amount of transmission (or absorption) into the walls and ceilings. The most reflective materials
are those that are smooth and hard; such materials are very dissimilar to air and thus reduce the
amount of transmission and increase the amount of reflection. The best materials to use in the
design of concert halls and auditoriums are those materials that are soft. For this reason,
fiberglass and acoustic tile are used in such buildings rather than cement and brick.

134
2.3.4 Reflection, Refraction, and Diffraction
Like any wave, a sound wave doesn't just stop when it reaches the end of the medium or when it
encounters an obstacle in its path. Rather, a sound wave will undergo certain behaviors when it
encounters the end of the medium or an obstacle. Possible behaviors include reflection off the
obstacle, diffraction around the obstacle, and transmission (accompanied by refraction) into the
obstacle or new medium. In this part of Section 3, we will investigate behaviors that have already
been discussed in a previous chapter and apply them towards the reflection, diffraction, and
refraction of sound waves.

When a wave reaches the boundary between one medium another medium, a portion of the wave
undergoes reflection and a portion of the wave undergoes transmission across the boundary. As
discussed in the previous part of Section 2.3, the amount of reflection is
dependent upon the dissimilarity of the two media. For this reason,
acoustically minded builders of auditoriums and concert halls avoid the use of
hard, smooth materials in the construction of their inside halls. A hard
material such as concrete is as dissimilar as can be to the air through which
the sound moves; subsequently, most of the sound wave is reflected by the
walls and little is absorbed. Walls and ceilings of concert halls are made softer materials such as
fiberglass and acoustic tiles. These materials are more similar to air than concrete and thus have
a greater ability to absorb sound. This gives the room more pleasing acoustic properties.

Reflection of sound waves off of surfaces can lead to one of two phenomena - an echo or a
reverberation. A reverberation often occurs in a small room with height, width, and length
dimensions of approximately 17 meters or less. Why the magical 17 meters? The affect of a
particular sound wave upon the brain endures for more than a tiny fraction of a second; the
human brain keeps a sound in memory for up to 0.1 seconds. If a reflected sound wave reaches
the ear within 0.1 seconds of the initial sound, then it seems to the person that the sound is
prolonged. The reception of multiple reflections off of walls and ceilings within 0.1 seconds of
each other causes reverberations - the prolonging of a sound. Since sound waves travel at about
340 m/s at room temperature, it will take approximately 0.1 s for a sound to travel the length of a
17 meter room and back, thus causing a reverberation (recall from Section 2.2, t = v/d = (340
m/s)/(34 m) = 0.1 s). This is why reverberations are common in rooms with dimensions of
approximately 17 meters or less. Perhaps you have observed reverberations when talking in an
empty room, when honking the horn while driving through a highway tunnel or underpass, or
when singing in the shower. In auditoriums and concert halls, reverberations occasionally occur
and lead to the displeasing garbling of a sound.

But reflection of sound waves in auditoriums and concert halls do not always lead to displeasing
results, especially if the reflections are designed right. Smooth walls have a tendency to direct
sound waves in a specific direction. Subsequently the use of smooth walls in an auditorium will
cause spectators to receive a large amount of sound from one location along the wall; there
would be only one possible path by which sound waves could travel from the speakers to the
listener. The auditorium would not seem to be as lively and full of sound. Rough walls tend to
diffuse sound, reflecting it in a variety of directions. This allows a spectator to perceive sounds

135
from every part of the room, making it seem lively and full. For this reason, auditorium and
concert hall designers prefer construction materials that are rough rather than smooth.

Fig. 2.29

Reflection of sound waves also leads to echoes. Echoes are different than reverberations. Echoes
occur when a reflected sound wave reaches the ear more than 0.1 seconds after the original
sound wave was heard. If the elapsed time between the arrivals of the two sound waves is more
than 0.1 seconds, then the sensation of the first sound will have died out. In this case, the arrival
of the second sound wave will be perceived as a second sound rather than the prolonging of the
first sound. There will be an echo instead of a reverberation.

Reflection of sound waves off of surfaces is also affected by the shape of the surface. As
mentioned of water waves in Chapter 1, flat or plane surfaces reflect sound waves in such a way
that the angle at which the wave approaches the surface equals the angle at which the wave
leaves the surface. This principle will be extended to the reflective behavior of light waves off of
plane surfaces in great detail in Chaoter 13. Reflection of sound waves off of curved surfaces
leads to a more interesting phenomenon. Curved surfaces with a parabolic shape have the habit
of focusing sound waves to a point. Sound waves reflecting off of parabolic surfaces concentrate
all their energy to a single point in space; at that point, the sound is amplified. Perhaps you have
seen a museum exhibit that utilizes a parabolic-shaped disk to collect a large amount of sound
and focus it at a focal point. If you place your ear at the focal point, you can hear even the
faintest whisper of a friend standing across the room. Parabolic-shaped satellite disks use this
same principle of reflection to gather large amounts of electromagnetic waves and focus it at a
point (where the receptor is located). Scientists have recently discovered some evidence that
seems to reveal that a bull moose utilizes his antlers as a satellite disk to gather and focus sound.
Finally, scientists have long believed that owls are equipped with spherical facial disks that can
be maneuvered in order to gather and reflect sound towards their ears. The reflective behavior of
light waves off curved surfaces will be studies in great detail in Book 6.

Diffraction of Sound Waves

Diffraction involves a change in direction of waves as they pass through an opening or around a
barrier in their path. The diffraction of water waves was discussed in Chapter 1. In that chapter,
we saw that water waves have the ability to travel around corners, around obstacles and through

136
openings. The amount of diffraction (the sharpness of the bending) increases with increasing
wavelength and decreases with decreasing wavelength. In fact, when the wavelength of the wave
is smaller than the obstacle or opening, no noticeable diffraction occurs.

Diffraction of sound waves is commonly observed; we notice sound


diffracting around corners or through door openings, allowing us to hear
others who are speaking to us from adjacent rooms. Many forest-dwelling
birds take advantage of the diffractive ability of long-wavelength sound
waves. Owls for instance are able to communicate across long distances due to
the fact that their long-wavelength hoots are able to diffract around forest trees
and carry farther than the short-wavelength tweets of songbirds. Low-pitched
(long wavelength) sounds always carry further than high-pitched (short
wavelength) sounds. Fig. 2.30

Scientists have recently learned that elephants emit infrasonic waves of very low frequency to
communicate over long distances to each other. Elephants typically migrate in large herds that
may sometimes become separated from each other by distances of several miles. Researchers
who have observed elephant migrations from the air and have been both impressed and puzzled
by the ability of elephants at the beginning and the end of these herds to make extremely
synchronized movements. The matriarch at the front of the herd might make a turn to the right,
which is immediately followed by elephants at the end of the herd making the same turn to the
right. These synchronized movements occur despite the fact that the elephants' vision of each
other is blocked by dense vegetation. Only recently have they learned that the synchronized
movements are preceded by infrasonic communication. While low wavelength sound waves are
unable to diffract around the dense vegetation, the high wavelength sounds produced by the
elephants have sufficient diffractive ability to communicate long distances.

Bats use high frequency (low wavelength) ultrasonic waves in order to enhance their ability to
hunt. The typical prey of a bat is the moth - an object not much larger than a couple of
centimeters. Bats use ultrasonic echolocation methods to detect the presence of bats in the air.
But why ultrasound? The answer lies in the physics of diffraction. As the wavelength of a wave
becomes smaller than the obstacle that it encounters, the wave is no longer able to diffract
around the obstacle, instead the wave reflects off the obstacle. Bats use ultrasonic waves with
wavelengths smaller than the dimensions of their prey. These sound waves will encounter the
prey, and instead of diffracting around the prey, will reflect off the prey and allow the bat to hunt
by means of echolocation. The wavelength of a 50 000 Hz sound wave in air (speed of
approximately 340 m/s) can be calculated as follows

wavelength = speed/frequency

wavelength = (340 m/s)/(50 000 Hz)

wavelength = 0.0068 m

The wavelength of the 50 000 Hz sound wave (typical for a bat) is approximately 0.7
centimeters, smaller than the dimensions of a typical moth.

137
Refraction of Sound Waves

Refraction of waves involves a change in the direction of waves as they pass from one medium
to another. Refraction, or bending of the path of the waves, is accompanied by a change in speed
and wavelength of the waves. So if the media (or its properties) are changed, the speed of the
wave is changed. Thus, waves passing from one medium to another will undergo refraction.
Refraction of sound waves is most evident in situations in which the sound wave passes through
a medium with gradually varying properties. For example, sound waves are known to refract
when traveling over water. Even though the sound wave is not exactly changing media, it is
traveling through a medium with varying properties; thus, the wave will encounter refraction and
change its direction. Since water has a moderating affect upon the temperature of air, the air
directly above the water tends to be cooler than the air far above the water. Sound waves travel
slower in cooler air than they do in warmer air. For this reason, the
portion of the wavefront directly above the water is slowed down,
while the portion of the wavefronts far above the water speeds
ahead. Subsequently, the direction of the wave changes, refracting
downwards towards the water. This is depicted in the diagram at the
right. Fig. 2.31

Refraction of other waves such as light waves will be discussed in more detail in a later chapter.

138
2.4 Resonance and Standing Waves
Natural Frequency | Forced Vibration | Standing Wave Patterns
Fundamental Frequency and Harmonics

2.4.1 Natural Frequency


As has been previously mentioned in this unit, a sound wave is created as a result of a vibrating
object. The vibrating object is the source of the disturbance that moves through the medium. The
vibrating object that creates the disturbance could be the vocal cords of a person, the vibrating
string and soundboard of a guitar or violin, the vibrating tines of a tuning fork, or the vibrating
diaphragm of a radio speaker. Any object that vibrates will create a sound. The sound could be
musical or it could be noisy; but regardless of its quality, the sound wave is created by a
vibrating object.

Nearly all objects, when hit or struck or plucked or strummed or somehow disturbed, will
vibrate. If you drop a meter stick or pencil on the floor, it will begin to vibrate. If you pluck a
guitar string, it will begin to vibrate. If you blow over the top of a pop bottle, the air inside will
vibrate. When each of these objects vibrates, they tend to vibrate at a particular frequency or a
set of frequencies. The frequency or frequencies at which an object tends to vibrate with when
hit, struck, plucked, strummed or somehow disturbed is known as the natural frequency of the
object. If the amplitudes of the vibrations are large enough and if natural frequency is within the
human frequency range, then the vibrating object will produce sound waves that are audible.

All objects have a natural frequency or set of frequencies at which they vibrate. The quality or
timbre of the sound produced by a vibrating object is dependent upon the natural frequencies of
the sound waves produced by the objects. Some objects tend to vibrate at a single frequency and
they are often said to produce a pure tone. A flute tends to vibrate at a single frequency,
producing a very pure tone. Other objects vibrate and produce more complex waves with a set of
frequencies that have a whole number mathematical relationship between them; these are said to
produce a rich sound. A tuba tends to vibrate at a set of frequencies that are mathematically
related by whole number ratios; it produces a rich tone. Still other objects will vibrate at a set of
multiple frequencies that have no simple mathematical relationship between them. These objects
are not musical at all and the sounds that they create could be described as noise. When a meter
stick or pencil is dropped on the floor, it vibrates with a number of frequencies, producing a
complex sound wave that is clanky and noisy.

139
Table 2.4

The actual frequency at which an object will vibrate at is determined by a variety of factors. Each
of these factors will either affect the wavelength or the speed of the object. Since

frequency = speed/wavelength

an alteration in either speed or wavelength will result in an alteration of the natural frequency.
The role of a musician is to control these variables in order to produce a given frequency from
the instrument that is being played. Consider a guitar as an example. There are six strings, each
having a different linear density (the wider strings are more dense on a per meter basis), a
different tension (which is controllable by the guitarist), and a different length (also controllable
by the guitarist). The speed at which waves move through the strings is dependent upon the
properties of the medium - in this case the tightness (tension) of the string and the linear density
of the strings. Changes in these properties would affect the natural frequency of the particular
string. The vibrating portion of a particular string can be shortened by pressing the string against
one of the frets on the neck of the guitar. This modification in the length of the string would
affect the wavelength of the wave and in turn the natural frequency at which a particular string
vibrates at. Controlling the speed and the wavelength in this manner allows a guitarist to control
the natural frequencies of the vibrating object (a string) and thus produce the intended musical
sounds. The same principles can be applied to any string instrument - whether it is the harp,
harpsichord, violin or guitar.

As another example, consider the trombone with its long cylindrical tube that is bent upon itself
twice and ends in a flared end. The trombone is an example of a wind instrument. The tube of
any wind instrument acts as a container for a vibrating air column. The air inside the tube will be
set into vibration by a vibrating reed or the vibrations of a musician's lips against a mouthpiece.
While the speed of sound waves within the air column is not alterable by the musician (they can
only be altered by changes in room temperature), the length of the air column is. For a trombone,
the length is altered by pushing the tube outward away from the mouthpiece to lengthen it or
pulling it in to shorten it. This causes the length of the air column to be changed, and
subsequently changes the wavelength of the waves it produces. And of course, a change in
wavelength will result in a change in the frequency. So the natural frequency of a wind

140
instrument such as the trombone is dependent upon the length of the air column of the
instrument. The same principles can be applied to any similar instrument (tuba, flute, wind
chime, organ pipe, clarinet, or pop bottle) whose sound is produced by vibrations of air within a
tube.

Fig. 2.32

There were a variety of classroom demonstrations (some of which are fun and
some of which are corny) that illustrate the idea of natural frequencies and their
modification. A pop bottle can be partly filled with water, leaving a volume of
air inside that is capable of vibrating. When a person blows over the top of the
bottle, the air inside is set into vibrational motion; turbulence above the lip of
the bottle creates disturbances within the bottle. These vibrations result in a
sound wave that is audible to students. Of course, the frequency can be modified
by altering the volume of the air column (adding or removing water), which
changes the wavelength and in turn the frequency. The principle is similar to the frequency-
wavelength relation of air columns; a smaller volume of air inside the bottle means a shorter
wavelength and a higher frequency.

A toilet paper roll orchestra can be created from different lengths of toilet paper rolls (or
wrapping paper rolls). The rolls will vibrate with different frequencies when struck against a
student's head. A properly selected set of rolls will result in the production of sounds that are
capable of a Tony Award rendition of "Mary Had a Little Lamb."

Maybe you are familiar with the popular water goblet prom trick that is often demonstrated in a
Physics class. Obtain a water goblet and clean your fingers. Then gently
slide your finger over the rim of the water goblet. If you are fortunate
enough, you might be able to set the goblet into vibration by means of
slip-stick friction. (It is not necessary to use a crystal goblet. It is often
said that crystal goblets work better; but the trick is just as easily
performed with clean fingers and an inexpensive goblet.) Like a violin
bowstring being pulled across a violin string, the finger sticks to the glass
molecules, pulling them apart at a given point until the tension becomes so
great. The finger then slips off the glass and subsequently finds another microscopic surface to
stick to; the finger pulls the molecules at that surface, slips and then sticks at another location.
This process of stick-slip friction occurring at a high frequency is sufficient to set the molecules
of the glass into vibration at its natural frequency. The result is enough to impress your dinner
guests. Try it at home!!

141
Perhaps you have seen a pendulum bob vibrating back and forth about its equilibrium position.
While a pendulum does not produce a sound when it oscillates, it does illustrate an important
principle. A pendulum consisting of a longer string vibrates with a longer period and thus a
lower frequency. Once more, there is an inverse relationship between the length of the vibrating
object and the natural frequency at which the object vibrates. This very relationship carries over
to any vibrating instrument - whether it is a guitar string, a xylophone, a pop bottle instrument, or
a kettledrum.

To conclude, all objects have a natural frequency or set of frequencies at which they vibrate
when struck, plucked, strummed or somehow disturbed. The actual frequency is dependent upon
the properties of the material the object is made of (this affects the speed of the wave) and the
length of the material (this effects the wavelength of the wave). It is the goal of musicians to find
instruments that possess the ability to vibrate with sets of frequencies that are musically sounding
(i.e., mathematically related by simple whole number ratios) and to vary the lengths and (if
possible) properties to create the desired sounds.

Watch It!

A physics instructor makes a water goblet sing at its natural frequency.

2.4.2 Forced Vibration


Musical instruments and other objects are set into vibration at their natural frequency when a
person hits, strikes, strums, plucks or somehow disturbs the object. For instance, a guitar string is
strummed or plucked; a piano string is hit with a hammer when a pedal is played; and the tines of
a tuning fork are hit with a rubber mallet. Whatever the case, a person or thing puts energy into
the instrument by direct contact with it. This input of energy disturbs the particles and forces the
object into vibrational motion - at its natural frequency.

If you were to take a guitar string and stretch it to a given length and a given tightness and have a
friend pluck it, you would hear a noise; but the noise would not even be close in comparison to
the loudness produced by an acoustic guitar. On the other hand, if the string is attached to the
sound box of the guitar, the vibrating string is capable of forcing the sound box into vibrating at

142
that same natural frequency. The sound box in turn forces air particles inside the box into
vibrational motion at the same natural frequency as the string. The entire system (string, guitar,
and enclosed air) begins vibrating and forces surrounding air particles into vibrational motion.
The tendency of one object to force another adjoining or interconnected object into vibrational
motion is referred to as a forced vibration. In the case of the guitar string mounted to the sound
box, the fact that the surface area of the sound box is greater than the surface area of the string
means that more surrounding air particles will be forced into vibration. This causes an increase in
the amplitude and thus loudness of the sound.

This same principle of a forced vibration is often demonstrated in a Physics


classroom using a tuning fork. If the tuning fork is held in your hand and
hit with a rubber mallet, a sound is produced as the tines of the tuning fork
set surrounding air particles into vibrational motion. The sound produced
by the tuning fork is barely audible to students in the back rows of the
room. However, if the tuning fork is set upon the whiteboard panel or the
glass panel of the overhead projector, the panel begins vibrating at the
same natural frequency of the tuning fork. The tuning fork forces
surrounding glass (or vinyl) particles into vibrational motion. The vibrating
whiteboard or overhead projector panel in turn forces surrounding air Fig. 2.33
particles into vibrational motion and the result is an increase in the amplitude and thus loudness
of the sound. This principle of forced vibration explains why demonstration tuning forks are
mounted on a sound box, why a commercial music box mechanism is mounted on a sounding
board, why a guitar utilizes a sound box, and why a piano string is attached to a sounding board.
A louder sound is always produced when an accompanying object of greater surface area is
forced into vibration at the same natural frequency.

Now consider a related situation that resembles another common Physics demonstration.
Suppose that a tuning fork is mounted on a sound box and set
upon the table; and suppose a second tuning fork/sound box
system having the same natural frequency (say 256 Hz) is
placed on the table near the first system. Neither of the tuning
forks is vibrating. Suppose the first tuning fork is struck with a
rubber mallet and the tines begin vibrating at its natural
frequency - 256 Hz. These vibrations set its sound box and the
air inside the sound box vibrating at the same natural
Fig. 2.34
frequency of 256 Hz. Surrounding air particles are set into
vibrational motion at the same natural frequency of 256 Hz
and every student in the classroom hears the sound. Then the
tines of the tuning fork are grabbed to prevent their vibration and remarkably the sound of 256
Hz is still being heard. Only now the sound is being produced by the second tuning fork - the one
which wasn't hit with the mallet. Amazing!! The demonstration is often repeated to assure that
the same surprising results are observed. They are! What is happening?

In this demonstration, one tuning fork forces another tuning fork into vibrational motion at the
same natural frequency. The two forks are connected by the surrounding air particles. As the air
particles surrounding the first fork (and its connected sound box) begin vibrating, the pressure

143
waves that it creates begin to impinge at a periodic and regular rate of 256 Hz upon the second
tuning fork (and its connected sound box). The energy carried by this sound wave through the air
is tuned to the frequency of the second tuning fork. Since the incoming sound waves share the
same natural frequency as the second tuning fork, the tuning fork easily begins vibrating at its
natural frequency. This is an example of resonance - when one object vibrating at the same
natural frequency of a second object forces that second object into vibrational motion.

The result of resonance is always a large vibration. Regardless of the vibrating system, if
resonance occurs, a large vibration results. This is often demonstrated in a Physics class with an
odd-looking mechanical system resembling an inverted pendulum. The apparatus consists of
three sets of two identical plastic bobs mounted on a very
elastic metal pole, which are in turn mounted to a metal bar.
Each metal pole and attached bob has a different length, thus
giving it a different natural frequency of vibration. The bobs
are often color-coded to distinguish between them; they are
colored red, blue and green (a set of three colors that will be
significant later. The red bobs are mounted on the longer poles
and they have the lowest natural frequency of vibration. The
blue bobs are mounted on the shorter poles and have the
highest natural frequency of vibration. (Note the length-
wavelength-frequency relationship that was discussed earlier.) Fig. 2.35
When the red bob is disturbed, it begins vibrating at its natural frequency. This in turn forces the
attached bar to vibrate at the same frequency; and this forces the other attached red bob into
vibrating at the same natural frequency. This is resonance - one bob vibrating at a given
frequency forcing a second object with the same natural frequency into vibrational motion.
While the green and the blue bobs were disturbed by the vibrations transmitted through the metal
bar, only the red bob would resonate. This is because the frequency of the first red bob is tuned
to the frequency of the second red bob; they share the same natural frequency. The result is that
the second red bob begins vibrating with a huge amplitude.

Watch It!

Another common classroom demonstration of resonance involves


a plastic tube containing an air column. The length of the air
column was adjusted by raising and lowering a reservoir of water
(dyed red). The raising and lowering of the reservoir adjusts the
height of water in the open-air tube, and thus adjusts the length of
the air column inside the tube. As the length of the air column is
decreased, the natural frequency of the air column is increased.
(Again note the length-wavelength-frequency relationship that
Fig. 2.36
144
was discussed earlier.) While adjusting the height of the liquid in the tube, a vibrating tuning fork
is held above the air column of the tube. When the natural frequency of the air column is tuned to
the frequency of the vibrating tuning fork, resonance occurs and a loud sound results. Quite
amazingly, the vibrating tuning fork forces air particles within the air column into vibrational
motion. Once more in this resonance situation, the tuning fork and the air column share the same
vibrational frequency.

In conclusion, resonance occurs when two interconnected objects share the same vibrational
frequency. When one of the objects is vibrating, it forces the second object into vibrational
motion. The result is a large vibration. And if a sound wave within the audible range of human
hearing is produced, a loud sound is heard.

2.4.3 Standing Wave Patterns


As mentioned earlier, all objects have a frequency or set of frequencies with which they naturally
vibrate when struck, plucked, strummed or somehow disturbed. Each of the natural frequencies
at which an object vibrates is associated with a standing wave pattern. When an object is forced
into resonance vibrations at one of its natural frequencies, it vibrates in a manner such that a
standing wave is formed within the object. The topic of standing wave patterns was introduced in
Chapter 10. In that unit, a standing wave pattern was described as a vibrational pattern created
within a medium when the vibrational frequency of a source causes reflected waves from one
end of the medium to interfere with incident waves from the source, The result of the
interference is that specific points along the medium appear to be standing still while other points
vibrated back and forth. Such patterns are only created within the medium at specific frequencies
of vibration. These frequencies are known as harmonic frequencies or merely harmonics. At any
frequency other than a harmonic frequency, the interference of reflected and incident waves
results in a disturbance of the medium that is irregular and non-repeating.

So the natural frequencies of an object are merely the harmonic frequencies at which standing
wave patterns are established within the object. These standing wave patterns represent the
lowest energy vibrational modes of the object. While there are countless ways by which an
object can vibrate (each associated with a specific frequency), objects favor only a few specific
modes or patterns of vibrating. The favored modes (patterns) of vibration are those that result in
the highest amplitude vibrations with the least input of energy. Objects favor these natural modes
of vibration because they are representative of the patterns that require the least amount of
energy. Objects are most easily forced into resonance vibrations when disturbed at frequencies
associated with these natural frequencies.

145
The wave pattern associated with the natural frequencies of an object is characterized by points
that appear to be standing still. For this reason, the pattern is often called a "standing wave
pattern." The points in the pattern that are standing still are referred to as nodal points or nodal
positions. These positions occur as the result of the destructive interference of incident and
reflected waves. Each nodal point is surrounded by antinodal points, creating an alternating
pattern of nodal and antinodal points. Such patterns were introduced in Chapter 1. In this unit,
we will elaborate on the essential characteristics and the causes of standing wave patterns and
relate these patterns to the vibrations of musical instruments.

A common Physics demonstration utilizes a square metal plate (known as a Chladni plate), a
violin bow and salt. The plate is securely fastened to a table using a nut and bolt. The nut and
bolt are clamped to the center of the square plate, preventing that section from
vibrating. Salt (or sand) is sprinkled upon the plate in an irregular pattern. Then the
violin bow is used to induce vibrations within the plate; the plate is strummed and
begins vibrating. And then the magic occurs. A high-pitched pure tone is sounded out
as the plate vibrates. And, remarkably (as is often the case in a physics class), the salt
upon the plate begins vibrating and forms a pattern upon the plate. As we know, all
objects (even a silly little metal plate) have a set of natural frequencies at which they
vibrate; and each frequency is associated with a standing wave pattern. The pattern
formed by the salt on the plate is the standing wave pattern associated with one of the
natural frequencies of the Chladni plate. As the plate vibrates, the salt begins to
vibrate and tumble about the plate until it reaches points along the plate that are not vibrating.
Subsequently, the salt finally comes to rest along the nodal positions. The diagrams at the right
show two of the most common standing wave patterns for the Chladni plates. The white lines
represent the salt locations (nodal positions). Observe in the diagram that each pattern is
characterized by nodal positions in the corners of the square plate and in the center of the plate.
For these two particular vibrational modes, those positions are unable to move. Being unable to
move, they become nodal points - points of no displacement.

Fig. 2.37

Flickr Physics Photo


Salt is sprinkled onto a metal plate. The plate is strummed with a violin bow and set into
vibration. The salt crystals vibrate about the plate until they settle onto positions of nodes (points
of no displacement). Several patterns can be obtained, each associated with a unique frequency

146
of vibrations. These standing wave patterns are known as Chladni patterns, named in honor of a
19th century German physicist who advanced our understanding of acoustics and the physics of
music.

*
Fig. 2.38

Standing Wave Patterns for Vibrating Strings

The diagram below depicts one of the natural patterns of vibrations for a guitar string. In the
pattern, you will note that there are certain positions along the string (the medium) that appear to
be standing still. These positions are referred to as nodes and are labeled on the diagram. In
between each nodal position, there are other positions that appear to be vibrating back and forth
between a large upward displacement to a large downward displacement. These points are
referred to as antinodes and are also labeled on the diagram. There is an alternating pattern of
nodal and antinodal positions in a standing wave pattern.

147
Fig. 2.39

Because the antinodal positions along the guitar string are vibrating back and forth from a large
upward displacement to a large downward displacement, the standing wave pattern is often
depicted by a diagram such as that shown below.

Fig. 2.40

The pattern above is not the only pattern of vibration for a guitar string.
There are a variety of patterns by which the guitar string could naturally
vibrate. Each pattern is associated with one of the natural frequencies of
the guitar strings. Three other patterns are shown in the diagrams at the
right. Each standing wave pattern is referred to as a harmonic of the
instrument (in this case, the guitar string). The three diagrams at the right
represent the standing wave patterns for the first, second, and third
harmonics of a guitar string. (Harmonics will be discussed in more detail
in the next sub-section of this Section.) Fig. 2.41
There are a variety of other low energy vibrational patterns that could be
established in the string. For guitar strings, each pattern is characterized by some basic traits:

• There is an alternating pattern of nodes and antinodes.


• There are either a half-number or a whole number of waves within the pattern established
on the string.
• Nodal positions (points of no displacement) are established at the ends of the string where
the string is clamped down in a fixed position.
• One pattern is related to the next pattern by the addition (or subtraction) of one or more
nodes (and antinodes).

The standing wave patterns for other musical instruments share some these same traits or at least
similar traits. These patterns will be discussed in more detail in Section 2.5.

148
2.4.4 Fundamental Frequency and Harmonics
Previously in this Section 2.4, it was mentioned that when an object is forced into resonance
vibrations at one of its natural frequencies, it vibrates in a manner such that a standing wave
pattern is formed within the object. Whether it is a guitar sting, a Chladni plate, or the air column
enclosed within a trombone, the vibrating medium vibrates in such a way that a standing wave
pattern results. Each natural frequency that an object or instrument produces has its own
characteristic vibrational mode or standing wave pattern. These patterns are only created within
the object or instrument at specific frequencies of vibration; these frequencies are known as
harmonic frequencies, or merely harmonics. At any frequency other than a harmonic frequency,
the resulting disturbance of the medium is irregular and non-repeating. For musical instruments
and other objects that vibrate in regular and periodic fashion, the harmonic frequencies are
related to each other by simple whole number ratios. This is part of the reason why such
instruments sound pleasant. We will see in this sub-section of Section 2.4 why these whole
number ratios exist for a musical instrument.

First, consider a guitar string vibrating at its natural frequency or harmonic frequency. Because
the ends of the string are attached and fixed in place to the guitar's structure (the bridge at one
end and the frets at the other), the ends of the string are unable to move. Subsequently, these
ends become nodes - points of no displacement. In between these two nodes at the end of the
string, there must be at least one antinode. The most fundamental harmonic for a guitar string is
the harmonic associated with a standing wave having only one antinode positioned between the
two nodes on the end of the string. This would be the harmonic with the
longest wavelength and the lowest frequency. The lowest frequency
produced by any particular instrument is known as the fundamental
frequency. The fundamental frequency is also called the first
harmonic of the instrument. The diagram at the right shows the first
harmonic of a guitar string. If you analyze the wave pattern in the guitar
string for this harmonic, you will notice that there is not quite one
complete wave within the pattern. A complete wave starts at the rest Fig. 2.42
position, rises to a crest, returns to rest, drops to a trough, and finally returns to the rest position
before starting its next cycle. (Caution: the use of the words crest and trough to describe the
pattern are only used to help identify the length of a repeating wave cycle. A standing wave
pattern is not actually a wave, but rather a pattern of a wave. Thus, it does not consist of crests
and troughs, but rather nodes and antinodes. The pattern is the result of the interference of two
waves to produce these nodes and antinodes.) In this pattern, there is only one-half of a wave
within the length of the string. This is the case for the first harmonic or fundamental frequency of
a guitar string. The diagram below depicts this length-wavelength relationship for the
fundamental frequency of a guitar string.

149
Fig. 2.43

The second harmonic of a guitar string is produced by adding one more


node between the ends of the guitar string. And of course, if a node is
added to the pattern, then an antinode must be added as well in order to
maintain an alternating pattern of nodes and antinodes. In order to create
a regular and repeating pattern, that node must be located midway Fig. 2.44
between the ends of the guitar string. This additional node gives the second harmonic a total of
three nodes and two antinodes. The standing wave pattern for the second harmonic is shown at
the right. A careful investigation of the pattern reveals that there is exactly one full wave within
the length of the guitar string. For this reason, the length of the string is equal to the length of the
wave.

The third harmonic of a guitar string is produced by adding two nodes


between the ends of the guitar string. And of course, if two nodes are
added to the pattern, then two antinodes must be added as well in order to
maintain an alternating pattern of nodes and antinodes. In order to create a
regular and repeating pattern for this harmonic, the two additional nodes
must be evenly spaced between the ends of the guitar string. This places Fig. 2.45
them at the one-third mark and the two-thirds mark along the string. These additional nodes give
the third harmonic a total of four nodes and three antinodes. The standing wave pattern for the
third harmonic is shown at the right. A careful investigation of the pattern reveals that there is
more than one full wave within the length of the guitar string. In fact, there are three-halves of a
wave within the length of the guitar string. For this reason, the length of the string is equal to
three-halves the length of the wave. The diagram below depicts this length-wavelength
relationship for the fundamental frequency of a guitar string.

Fig. 2.46

After a discussion of the first three harmonics, a pattern can be recognized. Each harmonic
results in an additional node and antinode, and an additional half of a wave within the string. If

150
the number of waves in a string is known, then an equation relating the wavelength of the
standing wave pattern to the length of the string can be algebraically derived.

Fig. 2.47

This information is summarized in the table below.

Table 2.5
# of # of Length-
Harm. # of
Waves Anti- Wavelength
# Nodes
in String nodes Relationship
1 1/2 2 1 Wavelength = (2/1)*L
2 1 or 2/2 3 2 Wavelength = (2/2)*L
3 3/2 4 3 Wavelength = (2/3)*L
4 2 or 4/2 5 4 Wavelength = (2/4)*L
5 5/2 6 5 Wavelength = (2/5)*L

The above discussion develops the mathematical relationship between the length of a guitar
string and the wavelength of the standing wave patterns for the various harmonics that could be
established within the string. Now these length-wavelength relationships will be used to develop
relationships for the ratio of the wavelengths and the ratio of the frequencies for the various
harmonics played by a string instrument (such as a guitar string).

Determining the Harmonic Frequencies

Consider an 80-cm long guitar string that has a fundamental frequency (1st harmonic) of 400 Hz.
For the first harmonic, the wavelength of the wave pattern would be two times the length of the
string (see table above); thus, the wavelength is 160 cm or 1.60 m. The speed of the standing
wave can now be determined from the wavelength and the frequency. The speed of the standing
wave is

151
speed = frequency • wavelength

speed = 400 Hz • 1.6 m

speed = 640 m/s

This speed of 640 m/s corresponds to the speed of any wave within the guitar string. Since the
speed of a wave is dependent upon the properties of the medium (and not upon the properties of
the wave), every wave will have the same speed in this string regardless of its frequency and its
wavelength. So the standing wave pattern associated with the second harmonic, third harmonic,
fourth harmonic, etc. will also have this speed of 640 m/s. A change in frequency or wavelength
will NOT cause a change in speed.

Using the table above, the wavelength of the second harmonic (denoted by the symbol 2 ) would
be 0.8 m (the same as the length of the string). The speed of the standing wave pattern (denoted
by the symbol v) is still 640 m/s. Now the wave equation can be used to determine the frequency
of the second harmonic (denoted by the symbol f2 ).

speed = frequency • wavelength

frequency = speed/wavelength

f2 = v / 2

f2 = (640 m/s)/(0.8 m)

f2 = 800 Hz

This same process can be repeated for the third harmonic. Using the table above, the wavelength
of the third harmonic (denoted by the symbol 3 ) would be 0.533 m (two-thirds of the length of
the string). The speed of the standing wave pattern (denoted by the symbol v) is still 640 m/s.
Now the wave equation can be used to determine the frequency of the third harmonic (denoted
by the symbol f3 ).

speed = frequency • wavelength

frequency = speed/wavelength

f3 = v / 3

f3 = (640 m/s)/(0.533 m)

f3 = 1200 Hz

152
Now if you have been following along, you will have recognized a pattern. The frequency of the
second harmonic is two times the frequency of the first harmonic. The frequency of the third
harmonic is three times the frequency of the first harmonic. The frequency of the nth harmonic
(where n represents the harmonic # of any of the harmonics) is n times the frequency of the first
harmonic. In equation form, this can be written as

fn = n • f1
The inverse of this pattern exists for the wavelength values of the various harmonics. The
wavelength of the second harmonic is one-half (1/2) the wavelength of the first harmonic. The
wavelength of the third harmonic is one-third (1/3) the wavelength of the first harmonic. And the
wavelength of the nth harmonic is one-nth (1/n) the wavelength of the first harmonic. In equation
form, this can be written as

n = (1/n) • 1

These relationships between wavelengths and frequencies of the various harmonics for a guitar
string are summarized in the table below.

Table 2.6
Harm. Freq. Wavelength Speed
fn / f1 n / 1
# (Hz) (m) (m/s)
1 400 1.60 640 1 1/1
2 800 0.800 640 2 ½
3 1200 0.533 640 3 1/3
4 1600 0.400 640 4 ¼
5 2000 0.320 640 5 1/5
n n * 400 (2/n)*(0.800) 640 n 1/n

The table above demonstrates that the individual frequencies in the set of natural frequencies
produced by a guitar string are related to each other by whole number ratios. For instance, the
first and second harmonics have a 2:1 frequency ratio; the second and the third harmonics have a
3:2 frequency ratio; the third and the fourth harmonics have a 4:3 frequency ratio; and the fifth
and the fourth harmonic have a 5:4 frequency ratio. When the guitar is played, the string, sound
box and surrounding air vibrate at a set of frequencies to produce a wave with a mixture of
harmonics. The exact composition of that mixture determines the timbre or quality of sound that

153
is heard. If there is only a single harmonic sounding out in the mixture (in which case, it wouldn't
be a mixture), then the sound is rather pure-sounding. On the other hand, if there are a variety of
frequencies sounding out in the mixture, then the timbre of the sound is rather rich in quality.

In Section 2.5, these same principles of resonance and standing waves will be applied to other
types of instruments besides guitar strings.

Exercise 2.8
1. Anna Litical cuts short sections of PVC pipe into different
lengths and mounts them in putty on the table. The PVC pipes
form closed-end air columns that sound out at different
frequencies when she blows over the top of them. The actual
frequency of vibration is inversely proportional to the wavelength
of the sound; and thus, the frequency of vibration is inversely
proportional to the length of air inside the tubes. Express your
understanding of this resonance phenomenon by filling in the
following table.
Table Q. 2.8.1

2. In a rare moment of artistic brilliance, a Physics teacher pulls out his violin bow and
strokes a square metal plate to produce vibrations within the plate. Often times, he
places salt upon the plate and observes the standing wave patterns established in the
plate as it vibrates. Amazingly, the salt is aligned along the locations of the plate that
are not vibrating and far from the locations of maximum vibration. The two most
common standing wave patterns are illustrated at the right. Compare the wavelength
of pattern A to the wavelength of pattern B. Suppose that the fundamental frequency
of vibration is nearly 1200 Hz. Estimate the frequency of vibration of the plate when it
vibrates in the second, third and fourth harmonics.

3. When a tennis racket strikes a tennis ball, the racket begins to vibrate. There is a set of
selected frequencies at which the racket will tend to vibrate. Each frequency in the set is
characterized by a particular standing wave pattern. The diagrams below show the three of the
more common standing wave patterns for the vibrations of a tennis racket.

154
Fig.Q. 2.8.3
a. Compare the wavelength of pattern A to the wavelength of pattern B. Make your comparison
both qualitative and quantitative. Repeat for pattern C.

b. Compare the frequency of pattern A to the frequency of pattern B. Make your comparison
both qualitative and quantitative. Repeat for pattern C.

c. When the racket vibrates as in pattern A, its frequency of vibration is approximately 30 Hz.
Determine the frequency of vibration of the racket when it vibrates as in pattern B and pattern C.

155
2.5 Musical Instruments
Resonance | Guitar Strings | Open-End Air Columns | Closed-End Air Columns

2.5.1 Resonance
The goal of Chapter 2 is to develop an understanding of the nature, properties, behavior, and
mathematics of sound and to apply this understanding to the analysis of music and musical
instruments. Thus far in this unit, applications of sound wave principles have been made towards
a discussion of beats, musical intervals, concert hall acoustics, the distinctions between noise
and music, and sound production by musical instruments. In this Section 2.5, the focus will be
upon the application of mathematical relationships and standing wave concepts to musical
instruments. Three general categories of instruments will be investigated: instruments with
vibrating strings (which would include guitar strings, violin strings, and piano strings), open-end
air column instruments (which would include the brass instruments such as the trombone and
woodwinds such as the flute and the recorder), and closed-end air column instruments (which
would include some organ pipe and the bottles of a pop bottle orchestra). A fourth category -
vibrating mechanical systems (which includes all the percussion instruments) - will not be
discussed. These instrument categories may be unusual to some; they are based upon the
commonalities among their standing wave patterns and the mathematical relationships between
the frequencies that the instruments produce.

As was mentioned in Section 2.4, musical instruments are set into vibrational motion at their
natural frequency when a person hits, strikes, strums, plucks or somehow disturbs the object.
Each natural frequency of the object is associated with one of the many standing wave patterns
by which that object could vibrate. The natural frequencies of a musical instrument are
sometimes referred to as the harmonics of the instrument. An instrument can be forced into
vibrating at one of its harmonics (with one of its standing wave patterns) if another
interconnected object pushes it with one of those frequencies. This is known as resonance -
when one object vibrating at the same natural frequency of a second object forces that second
object into vibrational motion.

The word resonance comes from Latin and means to "resound" -


to sound out together with a loud sound. Resonance is a
common cause of sound production in musical instruments. One
of our best models of resonance in a musical instrument is a
resonance tube (a hollow cylindrical tube) partially filled with
water and forced into vibration by a tuning fork. The tuning fork
is the object that forced the air inside of the resonance tube into
resonance. As the tines of the tuning fork vibrate at their own
natural frequency, they created sound waves that impinge upon
the opening of the resonance tube. These impinging sound
waves produced by the tuning fork force air inside of the

Fig. 2.48
156
resonance tube to vibrate at the same frequency. Yet, in the absence of resonance, the sound of
these vibrations is not loud enough to discern. Resonance only occurs when the first object is
vibrating at the natural frequency of the second object. So if the frequency at which the tuning
fork vibrates is not identical to one of the natural frequencies of the air column inside the
resonance tube, resonance will not occur and the two objects will not sound out together with a
loud sound. But the location of the water level can be altered by raising and lowering a reservoir
of water, thus decreasing or increasing the length of the air column. As we have learned earlier,
an increase in the length of a vibrational system (here, the air in the tube) increases the
wavelength and decreases the natural frequency of that system. Conversely, a decrease in the
length of a vibrational system decreases the wavelength and increases the natural frequency. So
by raising and lowering the water level, the natural frequency of the air in the tube could be
matched to the frequency at which the tuning fork vibrates. When the match is achieved, the
tuning fork forces the air column inside of the resonance tube to vibrate at its own natural
frequency and resonance is achieved. The result of resonance is always a big vibration - that is, a
loud sound.

Another common physics demonstration that serves as an excellent model of resonance is the
famous "singing rod" demonstration. A long hollow aluminum rod is held at its center. Being a
trained musician, teacher reaches in a rosin bag to prepare for the event. Then with great
enthusiasm, he/she slowly slides her hand across the length of the aluminum rod, causing it to
sound out with a loud sound. This is an example of resonance. As the hand slides across the
surface of the aluminum rod, slip-stick friction between the hand and the rod produces vibrations
of the aluminum. The vibrations of the aluminum force the air column inside of the rod to vibrate
at its natural frequency. The match between the vibrations of the air column and one of the
natural frequencies of the singing rod causes resonance. The result of resonance is always a big
vibration - that is, a loud sound.

Fig. 2.49

The familiar sound of the sea that is heard when a seashell is placed up to your ear is also
explained by resonance. Even in an apparently quiet room, there are sound waves with a range of
frequencies. These sounds are mostly inaudible due to their low intensity. This so-called
background noise fills the seashell, causing vibrations within the seashell. But the seashell has a
set of natural frequencies at which it will vibrate. If one of the frequencies in the room forces air
within the seashell to vibrate at its natural frequency, a resonance situation is created. And
always, the result of resonance is a big vibration - that is, a loud sound. In fact, the sound is loud

157
enough to hear. So the next time you hear the sound of the sea in a seashell, remember that all
that you are hearing is the amplification of one of the many background frequencies in the room.

Musical instruments produce their selected sounds in the same


manner. Brass instruments typically consist of a mouthpiece
attached to a long tube filled with air. The tube is often curled in
order to reduce the size of the instrument. The metal tube merely
serves as a container for a column of air. It is the vibrations of this
column that produces the sounds that we hear. The length of the
vibrating air column inside the tube can be adjusted either by
sliding the tube to increase and decrease its length or by opening and closing holes located along
the tube in order to control where the air enters and exits the tube. Brass instruments involve the
blowing of air into a mouthpiece. The vibrations of the lips against the mouthpiece produce a
range of frequencies. One of the frequencies in the range of frequencies matches one of the
natural frequencies of the air column inside of the brass instrument. This forces the air inside of
the column into resonance vibrations. The result of resonance is always a big vibration - that is, a
loud sound.

Woodwind instruments operate in a similar manner. Only, the source of vibrations is not the lips
of the musician against a mouthpiece, but rather the vibration of a reed
or wooden strip. The operation of a woodwind instrument is often
modeled in a Physics class using a plastic straw. The ends of the straw
are cut with a scissors, forming a tapered reed. When air is blown
through the reed, the reed vibrates producing turbulence with a range of
vibrational frequencies. When the frequency of vibration of the reed
matches the frequency of vibration of the air column in the straw, resonance occurs. And once
more, the result of resonance is a big vibration - the reed and air column sound out together to
produce a loud sound. As if this weren't silly enough, the length of the straw is typically
shortened by cutting small pieces off its opposite end. As the straw (and the air column that it
contained) is shortened, the wavelength decreases and the frequency was increases. Higher and
higher pitches are observed as the straw is shortened. Woodwind instruments produce their
sounds in a manner similar to the straw demonstration. A vibrating reed forces an air column to
vibrate at one of its natural frequencies. Only for wind instruments, the length of the air column
is controlled by opening and closing holes within the metal tube (since the tubes are a little
difficult to cut and a too expensive to replace every time they are cut).

Resonance is the cause of sound production in musical instruments. In the remainder of Section
11.5, the mathematics of standing waves will be applied to understanding how resonating strings
and air columns produce their specific frequencies.

158
2.5.2 Guitar Strings
A guitar string has a number of frequencies at which it will naturally vibrate. These natural
frequencies are known as the harmonics of the guitar string. As mentioned earlier, the natural
frequency at which an object vibrates at depends upon the tension of the string, the linear density
of the string and the length of the string. Each of these natural frequencies or harmonics is
associated with a standing wave pattern. The specifics of the patterns and their formation were
discussed in Section 2.4. For now, we will merely summarize the results of that discussion. The
graphic below depicts the standing wave patterns for the lowest three harmonics or frequencies
of a guitar string.

Fig. 2.50

The wavelength of the standing wave for any given harmonic is related to the length of the string
(and vice versa). If the length of a guitar string is known, the wavelength associated with each of
the harmonic frequencies can be found. Thus, the length-wavelength relationships and the wave
equation (speed = frequency * wavelength) can be combined to perform calculations predicting
the length of string required to produce a given natural frequency. And conversely, calculations
can be performed to predict the natural frequencies produced by a known length of string. Each
of these calculations requires knowledge of the speed of a wave in a string. The graphic below
depicts the relationships between the key variables in such calculations. These relationships will
be used to assist in the solution to problems involving standing waves in musical instruments.

159
Fig. 2.51

To demonstrate the use of the above problem-solving scheme, consider the following problem
and its detailed solution.

Example Problem
The speed of waves in a particular guitar string is 425 m/s. Determine the fundamental frequency
(1st harmonic) of the string if its length is 76.5 cm.

The solution to the problem begins by first identifying known information, listing the desired
quantity, and constructing a diagram of the situation.

Given: Find: Diagram:

v = 425 m/s f1 = ??

L = 76.5 cm = 0.765 m

The problem statement asks us to determine the frequency (f) value. From the graphic above, the
only means of finding the frequency is to use the wave equation (speed=frequency • wavelength)
and knowledge of the speed and wavelength. The speed is given, but wavelength is not known. If
the wavelength could be found, then the frequency could be easily calculated. In this problem
(and any problem), knowledge of the length and the harmonic number allows one to determine
the wavelength of the wave. For the first harmonic, the wavelength is twice the length. This
relationship is derived from the diagram of the standing wave pattern (and was explained in
detail in Section 2.4). This relationship, which works only for the first harmonic of a guitar
string, is used to calculate the wavelength for this standing wave.

160
Wavelength = 2 • Length

Wavelength = 2 • 0.765 m

Wavelength = 1.53 m

Now that wavelength is known, it can be combined with the given value of the speed to calculate
the frequency of the first harmonic for this given string. This calculation is shown below.

speed = frequency • wavelength

frequency = speed / wavelength

frequency = (425 m/s) / (1.53 m)

frequency = 278 Hz

Most problems can be solved in a similar manner. It is always wise to take the extra time needed
to set the problem up; take the time to write down the given information and the requested
information and to draw a meaningful diagram. These preparatory steps become more important
as problems become more difficult.

Seldom in physics are two problems identical. The tendency to treat every problem the same way
is perhaps one of the quickest paths to failure. It is important to combine good problem-solving
skills (part of which involves the discipline to set the problem up) with a solid grasp of the
relationships among variables. Avoid the tendency to memorize approaches to different types of
problems. To further your understanding of these relationships and the use of the above problem-
solving scheme, examine the following problem and its solution.

Example Problem
Determine the length of guitar string required to produce a fundamental frequency (1st
harmonic) of 256 Hz. The speed of waves in a particular guitar string is known to be 405 m/s.

The solution to the problem begins by first identifying known information, listing the desired
quantity, and constructing a diagram of the situation.

Given: Find: Diagram:

v = 405 m/s L = ??

f1 = 256 Hz

161
The problem statement asks us to determine the length of the guitar string. From the graphic
above, the only means of finding the length of the string is from knowledge of the wavelength.
But the wavelength is not known. However, the frequency and speed are given, so one can use
the wave equation (speed = frequency • wavelength) and knowledge of the speed and frequency
to determine the wavelength. This calculation is shown below.

speed = frequency • wavelength

wavelength = speed / frequency

wavelength = (405 m/s) / (256 Hz)

wavelength = 1.58 m

Now that the wavelength is found, the length of the guitar string can be calculated. For the first
harmonic, the length is one-half the wavelength. This relationship is derived from the diagram of
the standing wave pattern (and was explained in detail in Section 11.4). It may also be evident to
you by looking at the standing wave diagram drawn above. This relationship between
wavelength and length, which works only for the first harmonic of a guitar string, is used to
calculate the wavelength for this standing wave pattern.

Length = (1/2) • Wavelength

Length = (1/2) • Wavelength

Length = 0.791 m

If you have successfully followed the logic in the above two example problems, take a try at the
following practice problems. As you proceed, be sure to be mindful of the numerical
relationships involved in such problems. And if necessary, refer to the graphic above.

Exercise 2.9
1. A guitar string with a length of 80.0 cm is plucked. The speed of a wave in the string is 400
m/sec. Calculate the frequency of the first, second, and third harmonics.

2. A pitch of Middle D (first harmonic = 294 Hz) is sounded out by a vibrating guitar string. The
length of the string is 70.0 cm. Calculate the speed of the standing wave in the guitar string.

3. A frequency of the first harmonic is 587 Hz (pitch of D5) is sounded out by a vibrating guitar
string. The speed of the wave is 600 m/sec. Find the length of the string.

162
2.5.3 Open-End Air Columns
Many musical instruments consist of an air column enclosed inside of a hollow metal tube.
Though the metal tube may be more than a meter in length, it is often curved upon itself one or
more times in order to conserve space. If the end of the tube is uncovered such that the air at the
end of the tube can freely vibrate when the sound wave reaches it, then the end is referred to as
an open end. If both ends of the tube are uncovered or open, the musical instrument is said to
contain an open-end air column. A variety of instruments operate on the basis of open-end air
columns; examples include the flute and the recorder. Even some organ pipes serve as open-end
air columns.

Standing Wave Patterns for the Harmonics

As has already been mentioned, a musical instrument has a set of natural frequencies at which it
vibrates at when a disturbance is introduced into it. These natural frequencies are known as the
harmonics of the instrument; each harmonic is associated with a standing wave pattern. In
Section 1.4 of Chapter 1, a standing wave pattern was defined as a vibrational pattern created
within a medium when the vibrational frequency of the source causes reflected waves from one
end of the medium to interfere with incident waves from the source in such a manner that
specific points along the medium appear to be standing still. In the case of stringed instruments
(discussed earlier), standing wave patterns were drawn to depict the amount of movement of the
string at various locations along its length. Such patterns show nodes - points of no displacement
or movement - at the two fixed ends of the string. In the case of air columns, a closed end in a
column of air is analogous to the fixed end on a vibrating string. That is, at the closed end of an
air column, air is not free to undergo movement and thus is forced into assuming the nodal
positions of the standing wave pattern. Conversely, air is free to undergo its back-and-forth
longitudinal motion at the open end of an air column; and as such, the standing wave patterns
will depict antinodes at the open ends of air columns.

So the basis for drawing the standing wave patterns for air columns is that vibrational antinodes
will be present at any open end and vibrational nodes will be present at any closed end. If this
principle is applied to open-end air columns, then the pattern for the fundamental frequency (the
lowest frequency and longest wavelength pattern) will have antinodes at the two open ends and a
single node in between. For this reason, the standing wave pattern for the fundamental frequency
(or first harmonic) for an open-end air column looks like the diagram below.

Fig. 2.52

163
The distance between antinodes on a standing wave pattern is equivalent to one-half of a
wavelength. A careful analysis of the diagram above shows that adjacent antinodes are
positioned at the two ends of the air column. Thus, the length of the air column is equal to one-
half of the wavelength for the first harmonic.

The standing wave pattern for the second harmonic of an open-end air column could be produced
if another antinode and node was added to the pattern. This would result in a total of three
antinodes and two nodes. This pattern is shown in the diagram below. Observe in the pattern that
there is one full wave in the length of the air column. One full wave is twice the number of
waves that were present in the first harmonic. For this reason, the frequency of the second
harmonic is two times the frequency of the first harmonic.

Fig. 2.53
And finally, the standing wave pattern for the third harmonic of an open-end air column could be
produced if still another antinode and node were added to the pattern. This would result in a total
of four antinodes and three nodes. This pattern is shown in the diagram below. Observe in the
pattern that there are one and one-half waves present in the length of the air column. One and
one-half waves is three times the number of waves that were present in the first harmonic. For
this reason, the frequency of the third harmonic is three times the frequency of the first harmonic.

Fig. 2.54

Summary of Length-Wavelength Relationships

The process of adding another antinode and node to each consecutive harmonic in order to
determine the pattern and the resulting length-wavelength relationship could be continued. If
doing so, it is important to keep antinodes on the open ends of the air column and to maintain an
alternating pattern of nodes and antinodes. When finished, the results should be consistent with
the information in the table below.

The relationships between the standing wave pattern for a given harmonic and the length-
wavelength relationships for open end air columns are summarized in the table below.

164
Table 2.7
# of Length-
Harm. # of # of
Waves in Wavelength
# Nodes Antinodes
Column Relationship
1 ½ 1 2 Wavelength = (2/1)*L
2 1 or 2/2 2 3 Wavelength = (2/2)*L
3 3/2 3 4 Wavelength = (2/3)*L
4 2 or 4/2 4 5 Wavelength = (2/4)*L
5 5/2 5 6 Wavelength = (2/5)*L

Problem-Solving Scheme

Now the aim of the above discussion is to internalize the mathematical relationships for open-
end air columns in order to perform calculations predicting the length of air column required to
produce a given natural frequency. And conversely, calculations can be performed to predict the
natural frequencies produced by a known length of air column. Each of these calculations
requires knowledge of the speed of a wave in air (which is approximately 340 m/s at room
temperatures). The graphic below depicts the relationships between the key variables in such
calculations. These relationships will be used to assist in the solution to problems involving
standing waves in musical instruments.

Fig. 2.55

165
To demonstrate the use of the above problem-solving scheme, consider the following example
problem and its detailed solution.

Example Problem
The speed of sound waves in air is found to be 340 m/s. Determine the fundamental frequency
(1st harmonic) of an open-end air column that has a length of 67.5 cm.

The solution to the problem begins by first identifying known information, listing the desired
quantity, and constructing a diagram of the situation.

Given: Find: Diagram:

v = 340 m/s f1 = ??

L = 67.5 cm = 0.675 m

The problem statement asks us to determine the frequency (f) value. From the graphic above, the
only means of finding the frequency is to use the wave equation (speed = frequency •
wavelength) and knowledge of the speed and wavelength. The speed is given, but wavelength is
not known. If the wavelength could be found then the frequency could be easily calculated. In
this problem (and any problem), knowledge of the length and the harmonic number allows one to
determine the wavelength of the wave. For the first harmonic, the wavelength is twice the length.
This relationship is derived from the diagram of the standing wave pattern (see table above). The
relationship, which works only for the first harmonic of an open-end air column, is used to
calculate the wavelength for this standing wave.

Wavelength = 2 • Length

Wavelength = 2 • 0.675 m

Wavelength = 1.35 m

Now that wavelength is known, it can be combined with the given value of the speed to calculate
the frequency of the first harmonic for this open-end air column. This calculation is shown
below.

speed = frequency • wavelength

frequency = speed / wavelength

frequency = (340 m/s) / (1.35 m)

frequency = 252 Hz

166
Most problems can be solved in a similar manner. It is always wise to take the extra time needed
to set the problem up. Take the time to write down the given information and the requested
information, and to draw a meaningful diagram.

Seldom in physics are two problems identical. The tendency to treat every problem the same way
is perhaps one of the quickest paths to failure. It is much better to combine good problem-solving
skills (part of which involves the discipline to set the problem up) with a solid grasp of the
relationships among variables. Avoid the tendency to memorize approaches to different types of
problems.

To further your understanding of these relationships and the use of the above problem-solving
scheme, consider the following example problem and its detailed solution.

Example Problem
Determine the length of an open-end air column required to produce a fundamental frequency
(1st harmonic) of 480 Hz. The speed of waves in air is known to be 340 m/s.

The solution to the problem begins by first identifying known information, listing the desired
quantity, and constructing a diagram of the situation.

Given: Find: Diagram:

v = 340 m/s L = ??

f1 = 480 Hz

The problem statement asks us to determine the length of the air column. When inspecting the
problem-solving scheme described above, one will notice that the only means of finding the
length of the air column is from knowledge of the wavelength. But the wavelength is not known.
However, the frequency and speed are given, so one can use the wave equation (speed =
frequency • wavelength) and knowledge of the speed and frequency to determine the
wavelength. This calculation is shown below.

speed = frequency • wavelength

wavelength = speed / frequency

wavelength = (340 m/s) / (480 Hz)

wavelength = 0.708 m

167
Now that the wavelength is found, the length of the air column can be calculated. For the first
harmonic, the length is one-half the wavelength. This relationship is derived from the diagram of
the standing wave pattern (see table above. The relationship may also be evident to you by
looking at the standing wave diagram drawn above. This relationship between wavelength and
length, which works only for the first harmonic of an open-end air column, is used to calculate
the wavelength for this standing wave.

Length = (1/2) • Wavelength

Length = (1/2) • Wavelength

Length = 0.354 m

If you have successfully managed the above two problems, take a try at the following practice
problems. As you proceed, be sure to be mindful of the numerical relationships involved in such
problems. And if necessary, refer to the graphic above.

Exercise 2.10
1. Stan Dinghwaives is playing his open-end pipe. The frequency of the second harmonic is 880
Hz (a pitch of A5). The speed of sound through the pipe is 350 m/sec. Find the frequency of the
first harmonic and the length of the pipe.

2. On a cold frigid day, Matthew blows on a toy flute, causing resonating waves in an open-end
air column. The speed of sound through the air column is 336 m/sec. The length of the air
column is 30.0 cm. Calculate the frequency of the first, second, and third harmonics.

3. A flute is played with a first harmonic of 196 Hz (a pitch of G3). The length of the air column
is 89.2 cm (quite a long flute). Find the speed of the wave resonating in the flute.

168
2.5.4 Closed-End Air Columns
In the previous sub-section of Section 2.5, the formation of a standing wave patterns in an open-
end instrument was discussed. The mathematics of the harmonic frequencies associated with
such standing wave patterns were developed. This sub-section will use similar principles to
develop the standing wave patterns and associated mathematics for closed-end air column. An
instrument consisting of a closed-end column typically contains a metal tube in which one of the
ends is covered and not open to the surrounding air. Some pipe organs and the air column within
the bottle of a pop-bottle orchestra are examples of closed-end instruments. Some instruments
that operate as open-end air columns can be transformed into closed-end air columns by covering
the end opposite the mouthpiece with a mute. As we will see the presence of the closed end on
such an air column will affect the actual frequencies that the instrument can produce.

As has already been mentioned, a musical instrument has a set of natural frequencies at which it
vibrates at when a disturbance is introduced into it. These natural frequencies are known as the
harmonics of the instrument. Each harmonic is associated with a standing wave pattern. In
Section 1.4 of Chapter 1, a standing wave pattern was defined as a vibrational pattern created
within a medium when the vibrational frequency of the source causes reflected waves from one
end of the medium to interfere with incident waves from the source in such a manner that
specific points along the medium appear to be standing still. In the case of stringed instruments
(discussed earlier), standing wave patterns were drawn to depict the amount of movement of the
string at various locations along its length. Such patterns show nodes - points of no displacement
or movement - at the two fixed ends of the string. In the case of air columns, a closed end in a
column of air is analogous to the fixed end on a vibrating string. That is, at the closed end of an
air column, air is not free to undergo movement and thus is forced into assuming the nodal
positions of the standing wave pattern. Air at the closed end of an air column is still. Conversely,
air is free to undergo its back-and-forth longitudinal vibration at the open end of an air column.
And as such, the standing wave patterns will depict vibrational antinodes at the open ends of air
columns.

So the basis for drawing the standing wave patterns for air columns is that vibrational antinodes
will be present at any open end and vibrational nodes will be present at any closed end. If this
principle is applied to closed-end air columns, then the pattern for the fundamental frequency
(the lowest frequency and longest wavelength pattern) will have a node at the closed end and an
antinode at the open end. For this reason, the standing wave pattern for the fundamental
frequency (or first harmonic) for a closed-end air column looks like the diagram below.

Fig. 2.56
The distance between adjacent antinodes on a standing wave pattern is equivalent to one-half of
a wavelength. Since nodes always lie midway in between the antinodes, the distance between an
antinode and a node must be equivalent to one-fourth of a wavelength. A careful analysis of the
diagram above shows that a node and an adjacent antinode are positioned at the two ends of the

169
air column. Thus, the length of the air column is equal to one-fourth of the wavelength for the
first harmonic.

The fundamental frequency is the lowest possible frequency that any instrument can play; it is
sometimes referred to as the first harmonic of the instrument. The second harmonic of any
instrument always has a frequency that is twice the frequency of the first harmonic. The fourth
harmonic of any instrument always has a frequency that is four times the frequency of the first
harmonic. As we will see, a strange pattern results for a closed-end air column. Just as for all the
instruments, the next harmonic for a closed-end air column is the harmonic that has one more
node. And just as for all the instruments, the addition of an extra node also means that an extra
antinode must also be added to the pattern. This would result in a total of two vibrational
antinodes and one vibrational node. This pattern is shown in the diagram below. Observe in the
pattern that there is three-fourths of a full wave in the length of the air column. That is three
times the number of waves in the first harmonic. Since, the frequency of this harmonic is three
times the frequency of the first harmonic, this is called the third harmonic.

Fig. 2.57
But what happened to the second harmonic? Unlike the other instrument types, there is no
second harmonic for a closed-end air column. The next frequency above the fundamental
frequency is the third harmonic (three times the frequency of the fundamental). In fact, a closed-
end instrument does not possess any even-numbered harmonics. Only odd-numbered harmonics
are produced, where the frequency of each harmonic is some odd-numbered multiple of the
frequency of the first harmonic.

The next highest frequency above the third harmonic is the fifth harmonic. It is the standing
wave pattern with the next smallest wavelength. The standing wave pattern for the fifth harmonic
of a closed-end air column is produced by adding another node to the pattern. This would result
in a total of three anti-nodes and three nodes. This pattern is shown in the diagram below.
Observe in the pattern that there are one and one-fourth waves present in the length of the air
column. That is five times the number of waves in the first harmonic. For this reason, the
frequency of the fifth harmonic is five times the frequency of the first harmonic.

Fig. 2.58

The process of adding another node and antinode to each consecutive harmonic in order to
determine the pattern and the resulting length-wavelength relationship could be continued. If
doing so, it is important to keep vibrational antinodes on the open ends and vibrational nodes on
the closed end of the air column and to maintain an alternating pattern of nodes and antinodes.
When finished, the results should be consistent with the information in the table below.

170
The relationships between the standing wave pattern for a given harmonic and the length-
wavelength relationships for closed-end air columns are summarized below.

Table 2.8
Length-
Harm. # of Waves # of # of
Wavelength
# in Column Nodes Antinodes
Relationship

¼ 1 1 = (4/1)*L
1

¾ 2 2 = (4/3)*L
3

5/4 3 3 = (4/5)*L
5

7/4 4 4 = (4/7)*L
7

9/4 5 5 = (4/9)*L
9

(The symbol represents the wavelength.)

Now the aim of the above discussion is to internalize the mathematical relationships for closed-
end air columns in order to perform calculations predicting the length of air column required to
produce a given natural frequency. And conversely, calculations can be performed to predict the
natural frequencies produced by a known length of air column. Each of these calculations
requires knowledge of the speed of a wave in air (which is approximately 340 m/s at room
temperatures). The graphic below depicts the relationships between the key variables in such
calculations. These relationships will be used to assist in the solution to problems involving
standing waves in musical instruments.

Fig. 2.59
171
To demonstrate the use of the above problem-solving scheme, consider the following example
problem and its detailed solution.

Example Problem
The speed of sound waves in air is 340 m/s. Determine the fundamental frequency (1st
harmonic) of a closed-end air column that has a length of 67.5 cm.

The solution to the problem begins by first identifying known information, listing the desired
quantity, and constructing a diagram of the situation.

Given: Find: Diagram:

v = 340 m/s f1 = ??

L = 67.5 cm = 0.675 m

The problem statement asks us to determine the frequency (f) value. From the graphic above, the
only means of finding the frequency is to use the wave equation (speed=frequency • wavelength)
and knowledge of the speed and wavelength. The speed is given, but wavelength is not known. If
the wavelength could be found then the frequency could be easily calculated. In this problem
(and any problem), knowledge of the length and the harmonic number allows one to determine
the wavelength of the wave. For the first harmonic, the wavelength is four times the length. This
relationship is derived from the diagram of the standing wave pattern (see table above). The
relationship, which works only for the first harmonic of a closed-end air column, is used to
calculate the wavelength for this standing wave.

Wavelength = 4 • Length

Wavelength = 4 • 0.675 m

Wavelength = 2.7 m

Now that wavelength is known, it can be combined with the given value of the speed to calculate
the frequency of the first harmonic for this closed-end air column. This calculation is shown
below.

speed = frequency • wavelength

frequency = speed / wavelength

frequency = (340 m/s) / (2.7 m)

frequency = 126 Hz

172
Most problems can be solved in a similar manner. It is always wise to take the extra time needed
to set the problem up; take the time to write down the given information and the requested
information, and to draw a meaningful diagram.

Seldom in physics are two problems identical. The tendency to treat every problem the same way
is perhaps one of the quickest paths to failure. It is much better to combine good problem-solving
skills (part of which involves the discipline to set the problem up) with a solid grasp of the
relationships among variables. Avoid the tendency to memorize approaches to different types of
problems.

To further demonstrate the use of the above problem-solving scheme, consider the following
example problem and its detailed solution.

Example Problem
Determine the length of an closed-end air column that produces a fundamental frequency (1st
harmonic) of 480 Hz. The speed of waves in air is known to be 340 m/s.

The solution to the problem begins by first identifying known information, listing the desired
quantity, and constructing a diagram of the situation.

Given: Find: Diagram:

v = 340 m/s L = ??

f1 = 480 Hz

The problem statement asks us to determine the length of the air column. From the graphic
above, the only means of finding the length of the air column is from knowledge of the
wavelength. But the wavelength is not known. However, the frequency and speed are given, so
one can use the wave equation (speed = frequency • wavelength) and knowledge of the speed
and frequency to determine the wavelength. This calculation is shown below.

speed = frequency • wavelength

wavelength = speed / frequency

wavelength = (340 m/s) / (480 Hz)

wavelength = 0.708 m

173
Now that the wavelength is found, the length of the air column can be calculated. For the first
harmonic, the length is one-fourth the wavelength. This relationship is derived from the diagram
of the standing wave pattern (see table above); it may also be evident to you by looking at the
standing wave diagram drawn above. This relationship between wavelength and length, which
works only for the first harmonic of a closed-end air column, is used to calculate the wavelength
for this standing wave.

Length = (1/4) • Wavelength

Length = (1/4) • Wavelength

Length = 0.177 m

If you have successfully followed the logic of the two solutions above, then take a try at the
following practice problems. As you proceed, be sure to be mindful of the numerical
relationships involved in such problems. And if necessary, refer to the problem solving scheme
presented above.

Exercise 2.11
1. Titan Tommy and the Test Tubes at a night club this weekend. The lead instrumentalist uses a
test tube (closed-end air column) with a 17.2 cm air column. The speed of sound in the test tube
is 340 m/sec. Find the frequency of the first harmonic played by this instrument.

2. A closed-end organ pipe is used to produce a mixture of sounds. The third and fifth harmonics
in the mixture have frequencies of 1100 Hz and 1833 Hz respectively. What is the frequency of
the first harmonic played by the organ pipe?

3. Pipin' Pete is playing at City Park next weekend. One of the closed-end pipes is capable of
sounding out a first harmonic of 349.2 Hz. The speed of sound in the pipe is 350 m/sec. Find the
length of the air column inside the pipe.

174

View publication stats

You might also like