Вы находитесь на странице: 1из 234

ARCHITECTURAL

ACOUSTICS
Principles and Design
Madan Mehta
University of Texas at Arlington
.James Johnson
WJHW Acoustical Consultants, Dallas, Texas
.Jorge Rocafort
University of Puerto Rico, San Juan

Freehand illustrations by Lily Sun

Prentice Hall
Upper Saddle River, New Jersey Columbus, Ohio

-~--------------------
Library of Congress Cataloging-in-Publication Data

Mehta, Madan.
Architectural acoustics : principles and design I Madan Mehta,
James Johnson, Jorge Rocafort.
p. cm.

Includes bibliographical references and index.


ISBN 0-13-793795-4
1. Architectural acoustics. I. Jolmson, James (James Allison).
II. Rocafort, Jorge. m. Title.
NA2800.M45 1999 98-22448
690'.2-dc21 CIP

Cover photo: 0 The Stock Market


Editor: Ed Francis
Production Editor: Christine M. Harrington
DesignCoordinator: Karri e M. Converse
Cover Designer: Susan Unger
Production Manager: Patricia A Tonneman
Marketing Manager: Danny Hoyt

Tiris book was printed and bound byCourier Kendallville, Inc. The cover was printed by Phoenix
Co1orCorp.

0 1999 by Prentice-Hall, Inc.


Simon & Schuster/A Viacom Company
Upper Saddle River, New Jersey 07458

All rights reserved. No part of this book may be reproduced, in any form or by any means. without
permission in writing from the publisher.

Printed in the United States of America

10 9 8 7 6 5 4 3 2 1

ISBN: 0-13-793795-4

Prentice-Hall International (UK) Limited, London


Prentice-Hall of Australia Pty., Limited, Sydney
Prentice-Hall ofCanada, Inc., Toronto
Prentice-Hall Hispanoamericana, S. A, Mexico
Prentice-Hall of India Private Limited, New Delhi
Prentice-Hall of Japan, Inc., Tokyo
Simon & Schuster Asia Pte. Ltd, Singapore
Editora Prentice-Hall do Brasil, Ltda., Rio de Janeiro

ii
Contents

Preface ix

1 1
1.1 Wave motion 1
The Physics of Sound
1.2 Sound Frequency 4
1.3 Sound Velocity, Particle Velocity and Wavelength 8
1.4 The Decibel Scale 11
1.5 Combining Sound Levels 14
1.6 Sound Attenuation by Distance 17
1.7 Sound Fields 20

2 2.1
25
Sound Level Meter 25
Sound Measurement and 2.2 The Ear's Sensitivity 28
Hearing 2.3 The Haas Effect 31
2.4 Sound Masking 35
2.5 Binaural Hearing 36

3 37
3.1 The Boundary Phenomena 37
Sound Reflection,
3.2 Absorption Coefficient of Sound 39
Diffraction and Diffusion
3.3 Sound Diffraction 39
3.4 The Relevance of Acoustical Shadows 41
3.5 Acoustical Transparency of a Screen 44
3.6 Diffuse and Specular Reflections 48
3.7 Sound Diffusion 49

iii
3.8 Sound Diffusers 52
3.9 Source-Image Relationship in Specular Reflection 57
3.10 Flutter Echo 58

4 61
4.1 Rating of Sound Absorbing Materials 61
Sound Absorbing Materials Types of Sound Absorbing Materials 63
4.2
4.3 Porous Absorbers 64
4.4 Applications of Porous Absorbers 68
4.5 Panel Absorbers 73
4.6 Volume Absorbers 76
4.7 Screens and Perforated Panels as Absorbers 80
4.8 Mounting Conditions 81
4.9 Acoustical Tile Ceiling 82
4.10 Acoustical Plaster and Draperies 83
4.11 Audience and Air Absorption in Halls 84

5 5.1
87
Airborne and Structure-borne Sounds 87
Principles of Airborne 5.2 Sound Transmission Loss 90
Sound Insulation 5.3 Sound Absorption and Sound Insulation 91
5.4 Single-Leaf Panel- The Mass Law 92
5.5 Single-Leaf Panel - The Coincidence Effect 93
5.6 Transmission Loss of a Two-Leaf Panel 97
5.7 Panels of Three or More Leaves 100
5.8 Sound Transmission Class 101
5.9 Overall Transmission Loss of an Assembly 103
5.10 Sound Leaks and Flanking Transmission 105
5.11 Laboratory STC Versus Field STC 108
5.12 Ceiling Sound Transmission Class 108
5.13 Outdoor-Indoor Transmission Class 109

6 115
6.1 Concrete Walls and Slabs 117
Airborne Sound Insulation
6.2 Single-Leaf Masonry Walls 118
Practice
6.3 Masonry Cavity Walls 120
6.4 Furred Masonry Walls 122
6.5 Lightweight Gypsum Board Assemblies 124
6.6 Sound Leaks and the Control of Flanking Paths 129
6.7 Sound Insulating Windows 134
6.8 Sound Insulating Doors 138

iv
7 7.1
143
Impact Insulation Class (llC) 144
Structure-borne Sound 7.2 Strategies to Increase Impact Insulation 146
Insulation 7.3 Soft or Resilient Floor Covering 146
(Impact Isolation) 7.4 Floating Floor 147
7.5 Resiliently Attached Ceiling 154
7.6 Discontinuity in Floor and Ceiling 154

8 8.1
159
Interior Noise Criteria 161
Noise Control in Buildings 8.2 Interior Noise Legislation 169
8.3 Interior Noise Control Through Architectural Design 170
8.4 Interior Noise Control Through Sound Absorptive Treatment 172
8.5 Interior Noise Control Through Barriers 175
8 .6 Exterior Noise Criteria 179
8.7 Exterior Noise Control Through Site Planning 180
8.8 Exterior Noise Control Through Barriers 182

9 9.1
189
HVAC Systems 189
Control ofHVAC Noise 9.2 Noise Attenuation in Ducts 193
9.3 Noise Generated by Air Flow 197
9.4 Sound Radiation by Duct Walls 199
9.5 Cross Transmission in Ducts 201
9.6 Estimating HVAC Noise Levels 202
9.7 Active Noise Control in HVAC Systems 205

10 10.1
207
Impulse Response of a Room 207
The Behavior of Sound in 10.2 Impplse Diagram and Sound Diffusion 210
Rooms 10.3 Reverberation Time 212
10.4 Significance of Reverberation Time 214
10.5 Reverberation Time Calculations 216
10.6 Optimum Reverberation Times 217
10.7 Coupled Rooms 220
10.8 Behavior of Sound in a Small Room 222

11 11.1
229
Speaker Listener Distance
- 230
Design of Rooms for Speech 11.2 Balcony and Hall Depth 232
11.3 Room Shape 236
11.4 Room Volume 238
11.5 Reflecting and Absorbing Parts of a Room 239
11.6 Floor Rake 242
11.7 Ceiling Reflections 248

V
11.8 Ambient Noise 254
11.9 Sound Reinforcement Systems 254
11.10 Special Considerations for Classrooms 254

12 12.1
259
Musical Attributes and Acoustical Phenomena 260
Design of Rooms for Music- 12.2 Early Decay Tune and Clarity 261
Concert Halls and Music 12.3 Intimacy 266
Practice Rooms 12.4 Spaciousness 268
12.5 Warmth and Brilliance 275
12.6 Loudness 276
12.7 Concert Hall Design Procedure 278
12.8 Preliminary Design of a Concert Hall 279
12.9 Music Practice Rooms 283

13 13.1
289
Drama Theaters 289
Tbeaters, Multipurpose 13.2 Multipurpose Halls 294
Halls, Cinema Halls and 13.3 Cinema Halls 296
Studios 13.4 Recording and Broadcasting Studios 301

14 307
307
14.1 Speech Intelligibility
Speech Intelligibility, Speech 14.2 Speech Privacy 313
Privacy and Open-Plan 14.3 Acoustical Design Principles of Open-Plan Offices 317
Offices 14.4 Speech Privacy in an Open-Plan Office 320

15 327
15.1 Vibration Control Strategies 328
Vibration Isolation 15.2 Fundamentals of Vibration Isolation 329
15.3 Types of Isolators 334
15.4 Isolator Bases 337
15.5 Isolation System Design 339

16 341
16.1 Loudspeakers 343
Sound Reinforcement
16.2 Microphones 348
Systems
16.3 Amplifiers 349
16.4 Fundamentals of Sound Reinforcement 350
16.5 Loudspeaker Location 355
16.6 Computer-aided Sound System Design 360
16.7 Electronic Architecture 360

vi
Review Questions 363

Appendixes A SI System of Units 375


B Decibel Addition - The Exact Procedure 379
c Standing Waves and Normal Modes of Vibration 383
D Princ iples of a Quadratic Residue Diffuser 393
E Flat Floor Depth and Raked Floor Slope 395
F Overall Transmission Loss of an Assembly 399
G Reverberation Tune Equations 403
H Sound Absorption Coefficient Data 407
I Impact Insulation Class Data 413
J Sound Transmission Loss Data 414
K Important Formulas in Architectural Acoustics 425

Glossary of Terms 433

Index 443

vii
Preface

ANOTETO THE Although our expertise is different- two of the authors are university
professors and one, an acoustical consultant - all three of us have
READER taught acoustics for several years. Over these years, we have observed
a growing diversity in the backgrounds of students entering
architecture and interior design programs. For some students,
architecture and interior design is a second career. Consequently,
such students have excellent preparation in science and mathematics.
We believe that a growing number will specialize in acoustics or
other technological disciplines in their working lives.
On the other hand, most students enter architecture with a high
school preparation in science and mathematics. 'fP.ese students will
work in the traditional roles of architects and interior designers, who
need a basic understanding of acoustical principles and practice -
not a comprehensive knowledge of the subject.
Architectural Acoustics - Principles and Design has been
written with the above diversity in mind. The text portion of the
book will satisfy a reader who desires an exhaustive coverage of the
subject, while the illustrations - over 500 in number - are aimed
at a reader who does not have the time pr the desire to go through
long descriptions of complex acoustical phenomena.
The text and illustrations are fully integrated with each other.
However, an attempt has been made to lrthe illustrations stand alone
so that the reader can obtain a good working knowledge of the subject
by studying the illustrations along with a cursory reading of the text.
Our teaching experience indicates that an inaqequately illustrated
book is as frustrating as one with numerous illustrations, but with an
insufficient explanatory text.

ix
To improve the book ' s reader friendliness, topics that do not
require a detailed study during its first reading, such as the numerical
examples, description of acoustical standards etc . , have been
separated from the main text by enclosing them in a box with a gray
screen. Although references are given at the end of each chapter,
review questions are placed at the end of the book.
The primary aim of the book is to serve as a text during the
college years of an architect, interior designer, or engineer. However,
we have been able to synthesize academic rigor with practical details.
Therefore, the book should remain useful to the readers during their
working careers also.
A special feature of interest to practicing architects, interior
designers, and acoustical consultants is a large body of acoustical
data gathered by us over the years, and presented in Appendixes H
through J. Additionally, a glossary of acoustical terms and a
comprehensive listing of important formulas in architectural acoustics
should make the book attractive as a reference source.
The book contains more than adequate material for one
semester ' s course in acoustics in architecture, interior design, and
engineering programs. However, it can also be adapted for a half
semester course. In that case, we suggest that the instructor cover
mainly the principles, leaving the practical aspects for self study by
the reader. A suggested list of chapters for a half semester course is:
Chapters 1 to 5, 7, 10, 1 1 and 12.
Since the Imperial system is being phased out in the United
States, we have used the SI system of units as the primary system,
and the Imperial system as the secondary system. This format
corresponds with most contemporary U.S. publications in architecture
and engineering.
Although this is a joint and shared effort, the primary
responsibilities for various parts of the book are:

Madan Mehta: Chapters 1 through 8, 10, 1 1 and 1 3, Review


Questions, Appendixes, and Glossary of
Terms.
lames Johnson: Chapters 9, 12, 1 3, 14 and 1 5.
Jorge Rocafort: Chapter 16.

ACKNOWLEDGMENTS Although it is impossible to acknowledge every individual who has


helped us in this endeavor, we would like to particularly thank the
following persons for a careful and thorough review of the
manuscript. They are: Dr. Sidney Stotesbury, Department of
Architecture, Kansas State University, Manhattan, Kansas; Professor
David Hanna, De partment of Con struction Technology and
Management, Ferris State University, Big Rapids, Michigan ;
Professor Theodore Ceraldi, School of Architecture, Syracuse
University, Syracuse, New York; and Professor Glenn Goldman,
School of Architecture, New Jersey Institute of Technology, Newark,
New Jersey.
We are grateful to several building produc manufacturers who
have provided us with photographs of their products along with the
technical data. Our thanks are also due to our students who gave
useful suggestions and picked up several errors in the manuscript
while preparing for tests in our courses.

X
A special ack nowledgment is due to our editor, Ed Francis of
Prentice Hall, who was always available to answer questions and
provide the necessary help.

DISCLAIMER The information in this book has been derived from several sources,
such as the reference book s, journals, manufacturers' literature,
authors' professional experience, etc. It is presented in good faith,
and although the authors and the publisher have made every
reasonable effort to present the information accurately, they do not
warrant or assume any liability for its accuracy, its completeness, or
its suitability for any specific purpose. It is the responsibility of the
user of this book to apply his/her professional k nowledge in the use
of the information presented here, and to consult original sources
for detailed information as needed.

Madan Mehta
James Johnson
Jorge Rocafort

xi
The Physics of
Sound

This chapter begins with an introduction to the wave phenomenon


and its application to sound waves. Physical characteristics of sound
such as velocity, wavelength, frequency and octave bands are
discussed. Subsequently, the decibel scale and a few other important
concepts are introduced.

1.1 WAVE MOTION Sound is the human ear ' s response to pressure fluctuations in the air
caused by vibrating objects. For example, a tap on a wall produces
sound because the tap makes the wall vibrate. The vibrating wall
produces pressure fluctuations in the air. The same phenomenon
occurs from the vibrations of a guitar string when the string is plucked.
Although most sounds in our environment are produced by
vibrating objects, some sounds do not involve mechanical vibrations.
They are produced by a sudden increase in air velocity, through
turbulence in air flow. Thus, sound is produced when air escapes
out of a compressed air line. Similarly, air escaping out of air
conditioning outlets produces sound because the area of outlets is
smaller than the area of the duct, which increases the velocity of air
at the outlets.
Sound travels in space by a phenomenon called wave motion.
Wave motion in air is similar to the motion of a ripple produced by
dropping a pebble into a pond. Consider a water pond, which is
undisturbed by any air movement so that the surface of water is calm
and free from all motion, and a pebble is dropped into the pond. As
soon as the pebble strikes the surface, a ripple radiates out from the
point of impact in an ever increasing circular ring.

1
2 Chapter 1

A closer examination of the ripple indicates that when the pebble


Water eurface strikes the surface of the water, it creates a local depression on the
surface. Since water, like most liquids, is incompressible, the particles
of water adjacent to the point where the pebble strikes are forced
upwards. Thus, a ripple consisting of a trough and two crests, as
shown in Figure 1.1, is formed at the surface of the water.
Once the pebble has passed through the surface of water, the
elasticity of bulk water tends to bring the water surface back to its
normal state, causing the water particles to oscillate about their
original undisturbed positions. This motion is transmitted to
neighboring particles, so that the the ripple, and the energy contained
in it, advances away from the point of the ripple's origin.
Observations of light floating objects in ponds indicate that while
1.1 Ripple produced on the ripple (or the wave) travels forward, the particles of water do
the surface of water by a not. The water particles simply oscillate up and down from their
pebble. original positions. Thus in a wave motion it is the energy contained
in the wave that travels, not the particles of the medium.
Most waves in nature do not contain a single ripple, but rather a
series of ripples, each following the preceding ripple after a constant
time interval. The physical quantity that is related to the number of
ripples generated in unit time is called frequency. Other important
quantities associated with wave motion are wavelength and the speed
of wave travel. In fact, frequency, wavelength and the speed with
which the wave travels are related to each other by a simple
relationship. This relationship can be obtained with reference to the
examination of successive ripples produced in a liquid medium such
as water.
Assume that the speed of travel of ripples (or waves) in a liquid
medium is 100 m per minute and the ripples are formed by dropping
pebbles at the same location at a constant interval of 12 sec, i.e., at a
frequency of 5 pebbles per min. Since the speed of wave travel is
100 m per min, the first ripple will have traveled a distance of 20 m
when the second pebble hits the surface of the water, Figure 1.2(b).
At the instant when the third pebble hits the surface of the water, the
first ripple will be at a distance of 40 m and the second ripple at a
distance of 20 m, Figure 1.2(c). Similarly, when the fourth pebble
hits the surface, the first, second and third ripples will be at distances
of 60 m, 40 m and 20 m respectively, as shown in Figure 1.2(d).
Thus we note that at any given instant, the distance between adjacent
ripples is constant.
If the frequency is increased to 10 cycles per min, we will see
that the distance between adjacent ripples is reduced to 10 m. If the
frequency is 2 cycles per min, the corresponding distance between
ripples increases to 50 m.
The distance between two adjacent ripples at any instant is called
the wavelength, which is denoted by the Greek symbol lambda (A).
If the frequency is denoted by f, and the speed by c, the relationship
between frequency, wavelength and the speed of a wave motion, as
obtained from the above observation, is given by:

c =fA. (1.1)

Although the above relationship was obtained by considering


ripples on the surface of a liquid, it applies to all kinds of wave
motion, including sound waves. Consider once again a wall set into
vibration by an impact. As the wall moves to one side, say to the
left, the air directly in contact with the left side of the wall is pushed
The Physics of Sound 3


- !..
,;;=----.-- ... \ \ \

(a) lnetsnt when the flret pebble hlte the (b) lnetant when the eec;ond pebble hlte the
water (time zero) water (time 12 eec)

Wavelength
(c;) Instant when third pebble hlte the water (d) lnetsnt when the fourth pebble hlte the
(time 24 eec) water (time 36 eec;).

Speed (c) = 100 m/min.


Frequency (f) = 5/mln.
Wavelength (A.)= 20 m

1.2 Location of ripples on water at various instants.


4 Chapter 1

to the left, creating compression in that layer of air. This compressed


layer pushes the layer of air further to its left, which in turn pushes
the next layer to the left, and so on. In this way, a compression
ripple (or compression wave) travels away from the left of the wall
in a domino-like manner.
When the wall reverses dirction, i.e., when it moves to the right,
the opposite condition occurs. The layer of air to the left of the wall
that was earlier in compression suffers a reduction in pressure, called
rarefaction. The rarefaction is transferred to successive layers and a
rarefaction wave travels away from the wall to the left exactly in the
same way as the compression wave traveled (see also Figure 1.9).
Compression and rarefaction waves also travel to the right side
of the wall. In fact when the air in contact with the left side of the
wall is under compression, the air on the right side is under
rarefaction, and vice versa.

1.2 SOUND FREQUENCY It is the succession of compression and rarefaction waves traveling
away from a vibrating source that we refer to as the sound wave. A
compression and rarefaction wave creates vibration (back-and-forth
motion) of air particles. One back-and-forth motion of an air particle
is called a cycle. Thus in one cycle, the particle starts from its original
position of rest (shown by the black dot in Figure 1.3a), moves to
the extreme right, back to the particle's original position, to the
extreme left, and finally back to its original position.
The above process repeats itself in the second and subsequent
cycles. Note that the velocity of the particle in a cycle is maximum
at its central position, and zero at extreme left and right, in the same
way as the velocity of a pendulum is zero at extreme positions and
maximum at the pendulum's central position, Figure 1.3(b). The
maximum velocity of the particle is called its velocity amplitude,
and the maximum displacement of the particle from its rest position
is called the displacement amplitude.
The number of cycles that the air particles move back and forth
in one second in a sound wave is called the frequency of the wave.
Its unit is cycles per second (c/s) which is also termed Hertz (Hz)
after the Austrian physicist Heinrich Hertz (1857 -94). Subjectively,
the frequency of a sound wave is perceived as its pitch. A high
pitched sound means that it has a high frequency. The female voice
is slightly higher pitched than the male voice.
Frequency of sound is an important acoustical concept since the
properties of building materials and construction assemblies vary
with the frequency of sound. Additionally, the behavior of sound in
an enclosure is also dependent on its frequency.
A normal young adult is capable of hearing sounds ranging from
20 Hz to 20 kHz (1 kHz= 1 kilo Hertz= 1 ,000 Hz). Frequencies
below 20 Hz are called infrasonic frequencies. They are not heard
but are perceived by humans as vibrations. Frequencies above 20
kHz, referred to as ultrasonic frequencies, are also not heard by
humans, although some animals can hear them. Dogs can detect
frequencies up to 30 kHz, and bats up to 90 kHz. In fact, bats produce
high frequency signals and use them as an "acoustic radar" to detect
obstacles at night.
Sounds in our environment do not generally consist of individual
frequencies (a single note or pure tone) such as that produced by a
tuning fork. If the frequency is represented on the horizontal axis
The Physics of Sound 5

>
(----E--
<
)
--

------7
- - - - - - - - - - - -8
---

G
Zero particle Maximum particle Zero particle
Maximum
velocity velocity velocity Zero velocity
velocity

l
Dleplacement amplitude
(a)
(1-J)

1.3 Three cardinal positions of a vibrating particle and the corresponding particle velocities.

and the energy contained in the sound is represented on the vertical


axis, the graphical representation of a pure tone is a vertical line.
For instance, if a tuning fork of 500 Hz is struck , the representation
of sound created by it will be as shown in Figure 1.4(a). The harder
the tuning fork is struck , the louder the sound emitted by the tuning
fork, i.e., the greater the sound energy produced. Therefore, the line
representing the sound will be taller.
If two pure tones of different frequencies, say one at 250 Hz and
the other at 500 Hz, are produced simultaneously, their graphical
representation will be two vertical lines. The two vertical lines will
be of the same height if the energy content of each tone is the same.
If the energy content of the two tones is different, their heights on
the graph will be unequal, Figure 1.4(b).
The frequency-energy relationship of a sound is called the sound
spectrum. Most sounds in our environment are complex, i.e., they

500 250 500


Frequency (Hz) Frequency (Hz)

(a) (1-J)

1.4 Sound spectrum of (a) 500Hz tuning fork, (b) 250Hz and 500
Hz tuning forks sounding simultaneously.
6 Chapter 1

consist of a continuum of frequencies, so that their spectrum is a


continuous curve. For example, the human voice consists of a
continuum of sounds ranging from nearly 100Hz to 5 kHz, Figure
1.5. The male voice peak s at nearly 400Hz and the female voice at
nearly 500Hz.

1.2.1 Octave Bands


63 250 1,000 4,000
125 500 2,000 8,000 Since the range of audible sound frequencies is rather large (20Hz
to 20 kHz), it is not possible to deal with individual frequencies.
Frequency (Hz)
Therefore, we divide the entire range of audible sound frequencies
into frequency bands, and study the properties of building materials,
1.5 Approximate sound spectra of male enclosures and sound sources with respect to these frequency bands.
A frequency band is simply a continuum of frequencies lying
and female speech (long-term average).
between two frequency limits. The width of the band is the difference
between the lower frequency limit and the upper frequency limit.
For example, if we arbitrarily choose a bandwidth of 200 Hz, the
lower frequency limit of the first band will be 20Hz and the higher
frequency limit, 220Hz. The lower and upper frequency limits of
the next band will be 220Hz and 420Hz respectively, followed by a
band ranging from 420Hz to 620Hz, and so on. In these frequency
800Hz bands, the center frequency of the first band (the band extending
from 20Hz to 220Hz) is 120Hz, the center frequency of the next
band is 320Hz, and so on. The center frequency of each band was
obtained by (arithmetically) averaging the upper and lower frequency
limits, i.e., by adding the lower and upper frequency limits and
Octave dividing the sum by 2.
In architectural acoustics, we divide the entire range of audible
frequencies into octave bands or simply octaves, a concept that has
been borrowed from music. An octave is a band of frequencies whose
400Hz upper frequency limit is twice the lower frequency limit1. For
example, the frequency interval from 200Hz to 400Hz is an octave,
Octave the interval from 400Hz to 800Hz is the next higher octave, and so
on, Figure 1.6.
The center frequency of an octave cannot be obtained by
200Hz
(arithmetically) averaging the upper and lower frequency limits since
Octave the bandwidths of successive octaves are not equal, but increase by
a factor of 2. For instance, the bandwidth of the octave ranging
100Hz
from 200Hz to 400Hz is 200Hz, the bandwidth of octave ranging
from 400Hz to 800Hz is 400Hz.

1.6 Definition of an octave.

1 The ear's perception of the interval between two frequencies is based on their
ratio, not on the arithmetic difference between them. For instance, the interval
between 200 and 300Hz is not perceived to be the same as that between 100 and
200Hz, although both pairs of frequencies are 100Hz apart. In fact, the interval
between 200 and 300Hz appears to be smaller than the interval between 100 and
200Hz. To obtain the same interval sensation as that between 100 and 200Hz (a
ratio of 1 :2), the other pair must also be in the ratio of 1:2. Thus, the interval
between 100 and 200Hz is perceived to be the same as that between 200 and 400
Hz, or between 1,000 and 2,000Hz, or beween 5,000 and 10,000Hz.
When equal interval sensations between two magnitudes of a quantity are
obtained by the same ratio, we say that the quantity follows a logarithmic scale.
Thus, frequency perception follows a logarithmic scale. In fact, as we shall see
later in this chapter, the subjective perception of most acoustical quantities follows
a logarithmic scale. A brief review of logarithms is given at the end of this chapter.
The Physics of Sound 7

The center frequency of an octave band is obtained by


geometrically averaging the upper and lower frequency limits, i.e.
by multiplying the upper and lower frequency limits and square
rooting the product. Thus, the center frequency of a band ranging
from 200Hz to 400Hz is given by:

J 200 x 400 = 282.8 Hz, or say 283 Hz

Since the upper frequency limit (fu)and lower frequency limit


(fL) of an octave are in the ratio of 2:1, fu and fL for an octave with
a given center frequency (fc) are obtained by the following
relationships.

fu = fc (J2 ) = 1.414 fc

= 0.707 fc

In architectural acoustics, frequencies below 50Hz and beyond


10 kHz are seldom important. Therefore, eight octaves with the
following center frequencies are used in architectural acoustics: 63
Hz, 125Hz, 250Hz, 500Hz, 1 kHz, 2 kHz, 4 kHz and 8 kHz. The
upper and lower frequency limits of each octave are shown in Figure
1. 7. For instance, the lower and upper frequency limits of the octave
centered at 250Hz are 177Hz and 354Hz. Octaves from 63Hz to
250Hz are generally referred to as low frequencies, 500Hz and 1
kHz as mid frequencies, and 2 kHz to 8 kHz as high frequencies.

89 177 707 1.4 k 2.8 k 5.6 k 11.:3 k

6:3 125 250

1.7 Octaves, their center frequencies,


and upper and lower frequency limits.

1.2.2 One-third Octave Bands

An octave band study (analysis) of materials, noise environments


and spaces may r1ot be adequately detailed for some situations. In
such situations, we perform one-third octave band analysis. A one
third octave band analysis is a finer analysis than an octave band
analysis. In a one-third octave band analysis, each octave is divided
in three parts. Once again, the division of an octave into three one
third octaves is geometric rather than arithmetic.
Just as we multiply the center frequency of an octave by 2 to
obtain the center frequency of the next higher octave, we multiply
8 Chapter 1

Table 1. 1 Center the center frequency of a one-third octave by 2113 to obtain the center
Frequencies of One-third frequency of the next higher one-third octave. Dividing by 2113 gives
the center frequency of the next lower one-third octave.
Octave Bands
For instance, when the octave centered at 125 Hz is divided in
50 800 three one-third octaves, the center frequencies are: 100Hz (obtained
63 1,000 by dividing 125 by 1.26), 125Hz and 160Hz (obtained by multiplying
80 1,250 125 by 1.26). The center frequencies of various one-third octave
bands are shown in Table 1.1. Note that 2113 = 1.26.
lOO 1,600
125 2,000
160 2,500 1.2.3 Frequency Ranges of Speech and Music

200 3,150 As indicated in Figure 1.5, the frequency range of speech extends
250 4,000 from nearly 100 Hz to 5 KHz, covering nearly 5 octaves.
315 5,000 Approximately 75% of sound energy in speech is contained in vowels,
which are low frequency components of speech. It is the sound in
400 6,300 vowels that accounts for the distinguishing quality of an individual's
500 8,000 speech.
630 10,000 The consonants are the high frequency components of speech.
The energy contained in consonants is relatively small, but it is the
consonants that provide intelligibility in human speech. Frequencies
below 500 Hz contribute negligibly to speech intelligibility. The
frequency range of speech is shown in Figure 1.8, which also gives
frequency ranges for music and acoustical laboratory tests.

1.8 Frequency ranges of :31 6:3 125 250 500 1.000 2,000 4,000 8,000 16,000
acoustical laboratory tests, Frequency (Hz)
human speech, and m usic.

1.3 SOUND VELOCITY, The speed of sound in air has been measured as 344 m/sec (1,130 ft/
sec). This corresponds to 1240 km/hr (770 milhr) which is extremely
PARTICLE VELOCITY small as compared to the speed of light (300,000 km/sec). An object
AND WAVELENGTH traveling at a speed greater than the speed of sound is said to be
traveling at supersonic speed.
The speed of sound in air does not vary with the frequency of
sound or its loudness. In other words, sounds at all audible
frequencies, regardless of their loudness, travel at the same speed.
The Physics of Sound 9

Table 1.2 Wavelength of Note that unlike light, which does not require a medium in which
So und at Center to travel, sound energy, being vibrational energy of the particles of
medium, requires a medium in which to travel - from one point in
Frequencies of Octaves
the medium to another. Thus, sound cannot travel in a vacuum, while
Frequency Wavelength
light can.
(ft)
In solids, the speed of sound (that is, the speed of travel of
(Hz) (m)
vibrational energy) is considerably greater than in gases or liquids.
For instance, the speed of sound in steel is nearly fifteen times greater
63 1 8.0 5.46
than that in air. This explains why we are able to hear the vibrations
125 9.0 2.75
of railroad tracks well before we receive the sound from the arriving
250 4.5 1.38 train, which reaches us by traveling through the air.
500 2.3 0.69
1,000 1.1 0.34
2,000 0.6 0.17 1.3.1 Wavelength of Sound
4,000 0.3 0.09
8,000 0.15 0.04 As stated in Section 1.2, the wavelength and the frequency of sound
are related to each other according to Equation ( 1 . 1 ). The greater
the frequency of sound, the smaller its wavelength. Thus, the
wavelength of sound at 20 Hz is 344/20 = 17.2 m (56.5 ft). At 20
kHz, the wavelength is 1 .72 cm (0.7 in.). ['he wavelength of sound
corresponding to the center frequencies of various octaves is shown
in Table 1.2.
Although the velocity of sound is 344 m/sec (1,130 ft/sec), for
most practical purposes we may regard it as 3 00 m/sec (or 1 ,000 ft/
sec). With that approximation, the wavelength of sound at 100 Hz is
nearly 3 m (10 ft), at 1 kHz the wavelength is nearly 0.3 m, or 30 cm
(1 ft), at 2 kHz, the wavelength is 15 cm (6 in.), and so on.
Physically, the wavelength of a sound wave is the distance
between adjacent compression peaks or adjacent rarefaction peaks,
Figure 1.9. In fact, the wavelength of sound is the distance between
two air particles that are in exactly the same vibrational situation.

1.9 The distribution of air particles at any instant in a


sound wave. The wavelength of sound is the distance
between adjacent compressions or adjacent rarefactions
in air.
10 Chapter 1

1.3.2 Particle Velocity and Displacement

Assume a sound source operating at a certain loudness and frequency


in an open space - a space in which there is no reflection2 of sound.
If we could see the vibration of air particles in such a space, we
would observe that each particle is oscillating back and forth about
its central position in an identical manner. More specifically, we
would observe that each particle has the same maximum displacement
(displacement amplitude) and the same maximum particle velocity
(velocity amplitude).
It is important to be reminded that particle velocity is not the
same as sound velocity. Sound velocity in a medium is constant
(344 m/sec) and is the velocity at which sound energy travels from
one point in the medium to another. Particle velocity, on the other
hand, varies from zero to a maximum value in a cy cle, as shown in
Figure 1.3, and is much smaller than sound velocity. Even in an
extremely loud (deafening) environment, the velocity amplitude is
less than 1 cm/sec.
The loudness of sound is a function of the energy contained in a
sound wave, which in turn is a function of the kinetic energy of air
particle oscillation. Therefore the loudness3 of sound is directly
related to velocity amplitude of particles. This fact is important
because in order to reduce noise levels, all that is necessary is to
reduce particle velocity. In fact, as we will see in Section 4.3, the
absorbing property of one class of sound absorbers is due to its ability
to reduce particle velocities.
A related and an equally important fact, discussed in detail in
Appendix C, is that the maximum particle velocity occurs at a distance
of 0.25A. from room boundaries. Particle velocity is obviously zero
at room boundaries, since room boundaries are rigid, unable to
oscillate.

2 A space in which there is no reflection of sound is also called a free space


- implying freedom from reflections. An anechoic chamber provides such a
space (see Figure 1.15).

3 Theoretically, the quantity that is related to velocity amplitude of particles is


sound intensity, but we may use loudness here, since sound intensity is introduced
later in this chapter, and our ears perceive sound intensity as the loudness of sound.
The Physics of Sound 11

1.4 THE DECIBEL The physical quantity associated with the loudness of sound i s its
intensity, which is defined as the amount of sound power falling on
SCALE (or passing through, or crossing) a unit area. Since the unit of power
is watt, the unit of sound intensity is watts per square meter (W/m2).
! sound intensity whih is just adible, called the threshold o,j
audzbzhty, has been determmed expenmentally to be 10-12 W/m ,
and the intensity that corresponds to the sensation of pain in the
human ear4 is approximately 10 W/m2. Thus, the ear responds to a
very large range of intensities since the loudest sound is
10,000,000,000,000 times (1013 times) louder than the faintest sound.
In a situation where one has to deal with a large range of
magnitudes of a given quantity, it is more convenient to express the
magnitude as a ratio of a reference magnitude. In the case of sound
intensity, the reference intensity is assumed to be 10-12 W/m2 and a
given sound intensity is expressed as a multiple of this reference
intensity. The multiple is obtained by dividing the given intensity
(I) by the reference intensity (lref), i.e., the multiple is given by the
ratio: (1/Iref).
For instance, a sound intensity of 10-10 W1m2 is expressed as
100, since 10-10 W/m2 is 100 times the reference sound intensity of
10-12wtm2. Similarly, 10-9W/m2 is 1,000, and 10-8W/m2 is 10,000,
and so on.
Dealing with a large range of magnitudes is further simplified if
we express them as the logarithm of the above ratio (a review of
logarithms i given at the end of this chapter). Thus, the intensity of
a given sound can be expressed as:

I
log where, I
ref 10 -12 W/m2
I ref
--
=

Note that since the above logarithm is a ratio of two intensities,


and since the ratio of two identical quantities has no units, the above
logarithm is a unit-less (dimension-less) quantity. However, we
arbitrarily assign to it the unit "Bel" in recognition of Alexander
Graham Bell (1847-1922), the inventor of the telephone.
The numbers obtained from the above logarithmic ratio are a
little too small. Therefore, it is more common to use one-tenth of a
Bel, called the decibel (a,bbreviated as dB) as the unit5 of this
logarithm. In other words, decibel is the unit of the following
logarithmic ratio:

I
IL = 10 log (1.2)
I ref
-

The quantity defined by the above logarithm is called the sound


intensity level (IL), to distinguish it from sound intensity (1). Note
once again that the unit of sound intensity level is dB, while the unit
of sound intensity is W1m2.

4 The sound that produces pain in the human ear is not the loudest possible
sound. The sound inteJlsity in the neighborhood of a space rocket during its
2
lift-off exceeds 108 W/m , which is equivalent to 200 dB.

5 The unit decibel is not unique to acoustics. It is used extensively in electronics


and communication epgineering - in fact in all fields where the range of
magnitude of a quantity is large.
12 Chapter 1

Expressing the loudness of sound in dB is far more convenient


than expressing it in W/m2. For instance, it is easier to express the
loudness of sound as 94 dB instead of 0.0025 W/m2 (see Examples
1.1 and 1.2). This fact may be further appreciated with reference to
Table 1.3 Perception of Change in Figure 1.10 which gives the sound intensity and sound intensity levels
Sound Intensity Levels of common environmental sounds.
Obtaining convenient numbers is not the only reason for the use
Change in level Human perception of the decibel scale. The decibel scale also corresponds more directly
(dB) to the ear's perception of loudness. For instance, a change of 1 dB
in sound intensity level is hardly perceived by the human ear. That
is why a sound intensity level is expressed in a whole number since
1 Imperceptible expressing it as a decimal number indicates an unnecessary precision.
3 Just perceptible Thus, a sound intensity level of 50.6 dB may be approximated to 51
5 Clearly noticeable dB witpout any loss of accuracy. The minimum change in sound
10 Substantial change intensity level that is just perceptible is 3 dB . A 5 dB change is quite
noticeable, and a 10 dB change is substantial, Table 1.3.

Sound lntenelty
Sound preeeur: Sound lntenelty level or e;ound Nolee In the
(Pa) (W J m2) preeeure level environment
(dB)

63.2 10 - 130 -
Threehold of pain
20 1 120
- Near a jet aircraft at take-off

6.32 0.1 110 Riveting machine

2.0 0.01 f-,- 100


- Pneumatic hammer

0.632 0.001 90
-
Dleeel truck at (15 m) 50 ft
0.2 0.0001 80 Shouting at 1 m (3 ft)

0.0632 0.00,001 - 70 -
Buey office

0.02 0.000,001 1-- 60


- Convereatlonal epuch at 1 m (3 ft)

0.00632 0.0,000,001 50 - Quiet urllan area during daytime

0.002 0.00,000,001 1-- 40 - Quiet urllan area at night

0.000632 0.000,000,001 30
- Quiet eullurllan area at night

0.0002 o.o.ooo.ooo.oo1 20 - Quiet countryelde

0.000632 0.00,000,000,001 - 10 - Human llreathlne

0.00002 0.000,000,000,001 L.......- 0- Threehold of audllllllty

1.10 Some typical noises in our environment and their sound intensifies and sound
intensity levels.
The Physics of Sound 13

Example 1.1 Sound Intensity from Sound Intensity Level

Determine the sound intensity (I) of a sound whose sound intensity level (IL) is 83 dB.

Solution: IL = 83. From Equation (1.2),

I I
83 = 10 log or, 8.3 = log or,
-12
--

Iref
10
Thus, I = 10-37 = 0.0001995 W/m2.

Example 1.2 Sound Intensity Level from Sound Intensity

Determine the sound intensity level (IL) of a sound whose intensity is (i) 0.0025 W/m2, and (ii) 4.5 W/
m2 .

Solution: (i) From Equation 1.2, IL = 10 lo. (0.0025/ l 0-12) = 10 log (2.5 x 109) = 94 dB.

(ii) IL = 10 log (4.5/10-12) = 10 log (4.5 x 10 2) = 126.53 dB, or 127 dB.

1.4.1 Sound Intensity Level and Sound Pressure Level

Highly complex instrumentation is required to measure sound


intensity, and thereby sound intensity level. On the other hand, the
measurement of sound pressure is fairly simple since the microphones
used for this purpose detect sound pressure, not sound intensity.
Therefore in practice, we measure sound pressure, not sound intensity.
This does not cause any real problem since the sound pressure and
In architectural acoustics, sound pressure sound intensity are related quantities. Incidentally, our ear drum,
level, SPL, and sound intensity level, IL, which behaves similar to the microphone, also detects sound pressure.
are equal. These two quantities are Sound pressure is measured through an instrument called the
therefore used interchangeably. However, sound level meter (described in Section 2.1). A sound level meter
the use of sound pressure level is more does not give a reading of sound pressure but that of sound pressure
common, since the sound level meters level (S PL). Sound pre ssure level i s simply a logarithmic
respond to the sound pressure. representation of sound pressure, in exactly the same way as sound
intensity level is of sound intensity (see Appendix B).
Thus, in practice, it is the SPL that is measured, not the sound
intensity level. Fortunately, in architectural acoustics, the SPL is
(approximately) equal to the sound intensity level. Since it is the
SPL that is measured, noise levels are usually quoted in terms of the
SPL. However, because of their equivalence, the SPL and the sound
intensity level will be used synonymously in this text and will be
referred to simply as the sound level.

1.4.2 Sound Power Level

The pressure or intensity in a sound wave is obviously related to the


power of the source. The power of a sound source is measured in
watts. The sound power of a faint whisper is approximately 1 o-3 W,
14 Chapter 1

and that of a space rocket at take-off is 108 W. This large range is


compressed by using sound power level in place of sound power.
The sound power level (PWL) of a source is given by the
following logarithmic ratio, and is measured in dB:

PWL = lOlog ( 1 .3)


Wref

where, W is the sound power of a source in watts, and W ref is the


reference sound power= 10" 12 w.

1.5 COMBINING SOUND It is often necessary to determine the overall sound level from several
sound level measurements. For example, we may need to know the
LEVELS combined sound level in an environment where the sound levels of
individual noise sources are known. The converse of the above would
be to determine the resultant sound level in an environment when
one or more noise sources of known sound levels are removed.
The addition or subtraction of sound levels cannot be done in
the same way as the addition or subtraction of ordinary numbers
(arithmetic addition and subtraction). Sound levels, being logarithmic
quantities, must be added or subtracted logarithmically. For instance,
the resultant sound level of two noise sources, each producing a sound
level of 80 dB, is not 160 dB. In fact, as explained below, the
combined sound level of these two sources is 83 dB. In other words,
80 dB+ 80 dB= 83 dB.

1.5.1 Adding Sound Levels

Two procedures exist to obtain the sum of two sound levels: the
exact procedure and the approximate procedure. The exact procedure
is explained in Appendix B. According to this procedure, we first
determine the sound intensities corresponding to the given sound
levels. These sound intensities are then added (arithmetically), and
the resultant sound level determined by calculating the logarithm of
the sum of sound intensities.
The approximate procedure is more commonly used because of
its simplicity. It is a two-step procedure as given below:
Table 1.4 Approximate Decibel
Step 1: Determine the difference between the two sound
Addition levels to be added.

Step 2 : Determine the amount to be added to the higher
Difference between Decibels to be level from Table 1.4. This gives the resultant sound level.
two levels to be added to higher
added (dB) level For example, let us determine the resultant sound level of 80 dB
and 84 dB. The difference between the two levels is 4 dB. Therefore,
from Table 1.41 we will add 2 dB to the 84 dB level to give us the
0 or 1 3
resultant sound level of 86 dB. Thus, 80 dB+ 84 dB= 86 dB. If the
2 to 4 2
two sound levels are equal (a difference of zero), we will add 3 dB to
5 to 9 1
the sound level to obtain the resultant sound level. Thus, 80 dB+ 80
10 or more 0
dB= 83 dB.
The Physics of Sound 15

From Table 1.4, we observe that if two sound levels differ by 10


dB or more, we add nothing to the higher level to obtain the resultant
level. In this case, the louder sound determines the overall sound
level entirely. Thus, 80 dB + 90 dB = 90 dB .
The above procedure can also be used to add a number of sound
levels by successively adding two levels, as shown in Figure 1.11
and Example 1.3.
However, if there are a number of sources of identical sound
levels, their addition can be simplified through the use of Table 1.5.
Thus, the sound level of 10 sources, each with an individual level of
70 dB is 80 dB ; 100 such sources will give a total level of 90 dB (see
also Example 1.4).

70 d6 + 72 d6 + 75 d6 + 00 d6

1.11 The sum of a number of sound


V 78

pressure levels may be obtained by 82 d6 THI515 THE TOTAL


adding two levels at a time. SOUND LEVEL

Table 1.5 Summation of Levels for


Multiple Sources

Number of Decibels to be added


identical sources to single level

2 3 (a) Nolee from 1 (1:1) Nolee from 2 Identical


3 5 vacuum cleaner = x d6 vacuum cleaner& (x + :3) d6
=

4 6
5 7
6 8
7 8
8 9
10 10
15 12
20 13
50 17
100 20
(c) Nolee from 4 Identical vacuum cleanere = (x + 6) d6

(d) Nolee from 10 Identical vacuum cleanere = (x + 10) d6


16 Chapter 1

1.5.2 Ten Times the Logarithm (10 log )

Ten times the logarithm of a quantity [10 1og(x)] is so pervasive in


architectural acoustics that it is difficult to overemphasize its
importance. For example the reader will observe that the values in
the second column of Table 1.5 are obtained simply from the
Table 1.6 Subtraction of Sound expression 10 log(n), where n is the number of identical sources.
Levels Thus if n = 2, 10 log(2) = 3, since log(2) = 0.3; if n = 10, 10 log(10)
= 10.

Decibels in dB Decibels to be
between the overall subtracted from
level and the level the (earlier) 1.5.3 Subtracting Sound Levels
of source to be overall level
subtracted to obtain the The subtraction of sound levels can be obtained from Table 1.4 by
resultant level reversing the steps needed for addition. However, the work is
simplified through the use of Table 1.6. For example, assume that
the overall sound level in a space is 85 dB and we wish to eliminate
0 10 or more a source whose level is 80 dB. What will be the resultant sound
1 7 level in the space?. In other words, what is 85 dB - 80 dB? From
2 4 Table 1.6, the resulting sound level is 85 - 2 = 83 dB (see also Example
3 3 1.5).
4 or 5 2 Although the answers obtained from the use of Tables 1.4
6 to 9 1 through 1.6 are approximate, they are reasonably accurate for all
10 or more 0 architectural acoustics work since the error in the final answer seldom
exceeds 1 dB, which is insignificant from the point of view of human
perception (see Table 1.3).

Example 1.3 Overall Sound Level

Determine the overall sound level of a noise whose octave band levels are given in the following table.

Solution: The octave band noise levels can be added as shown in the adjacent diagram, giving an
overall sound level of 97 dB.

Octave (Hz)

SPL(dB)
63

95
125

93
250

70
500

70
1,000

70
2,000

60
4,000

62
8,000

60
95
93
> 97
70 >
70
73
>
97 >
70
60
> 70 )
.. 71
97

62 > 64
60

Example 1.4 Effect of Adding Noise Sources

The existing noise level in a manufacturing space is 90 dB. It is proposed to add 10 machines in this
space, each producing 80 dB. Will the resulting noise level be excessive as compared to the existing
level?

Solution: From Table 1.5, the combined noise level of 10 machines, each producing 80 dB, is 80 + 10
= 90 dB . Since the existing noise level in the space is 90 dB, the final noise level in the space is 93 dB
(from Table 1.5). Thus, the additional! 0 machines increase the noise level by 3 dB. From Table 1.3, an
increase of 3 dB is small since it is only just perceptible.
The Physics of Sound 17

Example 1.5 Effect of Deleting Noise Sources

The noise level on an urban road (Road A) during rush hour is 90 dB, which has been determined to be
excessive. As a solution to the noise problem, it has been proposed to build another road (Road B) in
the neighborhood, which will reduce the traffic on the existing road by 50 percent. What is your
assessment of this proposal from the point of view of noise reduction?

Solution: Since the construction of Road B will reduce traffic on road A by half, the noise reduction
will be only 3 dB, giving a noise level of 87 dB on Road A. From Table 1 .3, we note that a minimum
of 5 dB reduction is necessary to be perceived as a noticeable difference. Thus, constructing Road B,
to carry 50 percent of the traffic, will not make a noticeable difference to noise level on Road A.
From Table 1.6, we observe that to achieve a 5 dB reduction, the traffic on Road A must be reduced
to one-third of its original value. In other words, Road B must carry two-thirds of the traffic. With two
thirds traffic on Road B, the noise level on Road B will be 88 dB, which is not much less than 90 dB on
Road A. Therefore, the proposal is acoustically not justified.

SOUND
1.6 One of the ways in which acousticians classify noise sources is by
the size of the source relative to the distance at which the effect of
ATTENU ATION BY the source is considered. According to this classification a noise
DISTANCE source is classified as: (i) a point source, or (ii) a line source.
Theoretically, a third classification could be an area source. In
practice, however, only the point source and line source are important.

1.6.1 Point Source and Inverse Square Law

A point source is one whose dimensions are much smaller than the
distance at which we wish to determine the sound level produced by
it. As a rough guide, at a distance greater than or equal to five times
the largest dimension of the source, the source behaves as a point
source. Thus, if the largest dimension of a sound source is 2 ft, it
will behave as a point source at a distance of 10 ft or greater from
the source. Indeed the greater the distance from the source, the closer
the behavior of the source to a point source. In fact, all sound sources,
regardless of size, behave as point sources at sufficient distance away
from the source.
More precisely, a point source is one which obeys the inverse
square law. To understand the inverse square law, imagine a point
source of sound which radiates sound equally in all directions. We
Arr:a of ephr:rr:
will assume that the source is hung in space (or is located in an
= 411:R2

anechoic chamber) so that no reflected sound is present. Assume


that the acoustic power of source is W watts. Let us now draw an
imaginary sphere of radius R around the source with the source as
1.12 Sound intensity due to a source
the center, Figure 1.12. Since the area of this sphere is 41tR2, the
of known power. intensity of sound6 on the surface of the sphere (I) is given by:

6 Remember (from Section 1.4) that the intensity of sound is equal to the
sound power per unit area, i.e., I (sound power)/area.
=
18 Chapter 1

I = (1 .4)

The above expression gives sound intensity at distance R from


Sound lntenelty the source. Since quantities W and 47t are constants, we see from
= 0.25 1 this expression that sound intensity is inversely proportional to
distance squared - hence the name inverse square law. Thus,
according to this law, if the distance from the source is doubled, the
intensity at the new point (which is at twice the distance from the
Sound previous point) is one-quarter times its intensity at the previous point.
The reason is that at twice the distance, the sound energy (power)
passes through an area that is four times the area through which it
passed at half the distance, Figure 1 . 13.
B y the same argument, if the distance from the source i s halved,
its intensity at this new point is four times the intensity at the previous
point. In other words, with reference to Figure 1 . 14, if the intensity
at point P, which is at distance 4 m from the source is I, the intensity
at point Q, a distance of 2 m from the source, is 41. If the sound
intensity at Q is four times that at A, what can we say about the
sound level (or SPL) at Q, as compared to the sound level at P?
From Table 1 .5, we observe that four identical sources increase
the sound level by 6 dB. Thus, the sound level at Q is 6 dB greater
1.13 The area through which a given than that at P. Thus, if the sound level at P is 60 dB, that at Q will be
amount of sound power flows is four 66 dB. Similarly, the sound level at point R, which is 8 m away
from the source, will be 6 dB less than that at P, i.e. 54 dB, because
times the area through which the same the intensity at R is 0.251.
powerflows at twice the distance from In conclusion, therefore, the inverse square law states that in a
the source. space where there are no reflections (free field), sound level due to a
point source falls by 6 dB for every doubling of distance, and
increases by 6 dB for every halving of distance.

ANECHOIC SPACE


'"""' lty = ..
7
l""" olty iI = (IL + 6) .0
2 ..

lntenelty = 0.251 R p lntenelty = 1


lntenelty hwt:l = (IL - 6) dB Intensity lllYt:l = (IL) dB

1.14 Sound intensifies and sound intensity levels in a free .field. The sound source
has been assumed to be non-directional, that is, it radiates equally in all directions.
The Physics of Sound 19

1.6.2 Sound Levels at Different Distances from the Source

Often the manufacturers provide the noise level of their equipment


in terms of the sound pressure level produced by the equipment in a
free field (anechoic chamber) at some distance away from the
equipment - typically at a distance of 1 m or 1 ft. If we know the
sound level (sound pressure level or the sound intensity level) of a
sound source at distance D from it, we can determine the sound
level at any other distance, 02, by using the inverse square law.
By a simple mathematical manipulation, it can be shown that if
the sound level due to a source at distance D 1 is SPL 1 , then at distance
D2 from the source, SPL2, is7 :

SPLz = SPL1 - 20 log I ( 1 .5)

Equation ( 1 .5) can also be written as follows,which is a more easily


memorized format.

SPL2 + 20 log D2 = SPL1 + 20 log D1 ( 1 .6)

The use of the above equation is illustrated in Example 1 .6.

Example 1.6 The Use of Inverse Square Law

The sound level at lm away from a source in an open space (no reflections) is 80 dB. Calculate the
sound level due to the same source at a distance of 6 m from it. (Assume that the source is non
directional, that is, it radiates equally in all directions).

Solution: From Equation ( 1 .6),

SPL6 + 20 log(6) = SPL 1 + 20 log( l ). Thus, SPL6 = 80 - 15.6 = 64.4 dB, say 64 dB.

7 Equations (1.5) and (1.6) assume that the source is nondirectional, that is, it
radiates equally in all directions. If not, the source's directivity must be considered.
20 Chapter 1

1.6.3 Line Source and Inverse Law

A line source is one which contains a large number of point sources


spread out along a line. In practice, free flowing highway traffic
behaves as a line source. While the sound spreads spherically around
a point source as shown in Figure 1 . 13, it spreads cylindrically around
a line source.
Considering the cylindrical spreading of sound, it can be shown
that the sound intensity due to a line source is inversely proportional
to the distance from the source, a law called the inverse law. Thus,
if the intensity at distance D from a line source is I, the intensity due
to the same source at 20 away will be 0.5I. In terms of sound intensity
levels, it implies a 3 dB reduction for every doubling of distance, or
a 3 dB increase for every halving of distance from a line source (see
Example 1 .7) .

Example 1.7 The Use of Inverse Law

The sound level due to free-flowing traffic on a highway is 88 dB at


point A, which is 1 5 m (50 ft) away from the centerline of the highway.
How much will the sound level be at B , which is 30 m ( 1 00 ft) from
the highway?

Solution: Since the sound level due to a linear source decreases by


3 dB for every doubling of distance, the sound level at B = 88 - 3 =
85 dB.

1.7 SOUND FIELDS In describing the inverse square law, we assumed the source to be
suspended in space so that there were no reflections from the ground
or other nearby surfaces. A space (or sound field) in which all sound
comes directly from the source (with complete absence of any
reflected sound) is called a free field, implying 'freedom' from
reflections. From Section 1 .6. 1 , we see that in a free field, sound
level drops by 6 dB for every doubling of distance.
In practice, a free field is obtained in a room specially constructed
for this purpose, called the anechoic chamber. In an anechoic
chamber all walls, the ceiling, and the floor are covered with wedge
shaped sound absorbers so that all sound that falls on them is
absorbed, Figure 1 . 1 5 . The absorbing material on the floor is
protected by a structural wire mesh above it. The structural mesh is
strong enough to support the loads of equipment and technicians,
but allows virtually all sound to go through it.
Thus, we observe that in an ideal free field all sound is direct
sound. There is no reflected sound at all. By contrast, a sound field
in which all sound consists of reflected sound is called a reverberant
field. In a (true) reverberant field, the sound level is constant
throughout the space.
The Physics of Sound 21

1.15 Anechoic chamber. Courtesy of 3M Corporation, St. Paul, Minnesota, with permission.

At any point in an actual room, the sound consists of both the


direct sound and the reflected sound. Therefore, the sound field at a
point lies somewhere between the two extreme conditions: free field
and reverberant field conditions. Close to the source, the level of
direct sound is high. In this region, the direct sound dominates, and
the sound level obeys the inverse square law - 6 dB reduction for
each doubling of distance.
22 Chapter 1

The exception to the above behavior occurs in a region that is


extremely near the source. In this region, the sound level drops more
rapidly - by nearly 1 2 dB for each doubling of distance. This region
(a distance less than five times the dimension of the source), is called
the near field. The remaining region is, therefore, called the far
field.
We are usually interested in the far field region, since the near
field region is too close to the source to be of practical interest. The
far field region is further subdivided into free field and reverberant
field. The change from free field condition to reverberant field
condition is gradual.
The above discussion is summarized in Figure 1 . 16, which shows
the variation of sound level in a room with respect to the distance
from the source. In other words, if we draw a straight line between
a listener in a room and the sound source, and measure the sound
levels at different points along this line, the sound levels will be as
shown in Figure 1 . 1 6.
Note that at a large distance from the source, the sound level
becomes constant, showing that at these locations the reflected sound
dominates. Close to the source, the direct sound dominates, and in
this region the sound level decreases by 6 dB for every doubling of
distance. Still closer to the source (in the near field region), the
sound level drop is steep.
In an anechoic chamber, the reverberant sound level is zero.
Consequently, the change in sound level with distance will be
represented by the (broken) straight line throughout the room - a 6
dB drop for every doubling of distance.
The sound level corresponding to the horizontal line in Figure
1 . 1 6 is the reverberant (reflected) sound level, and depends on the
amount of absorption present in the room. The greater the amount
of absorption in the room, the lower the reverberant sound level line,
and vice versa. This is shown in Figure 1 . 1 7, which compares the
variation of sound levels with distance in two rooms - one with a
greater amount of sound absorption than the other.

Near
field Far field
Sound lvl variation In a
r- room with a emall amount
of aueorptlon
Reverberant field

Sound on thle lln


.
...

.. ___

decreaee l7y 6 dB for ry


douulln!! of dletanc from

t h eourc
Dle;tance from e;ource

..
.

.
. ..
.

Die;tance from e;ou rce


1.17 Variation of sound level with distance in
1.16 Near field, far field, free field and reverberant field in two rooms - one with a greater amount of
a room. absorption than the other.
The Physics of Sound 23

REVIEW OF LOGARITHMS

Any number can be expressed in exponential form - in the form of 1 ox, where x is called the exponent
of 1 0. For instance, 1 00 can be expressed as 1 02, 1 ,000 as 103 , and 1 0,000 as 1 04 , and so on. Not only
can we express numbers such as 1 ,000 or 10,000 in exponential format, we can also express numbers
such as 264 or 4,669 jn the form of 10x. Using a modem calculator, it is fairly easy to see that 264 =
1 o2.422, and 4669 = 10 669. Before the advent of calculators, however, expressing numbers in exponential
formats could be done only through the use of logarithm tables, referred to as log tables.
A logarithm is simply an operator that changes a given number into another number. By definition,
log(1 0)x = x. Thus, log(100) = 2, log( l ,OOO) = 3, log ( 10,000) = 4, and so on. Similarly, log(264) =
1
2.4216 and log(4,669) = 3.6692, since 102 .42 6 = 264, and 1 0H>692 = 4,669.
The real advantage of log tables was not in expressing numbers in exponential format, but in the
multiplication and division of large numbers. Log tables simplified the multiplication and division of
large numbers by converting the multiplication into addition of numbers, and the division of numbers
into subtraction.
For example, let us determine the product: 264.59 x 4,669.72. Using a log table, we would find that
log(264.59) = 2.4226, and log(4669.72) = 3.66932. We now add 2.4226 and 3.6693, giving us 6.091 9.
Using log tables once again, we would find that 1 ,235,663 is the number whose logarithm is 6.09 19. In
other words, log(l ,235,663) = 6.09 19, or (264.59 x 4,669.72) is (approximately) equal to 1 ,235,663. In
doing this multiplication with the help of log tables, we simply added numbers.
Fortunately, we do not have to learn the use of log tables today. All that we need to learn is the use
of a calculator in determining the logarithm of a number. Scientific calculators have a function "log".
This function gives us the logarithm of a number in a one-step operation. The reader is advised to
practice determining the logarithms of a few numbers using a scientific calculator. As an exercise,
determine: log(2,399); log(345,95) ; log ( 1 3 .33); log( 109.3). From thse exercises, the reader will note
that:

log(1) = 0
log( a number between 1 and 1 0) = a number that lies between 0 and 1
log(a number between 10 and 1 00) = a number that lies between 1 and 2
log(a number between 100 and 1 ,000) = a number that lies between 2 and 3
log(a number between 1 ,000 and 1 0,000) = a number that lies between 3 and 4, . .. etc.

Since ( 10X)( 10Y) = w<x+y) ' it follows that:

log(a x b) = log(a) + log(b)

Similarly,

log(alb) = log(a) - log(b), and

log(ab) = b'log(a) - by definition.

The above formulas are the only ones needed to solve any problem related to logarithms. A few
logarithms worth remembering are: log(2) = 0.3, log (3) = 0.48 and log 5 = 0.7. The reader should now
try to determine the following logarithms, using the above formulas, but not the calculator.

log(4); log(6); log( 1 5); log(20); log (60); log(200) ; log(800); log(2,000)

If the lofarithm of a number is given, that number can e determined. For instance, if log(n) = 3,
then n = 1 0 = 1 ,000. Similarly, if log(n) = 2.37 , n = 10 2 7 = 234.4. Using a calculator, the reader
should determine the value of n, given that log(n) is:
0.55, 1 .377, 3 .999
Sound
Measurement
and Hearing

Several acoustical quantities are routinely measured in an acoustic


laboratory. In practice, however, the most commonly measured
quantity is the sound pressure level of sound sources and noise
environments. This is done through a sound level meter whose
characteristics are briefly discussed in this chapter.
Since the human ear is the ultimate judge of all sounds, a sound
level meter must have reasonable correspondence with the hearing
characteristics of the ear. One important characteristic of hearing -
the ear's perception of sound pressure level changes- was discussed
in Chapter 1 (see Table 1.3). Other important characteristics of the
ear and related properties are described in this chapter.

2.1 SOUND LEVEL As stated in Section 1.4.1, the instrument used for measuring sound
pressure level (or sound intensity level) is called the sound level
METER meter. Various types of sound level meters are available depending
on their precision and versatility (which allow other accessories to
be attached to it). However, all of them consist of the following four
basic components: (i) a microphone, (ii) an input amplifier, (iii) an
output amplifier and (iv) a read-out device.
The microphone, through its diaphragm, converts the fluctuating
air pressure produced by a sound source into a fluctuating electrical
charge. The input amplifier converts this charge into a voltage signal
that is processed by the electronic circuitry of the instrument, and
converted into sound pressure level. The output amplifier increases
the extremely weak sound signal to a level that is adequate to display

25
26 Chapter 2

the signal on the meter's read-out panel. Most modem sound level
meters give sound pressure level reading in a digital format (using
light emitting diode, LED, or liquid crystal technology), as shown in
Figure 2.1. Some meters, however, include both a digital read-out
as well as analog (needle) read-out.
The sound level in an environment may either be virtually
constant or vary with time. For example, the noise in a transformer
room, or the noise created by air conditioning systems, electric motor
noise, etc., consists of a continuous noise of an almost constant leveL
On the other hand, noise in a manufacturing space, surface grinding,
welding, traffic noise, etc., varies with time.
The meter's display of the sound pressure level, whether digital
or the traditional needle type, cannot respond instantaneously to rapid
fluctuations of sound pressure levels. The meter's display mechanism
has a certain response time. Therefore, a typical sound level meter
is provided with three averaging speeds: slow, fast and impulse.
In the slow mode, the meter averages the sound level over one
second. In the fast mode, the averaging period is 1/8 second (125
milliseconds). The slow mode is most commonly used, particularly
if the noise level is changing rapidly. If the meter is set to the fast
mode for rapidly changing noise levels, the meter reading may vary
too fast for the eye to follow the changes. Thus, an assessment of
the average noise level may not be obtained.

2.1 Sound level meter. Courtesy of CEL Instruments,


Milford, North Hampshire, with permission.
Sound Measurement and Hearing 27

The fast mode is used where we wish to obtain the peak levels in
a variable noise environment. The impulse mode, with an averaging
period of 35 milliseconds, is used to monitor impulse noises such as
noise from pneumatic drills, automatic presses and guillotines,
riveting machines, etc.
To obtain sound levels within octave bands or one-third octave
bands, a filter mechanism is either connected to the sound level meter
externally, or built into the meter. A filter mechanism consists of a
number of filters; each filter allows only the sound contained within
that particular frequency band to pass through, cutting off all sound
lying outside that band.
A sound level meter without the filter mechanism gives the
overall sound level (in frequencies ranging from 20 Hz to 20 kHz).
The overall sound level is also obtained when the filter mechanism
is connected but not switched on. On the other hand, if the filter
mechanism is switched on for a particular band, the meter reading
gives the sound level contained within that band. The filters are
switched manually so that successive sound pressure level readings
can be obtained for each octave or one-third octave band. Readings
so obtained give the breakdown (analysis) of sound levels in various
octave or one-third octave bands, commonly referred to as octave or
one-third octave analysis.

2.1.1 Real-time Analysis

If the sound (or noise) environment is constant over time, both in


level as well as in frequency composition, one can use the sound
level meter, such as that of Figure 2. 1, and obtain the fequency-level
relationship of sound by sequentially switching the meter to different
octave or one-third octave bands. The above analysis must be carried
out at the location of noise.
However, if the level and frequency composition varies over
time, the above procedure cannot be followed. In such a case, the
traditional approach was to tape record the sound on a tape recorder
(specially made for this purpose) and to analyze the sound
subsequently, in the office or the laboratory.
The introduction of real time analyzers made tape recording
unnecessary, since the octave or one-third octave analysis can be
done instantaneously - that is, in real time. Such an analyzer is
called a real-time analyzer. Thus, a real-time analyzer, which looks
similar to a sound level meter, gives sound levels in various octave
or one-third octave bands. One such analyzer is shown in Figure
2.2 A real-time sound analyzer.
2.2.
Courtesy of CEL Instruments, A real-time analyzer is used not only for frequency analysis but
Milford, New Hampshire, with also for reverberation time1 measurements, in which the rate of decay
permission. of sound in an enclosure with respect to time is required. The
reverberation time of an enclosure usually varies with frequency.
Therefore, reverberation time is measured for different octave bands.

1 See Section 1 0.3 for a discussion of reverberation time.


28 Chapter 2

2.2 THE EAR'S The sound level measurements obtained from a sound level meter
refer to the physical phenomenon of sound. These measurements
SENSITIVITY do not bear a simple relationship with the sensation of loudness.
This lack of simplicity results from the human ear's unequal
sensitivity at various frequencies. For example,a given sound level,
of say 4 0 dB,at 1 00Hz does not sound as loud as 40dB at 1,000 Hz.
The unequal sensitivity of the human ear is shown in the curves
of Figure 2.3, which are known as equal loudness contours. These
contours have been obtained from extensive experimental data in
which the subjects were asked to compare the loudness of sounds at
different frequencies.
Each contour gives the sound level necessary to produce the
same loudness sensation at different frequencies. Contours are
labeled as 1 0, 20, 3 0,etc. - numbers that refer to the corresponding
sound levels at 1 kHz. Thus,in labeling these contours,the frequency
of 1 kHz has been taken as the reference frequency. A contour labeled
4 0 passes through the 4 0 dB point at 1 kHz frequency. Similarly, the
contour labeled 1 20 passes through 1 20 dB at 1 kHz frequency.
An arbitrary unit called the phon has been assigned to these
numbers, so that the contour labeled as 4 0 is a 4 0 phon contour, that
labeled as 5 0 is a 5 0 phon contour, and so on. Thus, we see that
phon is the unit of loudness. It is important to note the distinction
between phon and decibel: the decibel is the unit of (sound) stimulus
while phon is the unit of (sound) sensation.
Equal loudness contours indicate that the ear is most sensitive
between 4 to 5 kHz- the region where most information is contained
in human speech2 The sensitivity decreases in the low and high
frequency regions,the decrease in sensitivity being most pronounced
in the low frequency region. In fact, the ear is relatively deaf in
frequencies from 20 to 100 Hz. For example, examinipg the 3 0
phon contour,we see that to produce the same loudness sensation of
3 0 phons,we need a sound level of 87 dB at 20Hz,45 dB at 1 00Hz,
3 0 dB at 1 kHz, 23 dB at 4 kHz, and 4 0 dB at 1 0 kHz.
The contours tend to become flatter with greater loudness,
implying that the ear's sensitivity becomes more uniform at higher
sound leveis. The lowest (dashed) contour is the threshold of
audibility contour and the uppermost contour refers to the threshold
of pain.

2.2.1 Weighting Networks in Sound Level Meters

Since the human ear is the ultimate judge of all sounds, efforts have
been made to develop a sound level meter that will simulate the
equal loudness contours. However, this has not been done because,
first, the complexity of the equal loudness contours is so great that
such a meter is economically unviable. Second, the equal loudness
contours have been obtained with reference to pure tones although
real life sounds have complex spectra. The correspondence between
loudness perception of pure tones and complex sounds is not well
established.
A partial cloning of the ear's characteristics is obtained through
frequency weighting networks provided in a sound level meter. A
frequency weighting network simulates the unequal sensitivity of

2 Speech intelligibility is determined mainly by consonants, which are high


frequency components of sound and lie within the 2 kHz to 4 kHz octave bands
(see Figure 1 .8).
Sound Measurement and Hearing 29

Thresh o Id of pa In
I
120 " r--
Ll_ 120

'-r-... r-.
r--. -..... I
-V
1I'-,
."'-..
"' f'-
r--
r-- I J
100 -
-...... 1-"
'\[',.'r-. !'---.. 90 !'-... I '-...J
....... -
..
8;;...!'-... ...YI
--

80 r'-r-... !'---... - ......


'\.

Sound L\" r'-r--, 2Q !'--.1---V V


I
\.'-"
pre66ure
60
\t\'\.1'-r---
"

-
60 r-...r-- .. VVI/
"'r--, 50 r---.r-f.--'. Vv
- '-'
level (dB)
'\.
r---.I-
. VV
40
'"t\r--
0 40 V [\I
r-..; ""-""'r---... 30 r-f.--' v
v 1\J
1-V I
20 -...!'-..
.. r-- 20 rv,,

r--
Threehold of / r- 10r--.r--.1-./ I -I
/'
.../
1---r--
-- ......
...... -
audli:llllty curve
-- .....
0
Ill I
.. _ ...

20 50 100 500 1,000 5,000 10,000


2.3 Equal loudness Frequency (Hz)
contours.

the human ear at different frequencies. More specifically, a weighting


network reduces the sound pressure level at low frequencies to
compensate for the low sensitivity of the human ear at these
frequencies. Three such weighting networks are usually provided in
a sound level meter, referred to as A, B and C, Figure 2.4.
Ignoring minor details, weighting network A is an inversion of
the 4 0 phon contour, and weighting network B, that of the 70 phon
contour. Weighting network C, which is virtually flat, is an inverted

20
Center frequency A-weighting
A
of octave (Hz) (dB) ] 0 c
!--'-
.......
/

63 -26

:::1
Ill
-10 B
/
/
Ill V
1 25 -16
I:>.. -20 /
V
250 -9 -.::5
s:
500 -3
0 -30
:::1
1 ,000 0 Ill

2,000 +1 -40
+S
4,000 +1 (q

8,000 -1 -50
63 250 1,000 4,000
125 500 2,000 8,000
Frequency (Hz)

2.4 Standardizedfrequency weighting curves.


30 Chapter 2

version of the 100 phon contour. When originally developed, these


weighting networks were meant to be used for different noise levels:
A for low sound levels, B for mid-range levels and C for really loud
sounds.
However, observations have revealed that B and C networks do
not correspond well with loudness perceptions. This, once again, is
believed to be due to the lack of correspondence between the results
obtained for pure tones and those for complex sounds. Therefore,
almost all current noise evaluation is done on the A-weighting
network, because of its relatively good correlation with human
response. Weighting networks B and C are mere historical facts
now.
When a sound level is measured with A-weighting switched on,
it is designated as dBA or dB(A), to distinguish it from the unweighted
level which is given in dB. The weighting provided in different
octaves in the A-weighting network is given in Figure 2.4. If the
unweighted octave band levels are given, it is fairly easy to compute
the dBA levels by using the techniques described in Section 1.5.
The dBA levels are usually lower than dB levels.
Since the A-weighting curve is virtually flat for frequencies above
1 kHz, the difference between dB and dBA readings reveals the extent
of low frequency distribution in a noise environment. If the difference
between the dB and dBA levels is large (greater than 10 dB), it
indicates a greater contribution of low frequency component in the
noise, as illustrated in Example 2.1. On the other hand, if the low
frequency levels in a sound are relatively small, the difference
between dB and dBA levels is also small.
Note that although dBA level is a better indicator of the noise
environment than the unweighted overall noise level (in dB), it does
not give us as detailed a picture of the noise environment as that
provided by an octave or one-third octave analysis. However, being
a single number index dBA level is convenient to measure and use.

Example 2.1 dBA levels from Octave Band Levels

Determine the overall dBA sound level for the noise source of Example 1.3.

Solution: The unweighted noise levels are shown in row 2 and the A-weighted levels in row 3 of the
following table. The overall A-weighted level is obtained by adding the levels given in row 3, which
works out to 79 dBA. The overall unweighted level is 97 dB, see Example 1.3.

Octave (Hz) 63 125 250 500 1,000 2,000 4,000 8,000


69
77
> 78
SPL (dB) 95 93 70 70 70 60 62 60 78
61
> 68
67
SPL (dBA) 69 77 61 67 70 61 63 59 79dBA
70
> 71
61
72
63
59
> 65
Sound Measurement and Hearing 31

2.3 THE HAAS EFFECT The nonuniform sensitivity of the human ear is not the only property
of the ear that is of interest to us. A few other characteristics of
hearing that are of interest are discussed in the following sections.
One of these characteristics is the ability of the ear to integrate
all (identical) sounds that follow within brief intervals of each other,
as if these sounds were not separate sounds but one sound. The
ear's property to integrate sounds was first discovered by Helmut
Haas through experiments conducted on a large number of listeners.
In this experiment, the listener was set equidistant from two
loudspeakers, loudspeakers A and B in an anechoic chamber, so that
each loudspeaker subtended an angle of 45 at the listener, Figure
2. 5.
Haas's experimental set-up included a time delay mechanism in
the circuitry from the sound generator to loudspeaker B, so that the
sound at the listener from loudspeaker B could be delayed with respect
to that from loudspeaker A by varying amounts. In addition to varying
the delay time, the level of sound from loudspeaker B with respect
to the sound level from A could also be varied. Thus, in Haas's
experiments, the listener received a sound from loudspeaker A and
an identical sound from loudspeaker B - delayed by a time interval
and of different sound level.
Haas discovered that when the delay time was zero and the level
of both sounds was the same, i.e., when both sounds arrived at the
listener's ears at the same time and were of equal loudness, the listener
perceived them as one sound coming from an imaginary loudspeaker
C located right in front of him. In other words, the ear integrated
both sounds into one sound, and had the illusion of receiving the
sound from a source equidistant from the two sources.
The integration effect occurs even if the sound from B is delayed
with respect to that from A, provided the delay is less than 40
milliseconds, and the level of sound from B is not more than 10 dB
above that from A. Stated differently, if the delay between two sounds
is up to 40 milliseconds and if the delayed sound is no more than 10

Delay
mechanism

Sound
generator
rb. c
:.-...:-:.
Imaginary speaker

'

o'
'
'
'
'
'
'
, -
, -
'
'

ANECHOIC
CHAMBER

2.5 Experimental set-up for the Haas effect.


32 Chapter 2

dB above the level of the earlier sound, the ear does two things: (i)
it perceives both sounds as one sound, adding their loudness, and
(ii) it thinks that all the sound is coming from A - the loudspeaker
from which the sound came to the listener first.
In other words, the sound that arrives first at the listener
establishes the source of sound. The illusion of sound coming only
from the first (earlier) source is called the precedence effect.
Therefore, the Haas effect is also known as the integration and
precedence effect.
If the delayed sound is of the same level as the earlier sound (a
zero dB difference between the two sounds), the maximum delay
time during which integration and precedence effect occurs increases
to 50 milliseconds. If the delayed sound is 3 dB below the earlier
sound, the corresponding delay time for integration is increased
further to nearly 80 milliseconds. This is shown in Figure 2.6, in
which the integration zone has been indicated by the shaded area
the zone within which the blending of delayed sound occurs. Outside
this zone, the integration does not take place and the two sounds are
heard separately.
A sound that is heard separately from the earlier sound is referred
to as an echo. Echoes in listening spaces are disturbing. They
interfere with speech intelligibility and cause confused perception
of music, and therefore must be avoided.

10
8
2-
6
10
0
::::1-.::5
Ill
4 Echo zone
--

-.::5 ......
\)-.::5 2 \ --

\
S:
- 0
113 ::I
Ill 0
't=i -2 r--
L:
Integration .......
zone
-
:J -4 r-...
....._,
2-'t=i
-6 ['.,.
I I
10 20 30 40 50 60 70 M 90 100
2.6 Integration and echo zones for Delay time (in millie;econde;)
a delayed sound.

2.3.1 Practical Significance of the Haas Effect

The Haas effect is used in the design of auditoriums and in the design
of speech amplification systems. In an auditorium, the sound reaches
a listener in two ways: first the direct sound comes from the speaker,
and subsequently, the reflected sounds arrive at the listener after being
reflected from various surfaces of the room. Since the reflected
sounds travel through a longer path, they are delayed with respect to
the direct sound. The difference between the arrival times of direct
sound and a reflected sound is the delay time.
Sound Measurement and Hearing 33

In a typical auditorium or concert hall, a listener receives reflected


sounds from various surfaces, some of which are shown in Figure
2. 7. These reflected sounds arrive at the listener with different delay
times and levels. For instance, if the direct sound to a listener travels
through 10 m, and the sound reflected from one of the surfaces of
the room travels through 22 m to reach the same listener, then the
path length difference between the reflected and direct sounds is (22
- 10), i.e., 12 m. Since the speed of sound is 344 rnls, the reflected
sound is delayed by 35 milliseconds (obtained by dividing 12 by
344). Will this sound be integrated, or heard as an echo?
From Figure 2.6, we observe that the answer to the above question
depends on the level of reflected sound with respect to the direct
sound. Since a reflected sound travels through a longer distance
than the direct sound, its level is usually lower than that of the direct
sound. However, the sound reflected from some curved surfaces
(such as domes and vaults) can focus on a listener, increasing the
level of reflected sound above that of the direct sound.

r----- Sta wall reflctlon

Sid wall reflctlon

2. 7 Some of the various reflected


sound paths, and the direct sound eound

path, to a listener in an auditorium.

Since domes and vaults are generally avoided in auditoriums


and other listening spaces, the reflected sound is usually of a lower
level than the direct sound. Consequently, in the design of speech
auditoriums, we generally restrict the initial time delay between
reflected sound and direct sound to less than 50 milliseconds. Indeed,
from Haas's observations (Figure 2.6), this restriction is slightly
conservative since a 50 millisecond time delay is required when the
direct and reflected sounds are of the same level (zero dB difference).
A delay of 50 milliseconds corresponds to a path length difference
of nearly 17 m (55 ft) between direct and reflected sounds3. Thus, in
the design of speech auditoriums, we require that the path length
difference between a reflected sound and direct sound at the listener
should not exceed 17 m (55 ft). In practice, however, a round figure
of 20 m ( 65 ft) is used (see Section 11.7 .1 ). With this restriction, the

3 Since the speed of sound is 344 m/s, a time gap of 50 milliseconds (0.050 s)
corresponds to a path length difference of 344 x 0.050 1 7 m (55 ft).
=
34 Chapter 2

reflected sound at the listener is fully integrated and blended with


the direct sound. In addition, the listener gets the desirable illusion
that all the sound is coming directly from the speaker - not from
the direction of the reflected sound.
While a delay time of 50 milliseconds should preferably not be
exceeded in halls meant for speech, a longer delay time is acceptable
in halls meant for music. Typically, a delay time not exceeding 80
milliseconds is the criterion for music spaces (see Section 12.2.2).
A demonstration of Haas's effect can be accomplished simply if
we have a large free-standing sound-reflecting wall- say a concrete
or masonry wall, 30 m ( 100 ft) long and 30 m ( 100 ft) tall. If we
stand 25 m away from this wall and make a loud sound, we will hear
a distinct echo, because the reflected sound would travel 50 m and
reach us 145 milliseconds4 after the direct sound, Figure 2.8(a).
If we now move closer to the wall, say to a distance of 20 m
from the wall, and produce the same sound, the echo will be louder>,
since the reflected sound will travel a shorter distance (of 40 m),
reaching us after an interval of 116 milliseconds. Now if we move
3 m away from the wall, so that the reflected sound travels 6 m (20
ft), we will not hear any echo, just when we anticipated a louder
echo, Figure 2.8(b ). The reason for the absence of echo is that the
reflected sound reaches us after 17 milliseconds, which is within the
integration zone.

3m
I

(a) (b)

2.8 A simple Reflected eound le Integrated with the


Reflected eound le heard eeparately from direct eound
demonstration of the the direct eound
Haas effect.

2.3.2 Similarity with Seeing

The ability of the ear to blend sounds that follow at short intervals of
each other is similar to the ability of the eye to blend still
cinematographic pictures to give us an impression of motion. The
eye is able to blend discrete pictures into a continuous impression if
there are at least 16 pictures per second - an interval not exceeding
62 milliseconds between two still pictures. Therefore, some designers
assume the integration zone in acoustics to extend to 62 milliseconds,
although there is no known objective justification to assume similarity
between two distinct sensory perceptions.

4 Since the speed of sound is 344 rnls, it will travel 50 m in 0.145 seconds
( 145 milliseconds) - obtained by dividing 50 by 344.
5 The echo will be louder since the reflected sound will travel a smaller distance.
Remember, the sound level drops 6 dB for every doubling of distance in a free
field (see Section 1 .6).
Sound Measurement and Hearing 35

2.4 SOUND MASKING Another important aspect of hearing is masking. Masking of one
sound by another is an experience that we go through almost every
day. When we are unable to hear a speaker in an auditorium because
of high background noise, it is because the background noise is fully
or partially masking the sound of the speaker. When we have to
raise our voice to be heard in a noisy gathering, it is once again the
masking phenomenon that is in effect.
By definition, masking effect is measured by the number of
decibels a given sound must be raised above its normal threshold of
audibility to be heard in the presence of a masking sound. Masking
is related to the Haas effect, since the blending of a delayed sound
by the ear is a sort of masking- of delayed sound by the first arriving
sound.
Masking, however, is a more complex phenomenon than the Haas
effect, since masking has both neurological as well as sensory bases.
That is, masking is not simply the property of the ear but also of the
brain. For example, we are often able to hear distant conversations
of particular interest to us (or about us) in a noisy cocktail party. If
these conversations were not of interest to us, we might normally
not hear them.
Several studies have been reported on masking of pure tones by
pure tones, of pure tones by narrow band noise, of pure tones by
wide band noise, of noise by pure tones, etc. Only the following
conclusions from the above studies are of interest to us.
(i) A sound of a given frequency is more easily masked by a
sound of the same frequency. This means that the further away the
masking sound is in frequency from the frequency of the sound to be
masked, the greater the sound level of masking sound required. For
example, to fully mask a 65 dB, 400 Hz tone with another 400 Hz
tone requires a level of 80 dB. On the other hand, to completely
mask a 65 dB 1,000 Hz tone by a 400 Hz tone, a level far in excess
of 80 dB is required.
(ii) Low frequencies are generally more effective in masking
higher frequencies than vice versa, particularly if they are loud.
Excessive low frequency noises must, therefore, be avoided since
they constitute a serious source of interference for both speech and
music.
A certain amount of masking by background noise is necessary,
providing us with an "acoustical musk or aroma" since an extremely
quiet environment can be irksome and irritating. In an extremely
quiet environment, noises created by personal body movements and
breathing can be quite disturbing.
Excessive background noise levels are, however, undesirable
with one or two exceptions. For example, the masking phenomenon
is used to an advantage in open plan offices (large office areas with
low-height demountable partitions) where we intentionally provide
background noise with no information content to give speech privacy
at work stations, as discussed in Chapter 14.
36

2.5 BINAURAL Because we have two ears, human hearing is binaural. Binaural
hearing helps us locate a sound source in space, referred to as sound
HEARING localization. Specifically, it is because of our two ears that we are
able to qualitatively determine the angle that the direction of sound
makes with the line joining the two ears, Figure 2.9. In this respect,
our two-ear hearing mechanism is similar to our two-eye vision,
because of which we are able to perceive depth in space. One-eye
vision would not give us depth perception.
Studies indicate that the ears' ability to perceive the direction of
sound is due to: (i) different arrival times of sound at the two ears,
and (ii) different sound levels. Since the source-ear path lengths are
different for the two ears, the arrival times of sound at the two ears is
obviously different. The sound level differential is also caused by
the path length differential, but it is more due to the face and head
producing an acoustical shadow on the farther ear. Thus, in Figure
2.9, it is the left ear that is under acoustical shadow of the head, and
Sound consequently the sound level at the left ear is lower than at the right
ear. Because of the diffraction phenomenon6, acoustical shadows
are less pronounced at low frequencies.
Since sound localization is based on different arrival times and
sound levels, the ear has this ability only if the sound source is located
in the horizontal plane - the horizontal plane containing the sound
source and the two ears. In fact, the ear is able to localize the sound
source in the horizontal plane with an accuracy of 1 or 2 degrees.
If the sound source is located in the vertical plane - the vertical
plane passing through the center of the head and midway between
the ears - there is no difference between the arrival times of sound
to the two ears. Consequently, the ear cannot discriminate between
2.9 Direction localization by the ears. the direction of sounds in the vertical plane.
The inability of the ear to localize in the vertical plane is used to
advantage in establishing the location of loudspeakers in an
auditorium for sound amplification. The loudspeakers, which are
generally provided in the form of a loudspeaker cluster, are located
in the center of the proscenium. This locates the actual talker,
loudspeaker and the listener in a vertical plane. Therefore, the ears
are unable to distinguish between the directions of sound coming
from the talker and the loudspeaker.
If the ears were able to perceive directional differences in the
vertical plane, we would perceive two sounds, one from the talker
and the other from the loudspeaker, causing confusion. Sound
amplification systems are discussed in Chapter 16.

6 See Chapter 3 for a discussion of the diffraction phenomenon.


Sound
Reflection,
Diffraction and
Diffusion

Boundary elements of an enclosure have a profound influence on


the behavior of sound in an enclosure. In this chapter we will deal
with the effects of enclosure boundaries on sound. In particular, we
will qeal with the phenomena of sound reflection, diffraction and
diffusion. Sound absorption and sound absorbing materials are
covered in the next chapter.

3.1 THE BOUNDARY When sound energy falls on the boundary of an enclosure, such as a
wall or a ceiling, a part of the energy is reflected back into the
PHENOMENA enclosure, a part is absorbed within the material of the boundary and
converted into heat, and a part is transmitted through the boundary
element. The reflected sound expressed as a fraction of the total
sound energy falling on a boundary element is called the reflection
Tranemitted and al:>eorl:>lld coefficient of the element, denoted by the Greek symbol rho (p).
eound (80 unite) Thus:

Reflected sound energy


p =
Incident sound energy

incid11nt eound Rllfl11ct"d eound


(100 unite) (20 unite)
Since the reflected energy is always less than the energy incident
on the element, p is always less than 1. 0, and is generally expressed
as a decimal number. For instance, if the sound energy incident on a
3.1 Definition of reflection panel is 100 units, of which 2 0 units is reflected, then p = 2 0% or
coefficient (p) of an element. simply, p = 0.2, Figure 3. 1.

37
38 Chapter 3

The fraction of the incident sound that is transmitted through


the element is called the transmission coefficient, denoted by the
Greek symbol tau ('t). Similarly, the fraction of the incident sound
that is absorbed into the material is called its absorption coefficient,
represented by the Greek symbol alpha (a). Since the sum of the
reflected, absorbed and transmitted amounts of energy must be equal
to the incident energy, the following relationship must hold true:

p +a+ 't = 1.0

The most important property of a boundary element that affects


sound reflection, absorption and transmission characteristics is its
surface density (weight per unit area). Heavy weight elements are
more reflective, implying that they provide a strong reflection of
sound as compared with lightweight elements, Figure 3.2(a). In other
words, the value of p for a heavy weight element is high. Since a
heavy weight element is more reflective, less sound is available to
go through it. Thus, a heavy weight element transmits little sound,
i.e., the value of 't for a heavy weight element is small. Conversely,
a lightweight element transmits more sound and reflects less, Figure
3.2(b).

Transmitted and Tranemltted and


a11eorl1ed eound a11eorl1ed eound

Heavy weight
element t

Lightweight ---,
element

Reflected
eound
Incident
eound
/ ' ""''"""
eound

(a) REFLECTION FROM A HEAVY WEIGHT ELEMENT (11) REFLECTION FROM A LIGHTWEIGHT ELEMENT

3.2 Effect of an element's weight on sound reflection, transmission and absorption.

To appreciate the reason for this fact, consider a wall with a


sound source located as shown in Figure 3.3. Before the sound source
is turned on, the air on both the source side and the receiver side of
the wall is calm. However, when the source is turned on, the air
G) particles on the source side begin to vibrate. The vibrating air particles
Sound in proximity to the wall on the source side produce vibrations in the
eource
wall. The vibrating wall in turn produces vibrations in the air on the
receiver side, which are perceived as sound.
If the wall is heavy, the amplitude of vibrations in the wall is
small. Consequently, the vibrations in the air on the receiver side
have a small amplitude, implying that very little sound transmits
through the wall. On the other hand, a lightweight wall has larger
3.3 Vibrations produced in a wall by vibrations, and therefore, a greater amount of sound is transmitted
sound. through it.
Sound Reflection, Diffraction and Diffusion 39

3.2 ABSORPTION If we examine the reflection, absorption and transmission


characteristics of sound from the perspective of an enclosure's
COEFFICIENT OF interior, we find that both the absorbed and the transmitted parts of
SOUND sound energy are lost from the enclosure. Therefore, the absorbed
and the transmitted parts are grouped together and considered as the
absorbed part. In other words, in considering the acoustics of an
enclosure, we assume that of the sound energy that falls on an
enclosure boundary, a part is reflected and a part absorbed, i.e:

p +a = 1.0

In the above equation, the absorption coefficient (a) also includes


the transmission coefficient. Thus, the absorption coefficient is that
fraction of incident sound energy that is not reflected by the boundary
element. That is why, as far as the acoustics of an enclosure is
concerned, the two walls shown in Figure 3.4 are considered to have
the same absorption coefficient although the wall of Figure 3.4(a)
transmits more sound than that of Figure 3.4(b).
It is for the same reason that an open window, though not
absorbing any sound, is considered a perfect acoustical absorber
because all the sound falling on the window is transmitted outdoors.
Thus, for an open window, a = 1.0.

Reflected Reflected
eound 20% eound 20%

Abeorbed eound 79%


Abeorbed eound 507.

(/
Tranemltted eound 30% Tranemltted
eound 1'7.

Incident Incident
eound 100'7. eound 100%

(a) ( b)

3.4 Two assemblies with the same absorption coefficient (0.8) and reflection coefficient (0.2).

3.3 SOUND As stated previously, a building element must be heavy to provide a


strong reflection of sound. Another factor that affects reflection from
DIFFRACTION an element is the ability of sound to bend around an obstacle, referred
to as diffraction. If sound did not diffract, it would go over an obstacle
along a straight line path - like light, which produces an optical
shadow behind the obstacle.
40 Chapter 3

Diffraction causes a deviation from this straight line path and


the sound bends around the obstacle, so that the acoustical shadow

---t
t"
s::
is smaller than the optical shadow, Figure 3.5. Thus, it is because of
diffraction that we are able to hear a sound even when the sound
N
0
source is not visible to us.
s::
"
3: Diffraction is caused by the wave nature of sound, and although
0
-.:1
rce :;; 3: \IS we are representing sound to travel along straight line paths (or rays),
11) 0 ..1::
11)
=>-.:s
the reality is not exactly as shown in Figure 3.5. However, the straight
..,
0 \IS
"-'= Oi
< 11) line representation simplifies the discussion of the effects of
"R.
--+ 0 diffraction and is therefore a useful technique.
'
l 3.3.1 Wavelength of Sound and the Size of Reflector

3.5 Acoustical and optical shadows The degree of bending (diffraction) of sound around an obstacle is a
produced by a source. function of the sound's wavelength (or frequency, since wavelength
and frequency are related quantities). Low frequency (long
wavelength) sounds bend by a greater amount than high frequency
(short wavelength) sounds, Figure 3.6.

r OI:Ietacle

HIGH FREQUENCY
SOUND SOURCE
e LOW FREQUENCY
SOUND SOURCE
El:)
Acouetlcal ehadow

DIFFRACTION OF HIGH FREQUENCY SOUND DIFFRACTION OF LOW FREQUENCY SOUND

3.6 The effect of frequency on the diffraction of sound by an obstacle.

The consequence of the above fact is that if an element is small


in relation to the wavelength of sound, a large portion of sound energy
will bend around the edges of the element, creating virtually no
shadow behind the obstacle, Figure 3.7. Conversely, if the size of
the element is large compared with the wavelength of sound, most
of the sound incident on the element will be reflected by it; only a
Obetacle
small amount of sound will diffract. In other words, the diffraction
effect is a function of the dimensions of the element in relation to
the wavelength of sound.
Experimental observations indicate that for a plane (rectangular
3. 7 The effect of a small obstacle in panel) to reflect most of the sound falling on it, both its dimensions
the path of sound. must be at least SA.. Thus, the size of the panel must be at least 3 m
Sound Reflection, Diffraction and Diffusion 41

x 3 m (10 ft x 10 ft) if it is to be used as a reflector for a SOO Hz


sound, since A. for a SOO Hz sound is approximately 0.6 m (2 ft). If
the dimensions of the panel are progressively decreased from SA.,
increasingly greater diffraction will occur. When the panel size is
equal to A. in both directions, most of the sound will bend around the
panel with very little sound reflected from it, as shown in Figure 3.7.
Note, that regardless of the size of the obstacle, sound diffraction
will always occur around the edges of an obstacle. Thus, the reflection
of sound from a free-standing panel, even if it is quite large, is never
complete. However, for a panel greater than SA. in size, the amount
of diffracted sound is negligible.
Therefore, for a panel to function as an effective reflector, it is
necessary that both its dimensions be at least SA., Figure 3.8. As we
3.8 As far as possible, both dimensions shall see in Chapter 11, sound reflecting ceiling panels are commonly
provided in a speech auditorium to send reflected sound toward the
of a reflecting panel should be at least audience. Their size and stiffness1 are two important factors that
five times the wavelength ofsound to be determine their effectiveness as reflectors.
reflected.

3.4 THE RELEVANCE OF Since diffraction is wavelength (frequency) dependent, a high


frequency sound behaves like light, producing larger acoustical
ACOUSTICAL shadows than a low frequency sound. Thus, as shown in Figure 3.6,
SHADOWS the region of acoustical shadow behind an obstacle is larger for a
high frequency sound than for a low frequency sound.
An acoustical shadow has an unfavorable effect on hearing and
listening conditions in lecture and concert halls. For instance, an
acoustical shadow is formed by reflected sound under a deep balcony
in an auditorium, Figure 3.9. The shadow is deeper for high frequency
sounds than for low frequency sounds. Since it is the high frequency
component of speech that determines speech intelligibility (see
Section 1.2.3), poor hearing conditions are produced under a deep
balcony. Therefore, the depth of a balcony is usually limited in a
speech auditorium (see Section 11.2 for the definition of a deep
balcony).

Sound
0
!lourct:
3.9 Section through an auditorium with
a deep balcony showing the acoustical
shadow of ceiling reflected sound.

1 The stiffness of a panel is determined primarily by its surfaceweight, but also by


its profile. A curved or corrugated panel is stiffer than a planar panel.
42 Chapter 3

Music is also adversely affected. Although low frequency music


is able to penetrate into the balcony space by diffraction, high
frequency music cannot. Consequently, the music under a deep
balcony assumes undesirable tonal coloration. Since amplified music
is usually not acceptable, deep balconies must be avoided in concert
halls and other music spaces (see Section 12.8).
Although an acoustical shadow is undesirable for listening and
hearing, it is useful in the design of barriers to protect buildings and
neighborhoods from traffic noise. Since low frequency sounds
diffract substantially over the edges of an obstacle, a traffic noise
barrier must be high enough so that the barrier casts an acoustical
shadow over critical areas of the buildings to be protected, Figure
3.10(a). For the same reason, a traffic noise barrier must be long
and extend sufficiently beyond the end of the neighborhood, so that
the buildings to be protected fall within the acoustical shadow zone,
Figure 3.10(b). Traffic noise barriers are discussed further in Section
8.8.
Note that diffraction is not unique to sound. It occurs in all
kinds of wave motion including light, which is also a wave
phenomenon. However, we do not observe the bending of light
around the edges of objects because the wavelength of light is
extremely small in comparison with the objects in our environment.
For instance, the wavelength of light is less than 0.0007 mm. It is
because of the relatively very small wavelength of light with respect
to the dimensions of obstacles that light produces sharp shadows,
and hence the diffraction of light is usually not perceptible.
By comparison, the wavelength of sound (see Table 1.2) is of
the same order of magnitude as the dimensions of obstacles in our
environment, such as a freestanding column, a beam protruding below
the ceiling, a low-height partition, etc. It is because of the relatively
large wavelength of sound that sound diffraction is perceptible.

(a) SECTION

==llll ROAD
11

3.10 Sound d iffraction by a traffic


(b) PLAN ehadow zone
noise barrier.
Sound Reflection, Diffraction and Diffusion 43

3.4.1 Passage of Sound Through Openings

Diffraction effect also occurs when sound travels through an opening.


This is due to the bending of sound at the opening's edges.
Consequently, the amount of sound passing through an opening
consists of two parts:

that contained within the optical zone, and

that contained within the peripheral diffracted zone, Figure
3.11(a). The diffracted zone is a function of frequency,
increasing as the frequency decreases.
The consequence of the above fact is that: (i) the acoustical
transparency is always greater than the optical (or visual) transparency
of an opening, and (ii) for a given opening size, the acoustical
transparency increases with decreasing frequency.
Apart from the frequency, the size of the opening is also a
determinant of acoustical transparency. This is illustrated in Figures
3.11(a) to (c), which shows the passage of sound of the same
frequency through three openings of different sizes. Since the
frequency of sound is the same, the extent of diffraction (bending) at
the edges of all three openings is the same.
Therefore, the diffracted sound (as a percentage of the total sound
passing through the opening) increases as the opening size is reduced.
The smallest opening has the largest percentage of diffracted sound.
Through an extremely small opening, most of the sound passes by
diffraction.

Diffracted zone

Optical zone
Open In(!

Source
0:::::::::::_
o
Source .. _ .
Opening

Diffracted zone

(a) (b) (c)

3.11 Passage of sound of the same frequency through openings of three different sizes. As the
opening size decreases, an increasingly larger percentage ofsound passes through the opening by
diffraction.

In fact, when a sound wave falls on a small opening (much


smaller than the wavelength of sound), the opening behaves as a
(secondary) point source, radiating sound across the obstacle in a
hemispherical pattern, Figure 3.12. The amount of sound passing
through such an opening is no doubt small, but it is much greater
than the relative size of the opening. If there are several such openings
in a panel, the amount of sound passing through the panel can be
quite high.
e
Source
For instance, if the area of voids in a screen is 30% (i.e., a visual
transparency of 30%) the amount of sound that passes through the
,

screen (acoustical transparency) is much greater than 30%. In fact,


3.12 Passage of sound through a very virtually 100% of low frequency sound passes through a screen of
small opening. 30% visual transparency.
44 Chapter 3

3.5 ACOUSTICAL The acoqstical transparency of a screen is not merely a function of


its visual transparency and sound frequency, but also a function of
TRANSPARENCY OF A the distribution of voids in the screen. For a given visual transparency,
SCREEN small closely spaced voids provide greater acoustical transparency
than large voids spaced farther apart. For example, the screen of
Figure 3 . 1 3(a) has a greater acoustical transparency than the screen
of Figure 3 . 1 3(b), which in turn has greater acoustical transparency
than the screen of Figure 3. 13(c), although the visual transparency
of all three screens is the same - 30% . Indeed, the acoustical
transparency of all three screens increases with decreasing sound
frequency.
The reason for this fact is provided by the diffraction
phenomenon. In a screen with small closely spaced voids, such as
that of Figure 3 . 1 3 (a), the individual solid areas are small .
Consequently, a large amount of sound is able to bend around the
solid areas (obstacles) and pass through the voids. In a screen with
only a few large voids, such as the one of Figure 3. 13(c), the solid
areas are large. Therefore, a larger portion of sound is reflected by
these solid areas, and a smaller amount passes through the voids.
Knitted, woven or perforated fabrics with their small closely
spaced voids have excellent acoustical transparency. Such fabrics
are frequently used as protective and decorative coverings over
porous sound absorbing materials2 . By providing a high acoustical
transparency, these fabrics do not (greatly) reduce the effectiveness
of sound absorbing materials.

D c

(a) HIGHEST ACOUSTICAL (t:>) INTERMEDIATE ACOUSTICAL (c) LEAST ACOUSTICAL


TRANSPARENCY TRANSPARENCY TRANSPARENCY

3.13 Three perforated panel screens of the same (30%) visual transparency but different acoustical
transparencies.

3.5.1 Commonly Used Screens

Several manufacturers make fabric covered sound absorbing panels.


Typically, these consist of rigid fiberglass boards held in a wood or
metal frame and wrapped with perforated fire resistant fabrics, Figure
3 . 14. The panels can be easily hung on a wall, Figure 3. 15. The
availability of different sizes, fabric colors, patterns and textures
provide a great deal of design freedom.

2 Porous sound absorbing materials, consisting of fiberglass or mineral


wool, require protective facings (see Chapter 4).
Sound Reflection, Diffraction and Diffusion 45

Peripheral eupport
frame

High denelty
fll7erglaee

-----1--- Fal7ric
covering

3.14 Sound absorbing panel made out offabric-wrapped high


density fiberglass. Sample courtesy of Panel Solutions Inc.,
Hazleton, Pennsylvania. Photo by Madan Mehta.

Metal cllpe epoxled


to panel

/W::-- Metal cllpe


anchored to
wall

covered a17eorl71ng
panel

3.15 One of the several ways of hanging a fabric-wrapped sound


(a) HIGHER ACOUSTICAL
absorbing panel on a wall.
TRANSPARENCY

Perforated plywood, hardboard or metal panels are also used as


covering materials. Metal panels are particularly effective in dusty
environments, since the panels can be taken down, washed and put
back in place.
(b) LOWER ACOUSTICAL
The diffraction phenomenon also explains why a thicker screen
TRANSPARENCY
is less acoustically transparent than a thinner screen. Thus, of the
two perforated panels, the one of Figure 3 . 1 6(a) is more acoustically
3.16 Perforated panels of different transparent than that of Figure 3 . 1 6(b), although the visual
transparency of both screens is the same. For perforated metal panels
panel thickness.
Chapter 3

which are usually 0.75 to 1 .5 mm thick, an acoustical transparency


of 90% is achieved for a visual transparency of 30%, even at a high
b 12J 12J [] [] [] [] d frequency of 8 kHz[3 . l l . Similarly, of the two wood slat screens, the
one of Figure 3 . 1 7 (a) has a higher acoustical transparency than that
(a) HIGHER ACOUSTICAL
of Figure 3 . 1 7 (b) . Indeed, the acoustical transparency of both screens
TRANSPARENCY
decreases with increasing frequency.
A screen made of vertical wood slats is commonly used for the
protection of absorbing material. Its popularity lies in the warmth
and attractiveness of wood as an interior finish. Such a screen may
(1:1) LOWER ACOUSTICAL be obtained either in a prefabricated form, as shown in Figure 3 . 1 8,
TRANSPARENCY or custom designed to conform to the design philosophy of the
interior. There are many design possibilities, limited only by the
3.17 Acoustical transparencies of creativity of the designer. A few alternative patterns are shown in
Figures 3 . 1 9 and 3 .20.
wood slat screens with different slat
A typical construction detail of a wood slat screen is shown in
depths. Figure 3 .2 1 , in which the fiberglass pads are held between a wood
framework. The pads are covered with a wire mesh or grille cloth,
and finally covered with the wood screen.

3.18 Prefabricated wood slat screen, Keller


High School Auditorium, Keller, Texas. Screen
by Howard Manufacturing Company, Kent,
Washington. Acoustical Consultant: lames
Johnson. Photo by Madan Mehta.

H u jj l n 11p 11
n n

Jlllll 11 11 11

3.19 A few alternative wood slat screen


patterns in plan. 3.20 Two alternative elevational patterns ofwood slat screen.
Sound Reflection, Diffraction and Diffusion 47

,----- Supporting frame


Fll:>erglaee
[ B.:k -w=11------

3.21 Typical detail ofwood slat screen


covering over fiberglass blanket (as
sound absorbing material). Wire meeh or grille c;loth

Metal louver
ec;reen

(PART) SECTION THROUGH A METAL LOUVER


CEILING

3.22 Ceiling screen madefrom U-shaped metal louvers. Sample oflouver screen courtesy of Chicago Metallic,
Chicago, Illinois. Photos by Madan Mehta.

Screens made of U-shaped pressed metal louvers, Figure 3.22,


are an alternative to wood slat screens and provide a high acoustical
transparency. With black fiberglass pads placed above the screen
for sound absorption, a louvered screen yields an attractive ceiling.
Brick screens of the type shown in Figure 3.23 may be used, but
their acoustical transparency at high frequencies is low due to
reflection of sound from individual bricks.
As we will see in Section 4.7, screens and perforated panels
have sound absorbing properties of their own. Therefore, some
screens, particularly perforated panels, modify the sound absorption
characteristics of fiberglass or any other sound absorbing material
3.23 A brick screen. they protect.
48 Chapter 3

3.6 DIFFUSE AND As stated previously, to be a good sound reflector, a building element
must be sufficiently large in relation to the wavelength of sound and
SPECULAR also sufficiently stiff - of heavy weight construction. As we shall
REFLECTIONS see in Chapter 4, a good sound reflector must also be nonporous.
Sound reflection from a large, heavy and a nonporous surface can be
either:
specular reflection or
diffuse reflection.

3.6.1 Specular Reflection


Specular reflection is a mirror type reflection, similar to the reflection

-
of light from a mirror. In specular reflection, the incident sound
beam is reflected off the reflecting surface as per Snell 's law.
According to this law, the reflected beam makes the same angle with
orma (the normal to) the reflecting surface as the incident beam. In other
Incident eoun.:l
words, the angle of incidence (i) is equal to the angle of reflection
eoun.:l
(r), Figure 3.24.
Li = Lr For specular reflection to occur, the reflecting surface should be
"smooth". The word "smooth" here implies that surface texture and
3.24 Specular sound reflection. irregularities are much smaller than the wavelength of sound. Thus,
surfaces with recessed mortar joints in masonry walls and exposed
aggregates in a concrete wall behave as smooth walls, since their
irregularities are much smaller than the wavelength of sound of
interest to us.

3.6.2 Diffuse Reflection

Reflector The opposite of specular reflection is diffuse reflection. In diffuse


reflection, the incident sound is reflected equally in all directions
(uniform scattering), as shown in Figure 3.25. Diffuse sound
reflection is similar to the reflection of light by a matt surface or
Reflecte.:l eoun.:l
lncl.:lent frosted glass. For diffuse sound reflection, the reflecting surface
eoun.:l must be heavily textured and irregular - the dimensions of
irregularities should be nearly equal to the wavelength of sound.
3.25 Diffuse sound reflection. Thus, for a wall to provide diffuse reflection at 1 kHz (A
approximately equal to 0.3 m, i.e. 1 ft), its surface irregularities should
be of the order of 0.3 m ( l ft). Surface irregularities of a few
centimeters will provide specular reflection at 1 kHz frequency.
Consider once again a wall with surface irregularities of
approximately 0.3 m (1 ft), as shown in Figure 3.26(a). A sound
with a frequency of 100 Hz would be specularly reflected from such
a surface, since the wavelength of sound (approximately 3 m or 10
ft) is much greater than the dimensions of surface irregularities. In
other words, a 100 Hz sound will not see these irregularities and the
wall will behave as a smooth wall. On the other hand, a l kHz
sound will be diffusely reflected from this surface, Figure 3.26(b).
At a frequency of 500 Hz, the reflection from such a surface will
be partially diffuse and partially specular, implying that the scattering
of sound will be nonuniform with a large proportion of reflected
sound going in the direction of a specularly reflected beam. At a
frequency of 10 kHz, with a wavelength of approximately 30 mm
(nearly 1 in.), each individual irregularity is large enough to function
as an independent reflector. Therefore, sound will be specularly
reflected from each surface irregularity, Figure 3.26(c). Indeed, this
will provide some scattering of sound since the surface irregularities
are oriented in different directions.
Sound Reflection, Diffraction and Diffusion 49

0.3 m (1 ) t=

(a) Rcrllon of eound of 100 Hz frequency (1:1) Rcrllectlon


of eound of 1 kHz frequency
(5PECULAR REFLECilON) (DIFFUSE REFLECiiON)
Wavelength of eound (3.4 m, I.e. 11 ) ie much larger Wavelength of eound (approximately 0.3 m, I.e. 1 ) le
than eurface lrregularltlee. comparable to surface lrregularltlee.

(c) Rcrllon of eound of 10 kHz frequency


(5PECULAR REFLECiiON FROM EACH
INDIVIDUAL IRREGULARITY)
Wavelength of eound (30 mm, nearly 1 In.) is
much emaller than surface lrregularltlee. Hence
sound le epecularly rcrllected from each surface.

3.26 Effect offrequency on the reflection of sound from an irregular surface.

3.7 SOUND DIFFUSION It is obvious from the above discussion that if room boundaries consist
of sufficiently large surface irregularities, the sound field in such a
room will be diffuse. A perfectly diffuse sound field is defined as
one in which sound arrives at the listener from all possible directions
in equal strength, Figure 3 .27.
A perfectly diffuse sound field does not usually exist in a room
since the direct sound at most listeners' positions is stronger than the
reflected sound. However, an approximately diffuse sound field is
obtained in a room with highly reflecting surfaces at locations far
away from the source, because at these locations the reflected sound
predominates.
Sound diffusion is one of the important acoustical requirements
for rooms used for musical performances. A room with a few large
specularly reflecting surfaces, and which does not contain adequate
surface irregularities to diffuse sound, produces harsh reflections,
known as acoustic glare - an undesirable effect for music. On the
other hand, with adequate diffusion in the room, the listener receives
sound from various directions and has the feeling of being
"enveloped" by music - a desirable sensation for music (see Section
1 2.4.4).
Excessive diffusion, on the other hand, deprives the listener of
source localization, since in a diffuse field, the sound appears to
3.27 In a diffuse field, the sound arrives
come from all directions. Excessive diffusion is to be avoided in
at a listenerfrom all directions in equal rooms meant for speech, since in these rooms the sound must appear
strength. to come from the speaker.
50 Chapter 3

3.7.1 Effect of Room Geometry and Size on Sound Diffusion

Sound diffusion is a function of room geometry. Rectangular rooms


with flat parallel walls have poor diffusion. Even a slight splay ( 1 in
20) in one of the walls in an otherwise rectangular room improves
diffusion. It also eliminates flutter echo, discussed later in this
chapter. In fact, the more the room deviates from rectangularity, or
the more irregular the room shape, the greater the sound diffusion in
the room.
Size of the room is another factor that affects diffusion. Diffusion
is more easily obtained in a large room than in a small room, as
explained in Section 1 0.8. Thus, because of its small size, it is difficult
to achieve diffusion in a music recording studio or a control room
unless special sound diffusers are used on room surfaces.

3.7.2 Effect of Sound Absorption on Sound Diffusion

Reflective room surfaces increase diffusion in the room. The more


reflective the surfaces, the greater the diffusion. Conversely, the
provision of sound absorption decreases diffusion. A reverberation
chamber in an acoustical laboratory (used for measuring the
absorption coefficient of materials) is required to have a highly diffuse
sound field. Therefore, it is usually built of heavy concrete walls
and ceiling- with a nonrectangular geometry, and/or provided with
randomly placed reflecting surfaces, as shown in Figure 3.28.
Even when only one surface of the room is highly absorbing,
such as the floor (audience area) in an auditorium, sufficient diffusion
is difficult to obtain unless other means of increasing diffusion are
incorporated.

3.28 Reverberation chamber. Courtesy of Riverbank Acoustical Laboratory,


Geneva, Illinois, with permission.
Sound Reflection, Diffraction and Diffusion 51

Although sound absorption reduces diffusion, the alternate


Appr application of sound absorbing patches, as shown in Figure 3 .29,
M improves diffusion. The size of patches must be of the order of the
wavelength of sound. Therefore, to produce diffusion over a wide
band of frequencies, patches must be of various sizes. Note, however,
that alternate application of absorbing patches to obtain diffusion
should be used only in spaces where sound absorption is otherwise
required.

3. 7.3 Interior Ornamentation

t l 5ound reft.,ctlng
!5urfac"
Pilasters, piers, balconies, exposed beams, coffered ceilings, and any
other surface ornamentation that scatters sound increase diffusion.
Sufficient diffusion, provided by extensive ornamentation and
ound alleorlllng maurlal
protruding balconies is considered to be one of the reasons for the
good acoustics of the some of the older symphony halls, such as
Symphony Hall in Boston, Figure 3.30.
3.29 Alternate application of sound
absorbing materals.

3.30 Symphony Hall, home of the Boston Symphony Orchestra. Courtesy of Boston Symphony Orchestra,
Boston, Massachusetts, with permission. Photo by Bradford Herzog.
52 Chapter 3

3. 7.4 Diffusion and Convex Reflectors

Convex reflective surfaces also increase diffusion. They do so by


scattering sound, Figure 3 .3 1 . The convex exposed brick walls in
the hall of Figure 3.32 serve the purpose of improving diffusion. A
concave surface, on the other hand, tends to focus sound, Figure
3.33. Focusing is the opposite of diffusion since focusing tends to
concentrate sound into one direction and location, starving other
locations of adequate sound. Thus, a dome or similar concave surface
provides poor acoustics for an auditorium, unless the dome has deep
coffers to scatter sound.

3.31 Scattering of sound by a


convex reflector.

3.32 Irons Recital Hall, University of


Texas at Arlington, with convex profiled
brick walls. Photo by Madan Mehta.

3.33 S o u n d foc us in g by a
concave surface.

3.8 SOUND DIFFUSERS When sufficient diffusion cannot be obtained by the methods
described previously, sound diffusers may be used to increase
diffusion in the room. A sound diffuser is a surface element that
produces diffuse reflection.
Any reflective surface with irregularities of size comparable to
the wavelength of sound will work as a diffuser. The greater the
randomness in surface irregularities and sizes, the better the diffuser.
Sound Reflection, Diffraction and Diffusion 53

Thus, numerous geometries can be used to make a diffuser. Sheet


materials such as gypsum board, plywood, or hard board panels are
commonly used as diffuser material, but metals and masonry
materials may also be used. The diffuser shown in Figure 3.34 is
made of marble slabs, and those of Figure 3.35 may be made of
gypsum board or plywood.

(1:1) DOMICAL DIFFUSERS IN A WALL


3.34 Marble diffusers used in the walls OR CEILING
ofDe Doelen Concert Hall, Rotterdam, (a) CYLINDRICAL
The Netherlands. Courtesy of concert DIFFUSERS IN A A. (approx.)
en congresgebouw de Doelen, WALL

Rotterdam, The Netherlands, with


permission. Photo by Eric Spaans. See
s.,ctlon through a .::l om"
also Figure 12. 14.

3.35 Cylindrical and domical diffusers.

3.8.1 Quadratic Residue Diffuser

The diffusers made from surface modulations have two major


limitations. First, the surface protrusions and recesses have to be
large to provide good diffusion at low frequencies. Second, there is
no objective method of determining the extent of scattering produced
by such diffusers.
A diffuser that overcomes the above limitations is called a
quadratic residue diffuse.,-3 and is based on the theoretical work by
German acoustician M. R. Schroeder. It does not require large surface
modulations and the extent of scattering produced by it can be
ascertained fairly accurately.
A quadratic residue diffuser consists of an array of linear slits
(or wells) of constant width. The wells are separated by thin rigid
walls. The depths of wells vary according to a well-defined number
sequence. The sequence is repeated to produce a diffuser of the
required size. Each repetition of the sequence is called a period.

3 Appendix D gives the reason for the term: quadratic residue diffuser.
54 Chapter 3

Thus, if two repetitions are used, the diffuser consists of two periods,
Figure 3.36. Each period4 consists of a certain number of wells
which must be a prime number5 (a number that is divisible only by 1
or itself). The diffuser can be designed to provide sound scattering
within any required frequency band.
Quadratic residue diffusers are commercially available, in wood
and masonry, and are commonly used to increase sound diffusion in
auditoriums, music halls and recording studios, Figures 3 . 37 and
3.38.

One period One period

--tJf= Well width

3.36 A horizontal cross-section through a 1 -d quadratic residue diffuser with two periods. Each
period consists of a number of wells of constant width but varying depths.

tOne period 1 One period

3.37 Quadratic residue diffuser (N 7) used in the control room of the recording studio
=

of Collin County Community College (Spring Creek Campus), Plano, Texas. The difuser
was supplied by RPG Diffusor Systems Inc., Upper Marlboro, Maryland. Acoustical
consultant: lames Johnson. Photo by Madan Mehta.
Sound Reflection, Diffraction and Diffusion 55

.. ...
...
...

..

3.38 Quadratic residue diffusers used in the rea wall ofthe orchestral platform ofFrits Philips Muziekcentrum,
Eindhoven, The Netherlands. Courtesy of RPG Diffusor Systems Inc., Upper Marlboro, Maryland, with
fermission.

4 Increasing the number of periods of a quadratic residue diffuser increases


the length of the diffuser panel. However, the effectiveness of a quadratic
residue diffuser is higher if the periods are distributed over several panels,
separated from each other, rather than the same number of periods grouped
together in one long panel.

5 Some examples of odd prime numbers are 3, 5, 7, 1 1 , 1 3 , 17, 23, 29, etc.
56 Chapter 3

Quadratic residue diffusers, shown in Figures 3.37 and 3.38,


consisting of linear wells, are one-dimensional ( 1-d) diffusers. They
increase diffusion in a plane transverse to the wells. Thus, if the
wells are oriented horizontally, the sound is diffused in a vertical
plane. On the other hand, if the diffusers are oriented vertically, the
diffusion is produced in a lateral plane, Figure 3.39.
To produce three-dimensional diffusion, diffuser panels may be
oriented in both directions, as shown in Figure 3.40. Alternatively,
2-d quadratic residue diffusers may be employed, which consist of a
matrix of square shaped wells, Figure 3 .41.
A negative aspect of a quadratic residue diffuser is that it absorbs
sound (at low frequencies), which must be taken into account in the
design of rooms where sound absorption is a critical factor. The
manufacturers provide both the diffusion pattern as well as the
absorption data of quadratic residue diffusers. For an additional
discussion of quadratic residue diffusers, see Appendix D.

Incident

IJI!il::::=-:'f-t4 Normal to
plane of
dlffueer

Plane In which Incident


eound dlffueee 3.40 Diffus er panels consisting of
vertically and horizontally oriented
dlffueer 1-d quadratic residue diffusers.

3.39 The diffusion of sound by vertically


oriented wells in a quadratic residue
diffuser is produced in a semicircular disc
that is oriented in the same direction (8) to
the plane of the diffuser as the incident
sound.

3.41 A 2-d quadratic residue diffuser.


Courtesy of RPG Diffusor Systems Inc.,
Upper Ma rlboro, Maryland, with
permission.
Sound Reflection, Diffraction and Diffusion 57

3.9 SOURCE-IMAGE In Figure 3.24, we observed that specular reflection is governed by


Snell's law. Elementary geometrical considerations applied to Figure
RELATIONSHIP IN 3.24 show that the reflected ray can be considered to be originating
SPECULAR from behind the reflector - as if it originates from the image of the
source (1), as shown in Figure 3.42. This is identical to the image of
REFLECTION an object formed behind an optical mirror.
The distance of the image behind the reflector is equal to the
distance of the source in front of the reflector. Thus, SR = RI, where
S is the source and I its image. The straight line SI makes an angle
of 90 with the reflector.
Since the reflected sound can be assumed to originate from the
image, we can establish the direction of reflected sound at a listener
position (L) by simply joining I with L. The advantage of this
procedure is that once the image is located, we can disregard the
reflecting surface altogether. As we shall see in Chapter 1 1 , this
procedure simplifies determining the ceiling reflector profile in
auditoriums. It also simplifies the determination of the difference
between the path lengths of: (i) the direct sound and (ii) the reflected
sound to a listener.

I -.,
I lmag: of eourc: \ 5R = RI
L Lletlln:r poeltlon \
5 5ourc: 900 \ 1 R:fl:ctor
1 \,
R
Rllfl:ctllel sound

3.42 Source image relationship in specular reflection.

Remember from Section 2.3, that to prevent an echo, the path


length difference between the direct sound and the reflected sound
to a listener should not exceed 20 m (65 ft). If the listener is situated
at L in Figure 3.42, the path length difference between the direct and
reflected sound is given by:

Path length difference = reflected path - direct path


= (SP + PL) - SL = IL - SL

Thus, to ensure that an echo does not occur, all that is necessary
is to check that (IL - SL) does not exceed 20 m at any listener position
in an auditorium.
Note that in locating the image of the source, it is not necessary
that the reflector must extend up to point R. In other words, point R
need not lie on the reflector for a source to produce its image behind
the reflector. The image of the source can be located by an imaginary
extension of the reflector up to point R, Figure 3.43.
Figure 3.43 also indicates the part of the reflector that will supply
reflected sound between two listener positions, L l and L2. Thus, we
3.43 Reflected soundfrom a reflector
see that it is reflector length PQ that is respons1ble for providing
that does not extend up to the source. reflected sound between listener positions L 1 and L2. If L 1 and L2
58 Chapter 3

are two extreme listener positions in an auditorium, the above reflector


Ceiling area CD only need need only extend from P to Q.
l:>e reflective to provide The above concept6 is elaborated in Figure 3.44. For instance, if
ceiling reflected eound we wish to determine the portion of ceiling that will provide reflected
over the audience
sound in audience rows from A to B, we join points A and B with I.
The required reflector size is then given by CD. This implies that
only that part of ceiling which extends fr01 a C to D need be specularly
reflective; the remaining part of the ceiling can be treated differently.
If the reflector is smaller than CD, it will not send reflected sound
over the entire audience (from A to B). If it is larger than CD, a part
of the reflected sound will fall outside the audience area.

3.44 Section through an auditorium


showing reflector location. 3.9.1 Higher-order Images

If there is a set of two reflectors in a space, an image produced by


one reflector works as the source for the other reflector, producing
an image-of-an-image. Thus if there are two reflectors (reflectors 1
and 2) placed at right angles to each other, they produce three images,
Reflector 1 Figure 3.45 . Images 11 and 12 are the images produced by reflectors
1 and 2 respectively. I now works as a source for reflector 2,
lt ----
r producing an image 112. Similarly, 12 works as a source for reflector
.,eo0 1 , producing an image 121 .
-: - - - - - - - -1-'-
--- In this particular case (two reflectors at right angles to each other),
Rdfiector 2 112 and 121 lie at the same location. Images 11 and 12 are called first
0 oroer images, and 112 and 121 , second-order images.
lz A second-order Image simply locates the direction of the second
reflection of a sound ray. In a room, where a ray may be reflected
several times before being fully absorbed or weakening sufficiently,
there will be second, third, and higher order images. However, in
3.45 Images produced by two
designing an auditorium, we usually consider only the first order
reflectors at right angle to each other. images, since these are most significant.

3.10 FLUTTER ECHO If there are two parallel reflectors, we will obtain an infinite number
of images of the source since each image works as a source for the
other reflector. This may be confirmed by standing between two
parallel mirrors; an infinite number of images of the self will be seen.
The above fact implies that if a sound source is located between
two parallel reflecting walls, a listener will receive reflected sound
from an infinite number of images. This is simply another way of
stating that the sound will be reflected back and forth between two
parallel reflecting walls infinite number of times before exhausting
to inaudibility.
Now imagine a sound source (S) located between two reflective
parallel walls 1 5 m apart, as shown in Figure 3 .46. Obviously, this
situation produces an infinite number of images of the source. The
first-order images, images 11 and . are behind wall 1 and wall 2
respectively. The second order image, image 11 2, is the image of
image 11 and is formed behind wall 2. Image 121 is the image of
mage 12, behind wall l . Similarly, 1121 and 1212 represent third order
Images, and so on.

6 The image procedure for the study of reflected sounds is applicable only to a
planar reflector, not to a curved reflector. However, Snell's law is applicable to all
types of reflectors.
Sound Reflection, Diffraction and Diffusion 59

120 m

90 m

60 m

30 m
10 m 5m
+ +
1
0 0 0

1,2 12 s I,

Wall 2 Wall 1

3.46 Images produced through two long parallel reflecting walls.

If we determine the distance between images, we find that the


distance between successive order images increases by 30 m- twice
the distance between walls. Thus, the first order images are 30 m
apart, second order images are 60 m apart, third order images are 90
m apart, and so on.
Since the speed of sound is 344 m per second, the time ga.p
between each successive reflected sound will be 87 milliseconds .
This, according to the Haas effect (Section 2.3), will produce echoes.
Since these echoes recur after a regular interval of 87 milliseconds,
they produce a flutter effect; hence the term flutter echoes.
If the distance between walls were 5 m, successive order images
would be 10 m apart. Therefore the time gap between successive
reflections would be 1 0/344, i.e., 29 milliseconds, which (according
to the Haas effect) should not be perceived as echoes. However, the
flutter is heard all the same. The reason lies in our ears being
extremely sensitive to periodic repetition of sounds.
The existence of flutter echoes can be easily demonstrated by
clapping in a longish room, such as a corridor, which has two parallel
and highly reflecting walls, and an absorbent floor and ceiling.
Flutter echo is an acoustical defect and must be avoided in
auditoriums and other assembly spaces. It affects speech
intelligibility and produces tonal coloration of music. For example,
if the time gap between periodic reflections is 20 milliseconds, the
space will add a sound of 50 Hz frequency 8 to any sound produced
in this space.
In short, two parallel reflective walls should be avoided in an
auditorium. Splaying one or both walls of the room by as little as 5
degrees will usually eliminate flutter effect. Alternatively, one or
both walls may be treated with sound diffusers. Treating one of the
parallel walls with sound absorbing material will also eliminate
flutter.

7 This is obtained by dividing 30 m by 344 m/sec yielding 0.087 sec, i.e. 87


milliseconds.

8 One reflection every 20 milliseconds means that 50 reflections will be received


in one second - a frequency of 50 Hz.
60

3.1 Schultz, Theodore: Acoustical Uses for Perforated Metals, Industrial


REFERENCE
Perforators Association Inc., Milwaukee, Wisconsin, 1986.
Sound
Absorbing
Materials

All materials and objects absorb sound to some degree. However,


materials that are specifically employed for the purpose of absorbing
sound are called sound absorbing materials, or acoustical materials
- although the former term is preferable. For the same reason, the
term acoustical treatment usually implies sound absorptive treatment.
Sound absorbing materials are used to reduce noise in interior
spaces. In Chapter 8, we will examine the extent of noise reduction
achievable by their use. Sound absorbing materials are also used to
control reverberation in assembly spaces - a topic covered in
Cnapter 10.
In this chapter we will discuss various sound absorbing materials
and the mechanisms by which they absorb sound. The chapter begins
with a description of how sound absorbing materials are rated for
their effectiveness.

4.1 RATING OF SOUND The standard method of rating the effectiveness of a sound absorbing
material is by its absorption coefficient, which has been defined in
ABSORBING Section 3 .1. The absorption coefficient of a material varies with the
MATERIALS angle of incidence of sound- the angle at which the sound strikes
the surface of the material. However in most rooms, the sound strikes
its surfaces from all angles with almost equal probability. Therefore,
we are usually interested in the random incidence absorption
coefficient1

1 In air-conditioning ducts, the sound travels at grazing incidence (tangential)


to the sound absorbing material (duct lining).

61
62 Chapter 4

The random incidence absorption coefficient is the absorption


coefficient averaged over all the angles of incidence. In the following
discussion, the random incidence absorption coefficient will be
referred to simply as the absorption coefficient (a) of the material.
The sound absorption coefficient also varies with the frequency
of sound. In architectural acoustics, we are normally concerned with
the values of a in six octaves, ranging from 125 Hz to 4 kHz. Thus,
the values of a are generally quoted at 125 Hz, 250 Hz, 500 Hz, 1
kHz, 2 kHz, and 4k Hz. The values of a for commonly used building
materials are given in Appendix H.

4.1.1 Noise Reduction Coefficient

The noise reduction coefficient (NRC) of a material is the average


value of the absorption coefficients at 250, 500, 1,000 and 2,000 Hz,
rounded off to 0.05. Thus:

a a a
250 t.ooo 2.ooo
a +
+ 5 00 +
=
(4.1)
NRC
4

For example, if the values of a for a material are:

Frequency (Hz) a

125 0.16
250 0.31
500 0.52
1,000 0.83
2,000 0.9 1
4,000 0.97

the NRC of the material is:

0.31 + 0.52 + 0.83 + 0.91


= = 0.65
NRC
4

We see from Equation (4.1) that NRC is a single number rating


of the sound absorptive property of a material. The convenience of
NRC lies in its single number specification. However, since the low
and high frequencies are not represented in the NRC value, the
applicability of NRC data is limited to those noise control situations
where most of the sound energy lies between 250 Hz and 2 kHz.
Such situations include interiors where noise is primarily due to
speech, since most of speech energy lies in the four octaves ranging
from 250 Hz to 2 kHz (see Figure 1.5).
Sound absorptive materials used in offices, restaurants, airport
lounges, and such other assembly spaces are commonly specified
based on the NRC value. In situations where a significant amount
of sound energy lies outside the above four octaves, the values of a
at these other frequencies must also be examined.
Sound Absorbing Materials 63

4.1.2 The Unit of Sound Absorption

The product of the area of an absorber and its absorption coefficient


is called the sound absorption of the material. Thus, if the surface
area of a material is S, and its absorption coefficient a, then the
sound absorption provided by the material (A) is:

(4.2)

The unit of absorption (A) is called sabin, after the American


acoustician Wallace Clement Sabine (1868-1919). Thus, if the
surface area of a material is 10 m2 and its absorption coefficient
0.75, the amount of absorption provided by this material is 10(0.75)
= 7.5 sabins.

Since the surface area may be in square feet or square meters,


the unit sabin is either foot sabin or metric sabin. Since 1 m2 =
10.76 ft2, 1 metric sabin = 10.76 ft sabins.
For a room with its several boundary surfaces, the total absorption
provided by room boundaries (LA) is given by:

(4.3)

where, S 1, S 2, S3, ... are the surface areas of the room, and a1, a2,
a , ... are their respective absorption coefficients (see also Section
3
1U.3).

4.1.2 Average Absorption Coefficient of a Room

The average absorption coefficient of room boundaries (a3y) is


defined as the total absorption of room boundaries divided by the
total surface area of room boundaries. Thus:

LA (4.4)
=
St + Sz+ S3 +
LS

The use of Equations (4.3) and (4.4) is illustrated in Examples 4.1


and 4.2.

4.2 TYPES OF SOUND Sound absorbing materials may be classified under the following
ABSORBING
three types - a classification based on the mechanism by which
they absorb sound.
MATERIALS
Porous absorbers
Panel or membrane absorbers
Volume absorbers
64 Chapter 4

Example 4.1 Total Sound Absorption of a Room

The sound absorption coefficients of various surfaces of a manufacturing room measuring 25 m x 20 m


x 6 m (82 ft x 65.6 ft x 19.7 ft) are: aceiling = 0.6, all = 0.05, <Xfloor= 0.1. Calculate the total sound
absorption in the room.

Solution: The area of walls= 2(25 x 6) + 2(20 x 6)= 540 m2. Therefore, the amount of sound absorption
provided by walls= (540)0.05 = 27.0 sabins.
Similarly the sound absorption provided by the ceiling= (25 x 20)0.6= 300 sabins, and that provided
by the floor = (25 x 20)0.1 = 50.0 sabins.
From Equation (4.3), the total absorption of the room= LA = 27.0 + 300.0 + 50.0 = 377.0 (metric)
sabins.
If the areas are expressed in square feet, the resulting absorption is in (ft) sabins. The reader may
confirm that this is equal to 377.0 x 10.76 = 4057 (ft) sabins.

Example 4.2 Average Absorption Coefficient of a Room

Determine the average absorption coefficient of the room of Example 4.1.

Solution: From Example 4.1, LA= 377.0 sabins. The total surface area of room, LS= 540 + 500 + 500
= 1540.0 m2. Hence, from Equation (4.4),

377.0
= 0.25
1540.0

4.3 POROUS Almost any material whose surface is porous may be considered a

ABSORBERS
porous absorber. The porosity of the material may be either due to
the fibrous composition, or due to voids between granules or particles
of the material. Fiberglass and mineral wool2 are the most commonly
used porous absorbers. Other commonly used porous absorbers are
rigid mineral fiberboards with fissured or pierced surfaces - used
primarily as ceiling tiles in commercial interiors (see Section 4.9).
Wood fibers bonded with an inorganic hydraulic cement to form
rugged and high impact resistant boards is another type of porous
absorber (see Section 4.4.1).
A sound wave falling on a porous absorber causes the air in the
voids of the material to vibrate back and forth. As the air vibrates in
the voids, the vibrational energy of the air is converted into heat due
to friction between air particles and void walls.
For frictional losses to occur, it is important that the voids in the
material be interconnected and continuous so that the air can pump
back and forth within the material. Thus, only open-cell materials
can be good sound absorbers. Plastic foams such as expanded

2 Fiberglass and mineral wool are similar products, and are often used
interchangeably for sound absorption.
Sound Absorbing Materials 65

polystyrene or polyisocyanurate (used extensively for thermal


insulation) are poor sound absorbers because of their closed
unconnected cell structure. However, melamine, polyurethane, and
polyester foams with their interconnected cells make good sound
absorbers.
The advantage of plastic foams lies in their fiber-free composition
and light weight. An additional advantage is their moldability.
Consequently, sculpted foam sound absorbing tiles and other products
are available. The disadvantage of foams is their combustibility and
emission of toxic fumes upon combustion. Where these concerns
are unimportant such as in sound absorbing enclosures for machines
and appliances, plastic foams are preferable to fibrous products. Non
burning, fire resistive foam sound absorbing products are available.
However, local building codes must be consulted before specifying
plastic foams in interior construction.
The effectiveness of a porous material depends on its flow
resistance, defined as the degree of difficulty by which the air will
flow through the material. If the flow resistance is too low, the
frictional losses are small. Consequently, the value of a. is low. On
the other hand, if the flow resistance is too high, the air flow through
the material is so limited that the frictional losses are once again
low. Thus, there is an optimum value of flow resistance that provides
the highest value of a. for the material.
The flow resistance depends on the density of the material, and
if the product is fibrous it also depends on fiber diameter. In general,
the greater the density and the thinner the fiber, the greater the flow
resistance. A simple test to estimate the flow resistance of a porous
material is to blow through it. Materials that allow little or no air to
pass through provide little or no sound absorption.
Fiberglass (or mineral wool) with a density ranging from nearly
15 to 145 kg/m3 (1 to 9 lb/ft3) is commonly used as a sound absorbing
material. Higher density fiberglass (or mineral wool) is used in fabric
covered sound absorbing panels, where the panel must be rigid and/
or tackable. The lower density materials are used as blankets or
batts, and must be protected by rigid screens, such as perforated
panels, wood slat screens, metal screens, etc. (see Section 3.5).

4.3.1 Impervious Membrane Over a Porous Absorber

Since the absorption property of a porous absorber depends on the


air pumping back and forth into the material, a porous absorber can
be covered with an impervious membrane. This is particularly useful
in the case of fibrous materials such as fiberglass and mineral wool.
However, the membrane must be: (i) extremely light and flexible
- .012 to .025 mm (0.5 to 1 mil) thick plastic sheet3 - so that it
provides little resistance to the pumping of air into the body of the
material; and (ii) loosely wrapped over the absorber, not solidly
attached to it. A solidly attached membrane will not provide pumping
action, and will tend to reflect sound, particularly at high frequencies.

3 1 t
mt =
1/ 1,000 m.
.
66 Chapter 4

4.3.2 Thickness of Porous Absorber and the Value of a

The value of a for a porous absorber increases with the thickness of


the absorber, Figure 4.1. This is because the particle velocity4
maximizes at a distance of 0.25A. from the substrate- wall or ceiling
(see Appendix C).
If the thickness of the porous absorber is small, the maximum
particle velocity occurs outside the absorber, Figure 4.2(a), giving a
low value of a. Conversely, if the thickness of the absorber is large,
the maximum particle velocity is likely to occur within the absorber,
Figure 4.2(b). This likelihood increases as the thickness of absorber
increases, thereby increasing the value of a with absorber thickness.
Stated differently, the above fact implies that the layer of porous
absorber which is in direct contact with the substrate is useless in
absorbing sound since the particle velocity is zero there. The
effectiveness of a layer increases with its distance from the substrate
until it maximizes at 0.25A. away from the substrate. After that, the
effectiveness decreases and maximizes again at 0.75A., 1.25A., etc.,
away from the substrate (see Appendix C, Figure 2).

1.2 .................. ...


.J
100 mm /, ........ -r-,.,,.......... .
. , .......
- 41n. ..... ..
1.0 . ,-"5o m ::: ::::
: :::
: : : : : ::
: :
(2 In.)
1
I
0.8 I
/ / /"25 mm
hl----+-1+-, --+--/
+'
s::
\)
--.>Y-(1-. In.) -

0.6
V
"i3
ii:
0
\) / Fll:>erglaee
I
u
s:::
0.4 I
0 Sul:>etrate ---j.,,....;;...,.

I:>.. I I
/ /
....
0
0.2
ID

< /

0 L_ _L__i__---L-

125 250 500 1,000 2,000 4,000


_

Frequency (Hz)

4.1 Approximate frequency-a relationships for 25 mm (1 in.), 50 mm (2 in.)


and 100 mm (4 in.) thick fiberglass (density 48 kg!m 3- 3 lb/frl). Adapted
from Reference 4.1.

4 Since a porous absorber absorbs sound by converting the kinetic energy of


air particle vibration into heat, it is the decrease in particle velocity caused by
a porous absorber that determines the absorber's effectiveness.
Sound Absorbing Materials 67

Part;lcle velocity
amplitude variation

Poroua abaorber




.... -
..

- 5ubatrate
.
.

.. .
. .. ..
.... . .

0.25A

5ubatrate

(a) THIN POROUS ABSORBER (11) THICK POROUS ABSORBER

4.2 Particle velocity amplitude variation superimposed on a section through a wall covered with a porous
absorber.

Note from Figure 4.1 that for the commonly used thickness of
25 to 50 mm (1 to 2 in.), the absorption coefficient of a porous
absorber increases with frequency. For low frequencies, the absorber
thickness is small compared with 0.25A., giving a low value of a.
For high frequencies, the absorber thickness is comparable to 0.25A.,
Figure 4.3. At 100 mm (4 in.) thickness, a fiberglass blanket provides
fairly high values of a in both low and high frequency regions. A
200 mm (8 in.) thick fiberglass gives a virtually flat frequency-a
relationship.
The value of a may exceed 1.0. This apparent anomaly is due to
the diffraction of sound at the edges of the test sample- a peculiarity
of the test procedure for measuring the value of a. Values of a of up
to 1.3 may be obtained for thick porous absorbers. Since an a greater
than 1.0 means that more energy is absorbed than what falls on the
material (an obvious impossibility), it is a good practice to reduce it
to 1.0, or even 0.995.

Part;icle velocity amplitude


variation for low frequency Poroua abaorber

Part;lcle velocity amplitude 5ubatrate


variation for high frequency

4.3 Comparison of the particle velocity amplitude variation for


low and high frequencies.
68 Chapter 4

4.3.3 Effect of Space Between Porous Absorber and Substrate

If a porous absorber is spaced away from the substrate with an


intervening air cavity, its frequency-a relationship is similar to that
of a thicker material placed against the substrate without any cavity.
In other words, the provision of air cavity increases the sound
absorption of the material at low frequencies. At frequencies above
1,000 Hz, the increase in the value of a is usually insignificant, Figure
4.4. The slight decrease in the value of (l at high frequencies is due
to the presence of cavity space in locations of maximum particle
velocity.

Acouetlcally
.: 1.1
I
--

II
,. ..

25 mm (1 In.) :

flt1erglaee
j ____./_, /
25 mm (1 In.) :
:[ ...../ /
-- -----

air epace
.
./ J -_)3 .
0.5
/' /
0.:3 / / r-:=

/ / 25 mm (1 In.) =-
0.1 / /
flt1erglaee
V I I
4.4 The effect of cavity space on 125 250 500 1.000 2,000 4,000
frequency-a relationship of a 1 in. Frequency (Hz)
thickfiberglass blanket.

4.4 APPLICATIONS OF Porous absorbers such as fiberglass or mineral wool are commonly
POROUS ABSORBERS
used in low-height office partitions, and as wall- and ceiling-mounted
panels. In low-height partitions, Figure 4.5, relatively rigid, abuse
resistant, tackable, high density fiberglass (nearly 95 kg/m3 - 6 lb/
ft3) or wood fiberboards are used. A fabric wrapping provides a
decorative and protective finish over the absorber.
Medium density fiberglass (nearly 48 kg/m3- 3 lb/ft3) is semi
rigid. Typically, it is also fabric wrapped, but is used in protected
locations such as in ceiling suspended baffles, Figure 4.6. Ceiling
suspended baffles are particularly effective in spaces where a
conventional lay-in ceiling is impractical because of interference from
structural components, mechanical services, and other elements, or
where the ceiling is too high. By suspending the absorbers, they are
brought closer to the sound source, which improves their
effectiveness. Suspended baffles can be used in various different
arrangements, Figure 4.7.
Sound Absorbing Materials 69

4.5 Low-height sound absorbing office partitions. Courtesy of PGALArchitects,


Dallas, Texas, with permission.

4.6 Suspended absorbing baffles. Courtesy


ofEssi Acoustical Products Co., Cleveland,
Ohio, with permission.
70 Chapter 4

=== ===
=== ====
=== ===
4. 7 A few alternative
arrangements of suspended Instead of two-dimensional baffles, suspended absorbers can also
baffles in plan. be made of hollow three-dimensional shapes, called space absorbers.
Almost any shape can be fabricated, some of which are shown in
Figure 4.8.
Since a sound wave falls on all surfaces of space absorbers or
suspended baffles, their absorbing property is usually quoted in terms
of the absorption (in sabins) provided by each unit at various
frequencies, not by the value of a.
The total absorption provided by a group of suspended absorbers
is not simply a product of the number of absorbers in a space and the
absorption provided by each unit. For example, if there are 50
suspended absorbers in a room each with an absorption of 15 sabins
at a particular frequency, the total absorption in the room may not be
equal to (50 X 15), i.e., 750 sabins, since the total absorption is also
a function of the spacing of absorbers. A minimum spacing between
absorbers is required to realize their full potential. Small spacings
yield lower than the calculated absorption, particularly at high
frequencies. Manufacturer's data should be obtained for precise
values.
Low density fiberglass (15 kg/m3 to 25 kg/m3 - 1 to 1.5 lb/ft3)
is available in roll or blanket form. Being flexible, it needs adequate
protection. Wood slats (see Figure 3.18), perforated metal panels,
or similar screens are commonly used for the purpose. Fabric
wrapped low density fiberglass can also be used in sound absorbing
banners, Figure 4.9.

4.8 A few of the several possible space absorber profiles.


Sound Absorbing Materials 71

4.9 Sound absorbing banners. Courtesy of Essi Acoustical Products Co., Cleveland, Ohio, with
permission.

4.4.1 Sound Absorbing Roof and Floor Decks

Sound absorbing material may be integrated with a metal roof deck


or floor deck, which is referred to as acoustical metal deck . Although
there are some variations between manufacturers, a typical acoustical
metal deck consists of fluted channels in which the bottom plate is
perforated, Figure 4.10 (some manufacturers make acoustical decks
with perforated webs in place of perforated bottoms). Fiberglass or
mineral wool is placed in the channels. Like any other metal deck,
an acoustical deck is supported on steel trusses or beams. An
acoustical deck has an NRC rating ranging from 0.6 to 0.9 depending
on the depth of air space behind the absorber, and the absorber
thickness.
With sound absorption provided in the deck, there is no need for
a sound absorbing ceiling. Acoustical metal deck is particularly
suitable for spaces where exposed structural framework is
aesthetically acceptable, such as gymnasiums, skating rinks,
manufacturing spaces, etc.
Another effective acoustical deck is provided by rigid wood
fiberboards. Being porous, the boards have sound absorbing
properties. They are available in thickness ranging from 40 to 100
mm (11/2 in. to 4 in.) and are strong enough to span over roof trusses
or other supporting members and carry dead loads due to a concrete
slab. When used in this way, the fiberboard deck functions as a
permanent form work for a concrete slab, in addition to providing a
sound absorbing ceiling, Figure 4.11.
72 Chapter 4

Fil:lergiaaa

ACOUSTICAL ROOF DECK


aurface

ACOUSTICAL ROOF DECK AND ITS SUPPORT


STRUCTURE

Perforated
aurface

ACOUSTICAL FLOOR DECK

4.10 Acoustical metal decks. Courtesy of Epic Metals Corporation, Rankin, Pennsylvania ( 1995 Epic
Metals Corporation, all rights reserved), with permission.

Roof meml:lrane

Lightweight concrete -
Vapor l:larrier -----,

Wood fil:lerl:loard -

I ---

Supporting ateei frame

WOOD FIBERBOARD WOOD FIBERBOARD SECTION THROUGH A ROOF

WITH FABRIC COVERING

(a) (b)

4.11 (a) Woodfiberboard. Samples courtesy of Tectum Inc., Newark, Ohio. Photo by Madan Mehta.
(b) The use of woodfiberboard as an acoustical roof deck.
Sound Absorbing Materials 73

4.5 PANEL ABSORBERS A solid unperforated panel installed against a hard substrate with an
intervening air space acts as a panel absorber or a membrane
absorber, Figure 4.12. When a sound wave falls on such a panel, it
sets the panel into vibration. Since the panel is never fully elastic, it
looses some energy by damping.
The vibration of the panel is similar to the vibration of a mass
attached to a spring, Figure 4.13. A mass-spring assembly looses
energy due to damping forces and gradually stops oscillating after
the exciting force is withdrawn. Precisely the same phenomenon
occurs in the panel, and the panel's vibration is influenced by damping
forces.
Damping is a measure of the resistance of a vibratory system to
sustain vibrations. In the case of a mass-spring assembly, the damping
is caused by the viscosity of the medium. The more viseous the
4.12 A solid unperforated medium, the greater the energy loss by damping, and the quicker the
panel against a hard substrate return of the system to a motionless state. Thus, the damping forces
are larger if the mass-spring assembly oscillates in an oil medium
with an intervening cavity.
rather than in air because of the greater viscosity of oil. Additional
damping in a mass-spring assembly is provided by frictional losses
at the surface on which the mass oscillates. Damping also exists
within the material of the spring.
For a given mass-spring assembly, damping forces increase as
the velocity of the mass increases[421. The velocity of the mass is
maximum at resonant frequency. Therefore, the damping forces have
the maximum effect at the resonant frequency. The resonant
frequency of a mass-spring assembly is equal to its natural frequency
- the frequency at which the mass will oscillate when pulled to one
4.13 A mass-spring assembly. side and, thereafter, left to oscillate on its own5 .
In the case of the panel, the damping is provided by the medium
and at its edge fixing. An additional source of damping exists within
the panel - in the material of the panel - which comes into play
due to bending vibrations in the panel. All materials have some
internal damping. Among panels used in buildings, steel and
aluminum have a low internal damping as compared with plywood
and gypsum board, while lead has the highest internal damping.
Since the damping forces are maximum at the resonant frequency
of the panel, a panel absorber has a maximum value of a at the
resonant frequency. On both sides of the resonant frequency, the
value of a decreases.
In our analogy between the mass-spring assembly and the panel
absorber, the panel represents the mass, and the air behind the panel
acts as the spring. A smaller depth of air behind the panel represents

5 When a vibrating system (such as a mass-spring assembly) is subjected to a

natura/frequency, then the system vibrates with large displacement and velocity
periodically changing external force of the same frequency as the vibrating system's

amplitudes. Theoretically, if there is no damping in the system, the velocity and


displacement amplitudes will grow to an infinite value, causing a breakdown of

This phenomenon is known as resonance, and is similar to the resonance in


the system.

rooms caused by standing waves (see Appendix C). Resonance is of interest not
only in acoustics but also in other fields. For example, flexible buildings (such as
high-rise structures) must be able to resist resonance effects resulting from
earthquake or wind force induced vibrations.
74 Chapter 4

a stiffer spring (a thin cushion). A greater depth of air represents a


more supple spring (a thicker cushion). This analogy provides an
approximate relationship for calculating the resonant frequency of a
panel as:

fres =
1,900
(4.5)
.Jmd

where m is the mass of panel in kg/m2 and d is the air space behind
the panel in millimeters. In the U.S. system of units, the above
expression changes to:

fres =
170
(4.5')
.Jmd

where m is in lb/ft2 and d is in inches. For a 1 mm (1/ in.) thick


gypsum board panel weighing 20 kg/m2 (4 lb/ft2) with 252mm (1 in.)
air space, the resonant frequency of the panel from Equation 4.4), is:

fres =
1 900
= 85 Hz
.J (20)(25)

Similarly, the resonant freuency of a 28 p,auge steel panel


weighing 3.65 kg/m2 (0.75 lb/ft ) with 13 mm ( I in.) air space is
278 Hz. The resonant frequency of a practical panel lies somewhere
between 50 to 400 Hz. A lightweight membrane such as a leather
cloth or a thin plastic sheet will have a slightly higher resonant
frequency because of its lighter weightl43]. Typical panel absorbers
in building interiors include gypsum board partitions, wood paneling,
glass windows and suspended ceilings.
The absorption coefficient of a panel absorber, even at resonant
frequency rarely exceeds 0.3, Figure 4.14. This is because of the
vibrations of the panel, which make it act as a sound radiator.
However, if the panel's material lacks stiffness, i.e., if it is made of
an inelastic (limp) material such as a thin lead sheet or a built-up
roofing felt, the value of a at resonant frequency is nearly 1.0 because
of the large damping in the material.

4.5.1 Effect of a Porous Material in Air Cavity

The absorption coefficient of a panel absorber increases if a porous


absorber is placed in the cavity space between the panel and the
substrate. The inclusion of a porous material in the cavity introduces
damping into the system, which increases its absorption. In our
analogy of the panel absorber with the mass-spring assembly,
introducing a porous material into the cavity is like increasing the
viscosity of the medium in which the mass spring assembly oscillates.
The improvement in the absorption coefficient of the panel is quite
marked near the resonant frequency.
Sound Absorbing Materials 75

f-----+----+---+--1
r-f- With porou& at1eorl1er
0.6

'E
f- Without poroue at1eorber
0.5


\)


0 .4
...... ..
... ...
!F
.. ...
8 ....

0.3 ...,.,
..

.. .

v "'

5
l f----f-- ---""odl---
r----"'-.:-,,,,-
- ..,- ,, .--------j


. .

,........
0.2 .

,
---+-----1----4====::::;" ;,;....:.:::;, ..::.:;.:_;
. .;1"
4.1 4 A pproxi m ate frequency-a .. .. .
0.1
relationship of a 13 mm (112 in.) thick
fJlywood panel backed by 25 mm ( 1 in.)
125 250 500 1,000 2,000 4,000

Frequency (Hz)
air space -with and without a porous
ohsorber in the cavity. Adapted from
'-eference 4.4.

The most commonly used porous material is fiberglass or mineral


wool. It is immaterial where in the cavity space the porous material
is placed, although attaching the porous material to the panel's back
is a convenient way of doing so. The thickness of porous material
must be at least two-thirds of the cavity depth.

4.5.2 Applications of Panel Absorbers

Because of its low value of a, a panel absorber is not a sound


absorbing material in the same sense as a porous absorber. It is
seldom added to building interiors to control noise. However, the
absorption provided by panels such as interior drywall, windows,
wood paneling, wood flooring, suspended reflectors, etc., must be
taken into account in determining the reverberation times of
auditoriums (see Chapter 10).
In certain situations, the absorption characteristics of a panel
absorber are used to advantage. These situations arise in music
practice rooms and rehearsal rooms. Because of their small size
(see Section 10.8 and Appendix C), they tend to be "boomy"
excessively loud at low frequencies. Drywall construction helps to
control the boominess because a typical drywall, with its intervening
air space, functions as a panel absorber, which absorbs low frequency
sounds.
76 Chapter 4

4.6 VOLUME The third and last class of absorbers is known as a volume absorber.
Various other terms used for this absorber are: cavity absorber, cavity
ABSORBERS resonator and Helmholtz resonator. This type of absorber consists
of a volume of air connected to the general atmosphere through a
small volume of air called the neck. A volume absorber is similar to
an open bottle where the volume of air in the bottle is connected to
the outside atmosphere through the air in the bottle's neck, Figure
4.15.
When a sound wave falls on an open bottle, the mass of air in
the neck oscillates back and forth, similar to the mass-spring
assembly. The air in the neck acts as the mass and the air in the body
of the bottle as the spring. As the air in the neck oscillates, it loses

4.15
energy by friction against the walls of the neck. The oscillation of
An open bottle as a the neck peaks at the resonant frequency of the bottle, which implies
volume absorber. that the sound absorption provided by the bottle is maximum at the
resonant frequency.
That an open bottle has a resonant frequency can be verified by
blowing across its neck. The bottle will always produce the same
pitch of sound regardless of who blows across it and how it is blown.
Usually, the resonant frequency of a bottle lies in the low frequency
region, and is given by the following expressions[45l.

55,000 s
fres (4.6)
Jv'V.
=

where S is the cross-sectional area of the neck, v is the volume of the


neck, and V is the volume of the body of the bottle; all dimensions
are in millimeters.
The most commonly used volume absorber is a slotted concrete
masonry unit, more commonly referred to as an acoustical block.
The acoustical block consists of a concrete masonry unit with a long
narrow slot in each cell. Thus, in a two-cell unit, there are two slots.
The slot functions as the neck and the cell space as the interior of the
bottle. For a cell to function as a bottle, the cell cavity must be
closed. Therefore, unlike a normal concrete block in which the cells
are open on both sides, the cells in an acoustical block are closed on
one side. The blocks are laid in a wall with their closed sides facing
up, Figure 4.16.

4.16 The top of an acoustical block.


Photo by Madan Mehta.
Sound Absorbing Materials 77

Since the resonant frequency of an acoustical block depends on


the volume (v) and the surface area of slots (S), it can be varied (by

DD
an octave or so) by changing the width of the slot, its height and its
profile. Two profiles are commonly used - a rectangular profile
and a funnel-shaped profile, Figure 4.17. In fact, several different
types of acoustical blocks are available to suit different applications.

rD
.

4.6.1 Septum and Porous Material in a Volume Absorber


.

The sound absorption property of acoustical blocks can be improved


by adding a metal divider (referred to as metal septum) in each cell.
4.17 Two commonly used
which divides the cell space into two parts, Figure 4.18. The metal
septum improves high frequency absorption of the block6.
profiles of slots in an
acoustical block. The absorption property of a volume absorber is further improved
by adding a porous material (usually fiberglass or mineral wool) in
the neck of the absorber. If the porous material is added to the body
of the volume absorber, the improvement is a little smaller, but more
convenient. Hence, it is this arrangement that is commonly used, as
shown in Figure 4.19, where fiberglass is placed in the cells of the
block.
Where noise control in high humidity spaces is required such as
Metal r;eptum
in indoor swimming pools and manufacturing processes using a great
deal of water, acoustical blocks with metal septum only (without
4.18 Acoustical block
porous fillers) are recommended.
with metal septums in cells.

4.19 Acoustical block with fibrous


filler and metal septum. Photo by
Madan Mehta.

6 The metal septum divides the cell cavity into two parts. In Chapter 5, we
will see why the low frequency sounds pass through the septum, but not the
high frequency sounds, which are reflected back by the septum. Consequently,
high frequency sounds do not see the cavity beyond the septum while the low
frequency sounds behave as if the septum does not exist. In other words, the
metal septum creates two resonating volumes: the smaller volume that extends
up to the septum only, and the larger volume that extends the entire depth of
the cell cavity.
Since the resonant frequency is indirectly proportional to the volume (V)
of the resonator, the smaller volume yields a higher resonant frequency. Without
the septum, this higher frequency resonance would be absent. Thus, the
provision of a septum broadens the frequency range of the acoustical block's
absorption.
78 Chapter 4

4.6.2 Applications of Volume Absorber

Acoustical block s are used extensively for noise control in


manufacturing spaces, school gymnasiums, air conditioning rooms,
swimming pools, skating rinks, test facilities, auditoriums, etc.,
Figures 4.20 and 4.21. Their popularity lies in the fact that the blocks
provide sound absorption as an integral part of a fire rated structural
wall, unlike the porous and panel absorbers, which are mere addenda
to a wall or ceiling.

4.20 G eneral Motors Diesel Test Facility, Detroit, Michigan. Courtesy of


The Proudfoot Company, Monroe, Connecticut, with permission.

4.21 A-1 Classy Theater, University of Texas at Arlington. Photo


by Madan Mehta.
Sound Absorbing Materials 79

The acoustical block of Figure 4.22 is particularly suitable for


load bearing wall applications since it consists of two twin cells.
The cells at the back are open at the top and bottom, like the normal
units, so that they can be grouted and reinforced if required. The
slotted cells in the front part of the block are closed at the top surface
in order to function as resonators.
As stated previously, the maximum sound absorption of a sound
block lies in the low frequency region- in the 125 or 250 octave.
Approximate frequency-a relationships of an 8 in. thick unit with
and without fibrous fillers are shown in Figure 4.23. Sound absorbing
clay units are also available, Figure 4.24.
The use of volume absorbers in buildings is an extremely old
concept dating back to Greek and Roman empires. Large bronze
jars containing enclosed volumes of air (the Greek eceia) were
commonly used in Greek and Roman open air theaters l4 61 . Several
clay vases have been found in the walls and ceilin9s of a number of
old churches in Scandinavia, Russia and Francel4 1.

4.22 Acoustical block with additional cells


C:ll op:n top and l:lottom
for reinforcing and grouting. Photo by Madan
C: ll cloeM at top
Mehta.

1.1 ,------,--r--.---,

0.9 1\\
''
.\.
---- --+--
0.7 ..\
..........
.
.. ..........
\ . ..
. . .-..-.. . . .. ./
.
Without fll:lroue flll:r
0.5 f--_l-\1---t---l----;;;..1-'. . - =
. - .."'"'1-- ..
With fll:lrou e flll:r
/
0.3 f-------+--

0.1 f-------+--

125 250 500 1,000 2,000 4,000

Frequency (Hz)

4.23 Approximate frequency-a relationships of an 8 in. thick 4.24 Clay sound absorbing units.
acoustical block with and without a fibrous filler. Adapted from Sample courtesy of Stark Ceramics Inc.,
Reference 4.8. Canton, Ohio. Photo by Madan Mehta.
80 Chapter 4

4.7 SCREENS AND Apart from providing a protective but acoustically transparent screen
over a porous absorber, a perforated panel also acts as an absorber in
PERFORATED PANELS its own right. When spaced away from a substrate, a perforated
AS ABSORBERS panel behaves as a large ensemble of cavity resonators. In this
ensemble, each hole behaves as the "neck" and the space behind the
panel functions as a set of connected air v olumes, Figure 4.25.
As observed in Section 3.5, perforated panels in aluminum, steel,
plywood and hardboard are commonly used. Apart from surface
density and depth of air space, the resonant frequency of a perforated
panel also depends on the percentage of open area ( optical
transparency). Like an individual resonator, the absorption coefficient
of a perforated panel peaks at its resonant frequency, which usually
lies in the low frequency region.
A perforated panel is seldom used alone. It is more commonly
used as a protective cover over a porous absorber. When used in this
way, the perforated panel improves the low frequency absorption of
a porous absorber because of the panel's own effectiveness as an
absorber at low frequencies. However, at high frequencies, the
perforated panel decreases the porous absorber's effectiveness due
to the lower acoustical transparency of the panel at high frequencies
(refer to Section 3.5). The effect of a perforated panel on the
absorption characteristics of a porous absorber is shown in qualitative
terms in Figure 4.26. For a precise data, refer to the manufacturer's
literature.

Rigid 17acklng
Perforated panel
Poroue
al1sorl1er only

Air cavity

Each lndMdual hole together


with the air cavity functfone
as a cavity resonator (volume
Frequency
al1eorl1er),

4.26 Effect of a perforated panel on


the absorption coefficient of a porous
absorber.
4.25 A perforated panel as an absorber.
Sound Absorbing Materials 81

4.8 MOUNTING Since a cavity space behind the absorber alters its absorption
coefficient, the absorption coefficient of a material depends on how
CONDITIONS it is mounted. Therefore, several mounting conditions have been
standardized. In comparing the absorption data of two materials, it
is important to ensure that the mounting conditions are the same.
Nine different mounting conditions, designated as mounting
types A, B, C, D, E, G, H, I and K, are identified in ASTM Standard
E 795[4-9l. Mounting types A, B, C D and E are for prefabricated
products, and are shown in Figure 4.27. Mounting types G and H
are for drapery, I for spray-on or trowel applied sound absorptive
treatment, and K for office screens.
Type C, D, E and G are further designated by a numerical suffix
which indicates the distance (in millimeters) of the absorbing material
from the substrate to the nearest integral multiple of 5 mm. Thus,
D-20 mounting indicates that there is a 20 mm air space behind the
sample.

5ubetrate
5ubetrate

5ubetrate

Type A mounting: elmulatee Type B mounting: elmulatee Type C mounting: elmulatee eound
normal uee where an abeorblng acouetlcal ceiling tllee adhered abeorblng material behind a
material le laid directly on, or to a hard eurface with an perforated, expanded or other open
attached with mechanical adhesive. facing, and attached to wood
faetenere to, a eubetrate. furring etrlpe. The preferred elzee
for furring &trlf'l' are 20 mm x 40
mm (
3/4 In. x 1112 In.) or 40 mm x
40 mm (1 1/2 In. x 1112 In.), I.e., either
C-20 or C-40 mounting.

Specimen

Mounting
depth

Type D mounting: elmulatee


5ubetrate
al7!'orblng material laid with an air
!'pace. The commonly ueed
mounting condition In measurement
If' D-20 or D-40, I.e., 20 mm ( 3 /4 In.)
or 40 mm (11/2 In.) air epace. Type E mounting: elmulatee a euepended
ceiling with an open plenum above. The
commonly ueed type le E-400.

4.27 Mounting conditions A, B, C, D and E. For mounting conditions G, H, I and K, refer to ASTM Standard
E 795. Adapted from Reference 4.9.
Chapter 4

Mounting type E is for acoustical ceilings. Acoustical ceiling


products are commonly tested with an air space of 400mm (16in.),
i.e., E-400 is the common mounting condition for testing the
absorption property of ceiling products. Although a plenum space is
usually greater than 400mm, it has been observed that the absorption
properties of an acoustical ceiling do not change with a plenum depth
greater than 400mm (16in.).
In using the absorption coefficient of a material for a particular
situation, it is important that the mounting condition be identified .
Without this identification, the absorption coefficient value is
meaningless.

4.9 AC OUSTICAL TILE An acoustical tile ceiling, as shown in Figure 4.28, has become a
standard feature of most commercial, educational and institutional
CEILING interiors. An acoustical tile is a rigid board (usually of high density
fiberglass or mineral wool) fabricated in modular sizes, usually in
600mm x 600 mm (2ft x 2ft) or 600 mm x 1200 mm (2ft x 4ft)
units. Its thickness normally ranges from 13 mm to 40mm (112 in. to
11/2 in.) and it is simply laid-in on an aluminum supporting framework
suspended from the floor or roof.
Being made of a fibrous material, the absorption characteristics
of ceiling tiles correspond with those of a porous absorber. However,
since there is usually a large air space (plenum) above the ceiling,
the tiles also behave somewhat as panel absorbers. In other words,
the sound absorption characteristics of a prefabricated tile ceiling
combines the characteristics of a porous absorber as well as a panel
absorber. Manufacturers of these systems provide the values of a as
a function of frequency in addition to the NRC data.

4.28 A typical ceiling tile interior.


Courtesy of Celotex Corporation,
Tampa, Florida, with permission.
Sound Absorbing Materials 83

Ceiling tiles are usually prepainted. However, to improve their


absorption, the fibers are exposed by piercing, fissuring or scoring
the surface of tiles. A few of the commonly used surface textures of
tiles are shown in Figure 4.29. Note that in-service painting of ceiling
tiles will alter their absorption characteristics. Therefore, if the tiles
must be repainted, it should be done according to manufacturer's
recommendations.
The selection of an acoustical tile for a project depends not only
on its sound absorption characteristics but also on its fire resistance
rating, flame spread index, light reflectance7 and sound insulation.
Ceiling sound insulation is discussed in Section 5 .12.

r -
I
'I

'r

- .-

4.29 Commonly used surface textures of ceiling tiles. Courtesy of Celotex Corporation, Tampa, Florida, with
fermission.

4.10 ACOUSTICAL Acoustical plaster is used in situations where prefabricated sound


absorbing products cannot be used, such as on curved or irregularly
PLASTER AND shaped surfaces. An acoustical plaster is usually a proprietary product
DRAPERIES consisting of lightweight porous or fibrous matrix and a binder. The
plaster mix is usually sprayed on a rigid substrate, but if necessary it
can be troweled in place. Troweling, however, reduces the absorption
because of the resulting compaction. Manufacturer's literature
usually provides the absorption coefficient values.
Although draperies are not an absorbing material in the real sense,
since their absorption is usually very low, they are often used in
concert halls and broadcasting or recording studios to provide a
convenient way of changing the amount of the total absorption in
the room. They can be easily drawn over an otherwise reflective
wall to increase the absorption of the room, or collected together to
reduce the absorption.

7 Refer to a book on building construction for a discussion of fire resistance


rating, flame spread index and light reflectance, e.g., The Principles of Build
ing Construction by Madan Mehta, Prentice Hall, 1997.
84

To provide a reasonable amount of absorption, the draperies must


be made of a heavy interlaced material with a large gather (200% or
more) and hung with an air space behind the drape, to increase low
frequency absorption. The absorption coefficient of a drape increases
with frequency. However, it is difficult to obtain an absorption
coefficient greater than 0.4 even at high frequencies.

4. 11 AUDIENCE AND In addition to the absorption provided by room surfaces, two other
sources of absorption in the room are: occupancy absorption an d
AIR ABSORPTION IN absorption provided by the air in the room. The occupancy absorption
HALLS consists of absorption provided by the furniture (seats in an
auditorium) and human beings.

4.11.1 Audience Absorption

Audience absorption is usually the largest contributor to the total


absorption of an auditorium or concert hall. In fact for a concert
hall, nearlY. 7 5% of the hall's total absorption is provided by the
audience [4 10l.
There are two ways to quantify the absorption provided by the
audience in an auditorium - per seat, or per square meter (per square
foot) of the floor over which the audience is seated. The area method
is more commonly used since the absorption provided per seat is
greatly influenced by the density of seats . Sparsely spaced seats
provide greater absorption (per seat) than those that are closely spaced
because of the diffraction effect.
Some of the other factors affecting audience absorption are the
type of upholstery on seats, the type of dress worn by the audience,
and the slope of the floor. Experimental results indicate that sound
traveling at a small angle of incidence over the audience (grazing
incidence) gets more absorbed than if the same sound wave traveled
at a larger angle of incidence. Consequently, the audience absorption
is higher in a hall with a relatively flat floor than in a hall with a
large floor slope.
Because of the multitude of factors, an audience absorptioq
coefficient can be given only for average conditions. For thesd
conditions, the audience absorption coefficient values, given in Table
4.1, may be used. To obtain the absorption provided by the audience,
simply multiply the appropriate coefficient by the audience area.
The values given in Table 4. 1 are based on the audience area
surrounded by a space of 0. 5 m (20 in.) all around the actual seating
area to account for the diffraction of sound - the edge effect.
To illustrate the use of Table 4 . 1, let us calculate the audience
absorption at 500 Hz in a hall whose seats are organized in two blocks,
each block measuring 9 m x 1 5 m, Figure 4.30. The area to be used
for the audience area is 2(10 m x 16m) = 320 m2. From Table 4. 1,
a = 0.82. Hence, the audience absorption = 320(0.8 8) = 2 8 1.6
(metric) sabins = 3,030.0 (ft) sabins.
Sound Absorbing Materials 85

ummm t!km=u mjj,.--+-


r
i
..
..
i
..
...

.. .
15 16
l r- l
m m

1 l
!
1 -i-o---
-+-
4.30 Audience area used in
:'"L''''''''''''''''''''':
etrlp added to eeatlng
----------------------- -'--

block
o.5 m

calcul ating audience absorption.

Table 4.1 Absorption Coefficients of Audience and Unoccupied Seats

Audience/seats Frequency (Hz)

1 25 250 500 1 ,000 2,000 4,000

Audience on heavily upholstered seats 0.76 0.83 0.88 0.91 0.91 0.89

Audience on medium upholstered seats 0.68 0.75 0.82 0.85 0.86 0.86

Audience on lightly upholstered seats 0.56 0.68 0.79 0.83 0.86 0.86

Unoccupied heavily upholstered seats 0.72 0.79 0.83 0.84 0.83 0.79

Unoccupied medium upholstered seats 0.56 0.64 0.70 0.72 0.68 0.62

Unoccupied lightly upholstered seats 0.35 0.45 0.57 0.6 1 0.59 0.55

Adapted f rom Reference 4. 10.

4.11.2 Air Absorption

The air also absorbs sound, but air absorption is significant only at
high frequencies - 2,000 Hz and above. The air absorption (Aair) is
Tabl e 4.2 Air Attenuation
given by:
Coefficient (m) Aair = mV

Frequency Attenuation coefficient where V is the volume of the room in m3 (or ft3 ), and m is the air
(kHz) Sabins/m Sabins/ft attenuation coefficient, expressed in sabins/m (or sabins/ft). The
value of m is a function of the relative humidity of air, but for a
relative humidity of 40% to 60% (commonly obtained in climate
2 0.009 0.003 controlled interiors), the values of m, as given in Table 4.2 14- 11 1 , may
4 0.025 0.008 be used. Thus, if the volume of a hall is 6,000 m3 , the absorption
8 0.080 0.025 provided by the room's air at 4 kHz = 0.025(6,000) = 150 (metric)
sabins.
86

REFERENCES 4.1 Harris, Cyril: Noise Control in Buildings, McGraw Hill Inc., New York,
1 994, pp. 3.59 and 3.64
4.2 Kinsler, Lawrence, and Frey, Austin: Fundamentals ofAcoustics, John
Wiley & Sons, Inc., New York, 1962, p. 1 8.
4.3 Parkin, P. H., et a!: Acoustics, Noise and Buildings, Faber and Faber
Publishers, London, 1979, p. 50.
4.4 Cracker, M. J., and Price A. J.: Noise and Noise Control, Volume I ,
CRC Press, Cleveland, Ohio, 1 975, p . 2 1 7 .
4.5 Harris, Cyril: Noise Control in Buildings, McGraw Hill Inc., New York,
1 994, p. 3 . 3 1 .
4.6 Mehta, M. : Sound Diffusion, Reverberation and Non-uniformAcousticaJ
Boundaries, Ph.D. Thesis, University of Liverpool, United Kingdom,
1 974, p. 3 .
4.7 Cracker, M. J., and Price A. J.: Noise and Noise Control, Volume I ,
CRC Press, Cleveland, Ohio, 1 975, p . 2 1 9 .
4.8 Proudfoot Company's literature.
4.9 American Society for Testing and Materials: "Standard Practices for
Mounting Test Specimens During Sound Absorption Tests", E-795-93.
4.10 Beranek, L. L. : Concert and Opera Halls - How They Sound, The
Acoustical Society of America, Woodbury, New York, 1 996, p. 62 1 .
4.11 Harris, Cyril: Noise Control in Buildings, McGraw Hill Inc., New York,
1 994, p. 4.7.
Princi pies of
Airborne Sound
Insulation

Any sound that is annoying, distracting, or simply unwanted is


considered noise. The assessment of what "unwanted" is varies from
individual to individual depending on the type of sound, its loudness,
frequency content and the time of occurrence. Even music, if it
occurs at the wrong place or the wrong time, would be considered
noise. In fact, what may be music to one person may be noise to
another.
Issues related to noise control are discussed in Chapter 8.
However, one of the means to control noise in buildings is to contain
it within the space of its origin by constructing barriers around the
space through which sound will not easily transmit. Reducing sound
transmission by constructing barriers of low sound transmission is
referred to as sound insulation. "
In this and the following two chapters, we wilt cover the
principles and practices of sound insulating construction.

5.1 AIRBORNE AND In discussing sound insulation in buildings, we must first distinguish
between airborne sound and structure-borne sound. Most sounds
STRUCTURE-BORNE in buildings are airborne sounds, such as the sounds generated by
SOUNDS human conversation and musical instruments, Figure 5.1. Fans,
motors, machinery, airplanes, and automobiles are some of the other
sources that produce airborne sound.
Structure-borne sound is produced by an impact of some sort on
building elements - walls, floors, roofs, etc. The impact causes
building elements to vibrate, and as they vibrate, they radiate sound.
Since it is impact related, structure-borne sound is also referred to as

87
88 Chapter 5

5.1 Examples of airborne sound


sources.

impact sound. Thus, when a nail is hit on a wall, or a person walks


on a suspended floor, structure-borne sound is produced, Figure 5.2.
Some other examples of structure-borne sounds are vibrating
machinery rigidly connected to a floor, plumbing pipes attached to a
wall, and the slamming of a door.
In other words, a structure-borne sound originates in an impact
or vibration1 producing source that is in contact with a building
component. The building component works as an amplifier of the
sound generated by the impact or vibrating source; the impact or
vibrating source itself may not create much sound. For example, a
vibrating water tap does not create much sound of its own, but if it is
rigidly connected to a wall, its sound is amplified several fold because
of the vibrations it produces in the wall.

5.2 Examples of structure-borne sound


sources.

1 Although both impact and vibration producing sources produce structure-borne!


sounds, the means used to mitigate their effects are generally different. Therefore,
impact sound insulation is dealt with in Chapter 7, and vibration isolation is covered
in Chapter 15.
Principles of Airborne Sound Insulation 89

Similarly, the sound produced by a tuning fork is relatively faint


when held in the hand, but if its handle is placed on a table, the
sound is greatly amplified through the vibrations of the table, Figure
5.3. Indeed, the sound produced by a string instrument, such as a
guitar, is amplified several fold by the wooden body (sounding board)
on which the strings are mounted2 .
Once the structure-borne sound is produced by a building
component, it becomes airborne sound and reaches the receiver as
such. Although most sources produce either airborne sound or
structure-borne sound, several sources produce both airborne and
structure-borne sounds. For instance, a floor-mounted machine with
rotating parts generally creates both airborne and structure-borne
sounds. The fans, rotating belts and other moving parts create
airborne sound, while the vibrations of the entire machine create
structure-borne sound, Figure 5.4.
Despite several similarities, sound insulating construction for
airborne sound differs from that of structure-borne sound. Structure
5.3 Sound amplification of a
borne sound insulation is discussed in Chapter 7. The principles of
airborne sound insulation are covered in this chapter; its practical
tuning fork by a table top. aspects are discussed in Chapter 6.

(( (
5.4 Structure-borne and airborne
sounds from a floor-mounted machine.

2 In fact, the so-called amplification is not really an amplification but a more


efficient conversion of vibrational energy into sound. Vibrations are more efficiently
converted into sound if the dimensions of a sound source are at least of the same
order of magnitude as the wavelength of sound. Thus, an efficient radiator of
sound at 100 Hz must be nearly 3.4 m (11 ft) in both its dimensions (length and
width). If its dimensions are less than 3.4 m, it will still radiate sound at 100 Hz,
but with a lower efficiency. This explains why low frequency (bass) loudspeakers
have to be larger than high frequency speakers (tweeters). The size of a tuning
fork or the diameter of a guitar string is too small for the wavelength of sound they
produce. Consequently, when a tuning fork or string (that is not connected to a
sounding board) vibrates, it does not produce much sound.
90 Chapter 5

5.2 SOUND In Section 3.1, we defined transmission coefficient 't as that fraction
of incident sound energy that is transmitted through a component.
TRANSMISSION LOSS Thus, if 't is 0.2, the transmitted sound is 20% of the incident sound.
't = 0 means that no sound is transmitted through the component. 't
= 1 implies that the component is acoustically transparent since all
sound is transmitted through it, e.g. an open window.
The values of 't for commonly used walls and floors varies from
2
approximately w- to w-8. For instance, for a 1/4 in. thick glass
sheet, 't is nearly 7.8x10-4 , and for a 6 in. thick concrete slab, 't is
approximately 6.3x10-6. Since these are rather awkward numbers,
the transmission property of a component is given by a quantity
referred to as the sound transmission loss or simply the transmission
loss (TL), which is defined as:

1
TL 10 log (5.1)
T
=

According to the above relationship, if 't = 1, TL = 0, and if 't =


0, TL is infinite. Obviously, it is impossible to obtain a component
with 't exactly equal to zero, although for a very thick and heavy
component, 't is nearly zero, and hence its TL is a large number. TL
values are given in decibels (dB).
Note that according to the above definition transmission loss is
the loss in sound pressure level that occurs as the sound passes
through the panel. For example, if the sound pressure level on the
source side of a panel is 90 dB, and 40 dB on the receiver side, the
TL of the panel is 50 dB, Figure 5.5. If the sound pressure level on
the source side is 70 dB, then the sound pressure level on the receiver
side of the same panel will be 20 dB. The greater the TL of a panel,
the greater the sound insulation provided by it. A panel with a TL of
50 dB is a better sound insulator than one with a TL of 45 dB.
If we substitute the value of 't for a 6 mm e/4 in.) thick glass
sheet, and for a 200 mm (8 in.) thick concrete slab in Equation (5.1),
we find that their TL values are 31 dB and 52 dB respectively.
Detailed calculations are shown in Examples 5.1 and 5 .2.

Sound preeeure
TL =
SPLeo rce eide - SPLrecelver elde level =90dB
u
SOURCE SIDE

5.5 Definition of the transmission loss of a component.


rtinciples of Airborne Sound Insulation 91

Example 5.1 TL Value from the Value oft

Determine the transmission loss of: (i) a 6 mm (1/4 in.) thick glass sheet for which t=7.8 x 104 , and (ii)
for a 150 mm (6 in.) thick concrete slab for which t=6.3 x 106.

Solution: (i) From Equation (5.1), TL=10 log [11(7.8 x 104)]=10 log (1,282)= 31 dB.
(ii) TL=10 log [11(6.3 X 10"6)]=10 log (158,730)=52 dB.

Example 5.2 Value oft from the Value of TL

Determine the value oft for a component whose (i) TL=30 dB, (ii) TL=60 dB, and (iii) TL=36 dB.

Solution: (i) 30=10 log (1/'t). Hence, log (1/t)=3.0. 1/t=103 , or t=103 = 0.001
(ii) 60=10 log (1/t). Hence, log (lit)=6.0. 1/t=106, or t=106 = 0.000 001
(iii) 36=10 log (lit). Hence, log (lit)= 3.6. t= w-3 6' i.e., t=2.5 X 10"4 = 0.00025

5.3 SOUND Since a thick porous absorber provides high sound absorption, one
might assume that it can also provide a high transmission loss. After
ABSORPTION AND all, if most of the sound is absorbed by the absorber, there is little
SOUND INSULATION left to go through it. The fact, however, is that in spite of its high
absorption, the transmission loss of a porous absorber, even a thick
one - say a 100 mm (4 in.) thick fiberglass blanket- is extremely
small. The reason lies in the fact that the absorption coefficient (a)
is expressed on a linear scale, while the TL i s expressed
logarithmically.
For instance, the value of a for a 100 mm (4 in.) thick fiberglass
blanket at high frequencies is approximately 0.95, that is 95% of
incident sound is absorbed by fiberglass. This gives a transmission
coefficient (t) equal to (1 - 0.95) =0.05. Therefore from Equation
(5.1), the TL of a 100 mm thick fiberglass blanket is approximately
13 dB, which is extremely low3 . By comparison, the TL of a 6 mm
(1I4 in.) thick glass sheet is 31 dB, as shown in Example 5.1. A 13
mm cit in.) thick gypsum board provides a TL of nearly 30 dB.
In Tact, the TL of a porous material, regardless of its high
absorption, is relatively low. For example, the TL of a 100 mm (4
in.) thick fiberglass blanket- even of high density - is 4 to 5 dB at
low frequencies. As a general rule, if air can pass through a material,
so can sound. Since, a porous absorber lets air through, it is a poor
sound insulator. Thus, for a material to be a good sound insulator, it
must be of impervious and airtight construction. As we will see in
Section 5.8, even a small void or penetration in an otherwise highly
sound insulating construction can degrade its TL substantially.

3 Note that sound insulation should not be expressed in terms of percentage


reduction in noise which, although theoretically correct, is highly deceptive. A
95% reduction in noise levels appears to be a very large reduction, but in terms of
sound insulation, it is a mere 13 dB.
92 Chapter 5

5.4 SINGLE-LEAF Building panels usually consist of several (layers) leaves of different
materials. For example, a typical lightweight wall consists of two
PANEL- THE MASS leaves of gypsum board, one on each side of a cavity space formed
LAW by the studs. We shall refer to a panel with two or more leaves as a
multi-leaf panel. On the other hand, a brick wall or a concrete slab,
since it consists of one leaf of the same material, is referred to as a
single-leaf panel.
If a single-leaf panel behaves as if it is composed of several
interconnected parts, which oscillate independent of each other under
the influence of sound pressure fluctuations, as shown in Figure 5.6,
it can be shown mathematically that the TL of such an idealized
panel is given by:

(5.2)

where f is the sound frequency and m is the surface mass of the


panel in kg/m2. If the surface mass is expressed in lb/ft2, the above
expression changes to:

(5.2')
5.6 The vibrations in a barrier that
obeys the mass law behave as if the
barrier consists of several independent
In other words, if a single-leaf panel behaves as an ensemble of
but connected parts - a limp mass irtdependent parts, its TL is directly proportional to its surface mass
model. In an ideal limp panel, while and sound frequency. That is, the TL increases with an increase in
one part of the panel may be vibrating, the surface mass of the panel and also with an increase in sound
the other parts may be stationary. frequency. Since the quantity [20 log (2)] is equal to 6, Equation
(5.2) shows that doubling the surface mass or doubling the frequency
gives a 6 dB increase in the TL of a panel. This is called the mass
law. Thus, according to the mass law, if a 50 mm (2 in.) thick panel
has a TL of 37 dB at certain frequency, a 100 mm (4 in.) thick panel
of the same material and at the same frequency will have a TL of 43
dB, a 200 mm (8 in.) thick panel will have a TL of 49 dB, and so on,
Figure 5.7. Example 5.3 illustrates the use of Equations (5.2) and
(5.2').

37 dB 43dB 49 dB 55dB

50 mm 100 mm 200 mm 400 mm


(2 in.) (4 In.) (8 In.) (16 In.)

5.7 Theoretical mass law and the thickness of


a component.
Principles of Airborne Sound Insulation 93

Example 5.3 - Transmission Loss of a Single Leaf Barrier That Obeys the Mass Law

Using the mass law, determine the TL of (i) a 50 mm in. thick concrete panel at 500 Hz and 1 kHz, and
(ii) an 8 in. thick concrete panel at 500 Hz. The density of concrete=2,400 kg/m3 (150 lb/ft3).

Solution: (i) Since the density of concrete is 2,400 kg/m3, the surface mass (m) of 50 mm thick concrete
panel=2,400 x 0.05=120 kg/m2. Hence, from Equation 5.2,

TL500= 20 log [500 x 120]- 47=20 log(60,000)- 47=49 dB. Similarly,

TL1,000=20 log [1,000 x 120]- 47=55 dB.

(ii) For an 8 in. thick panel, m=100 lb/ft2. From Equation (5.2'),

TL500 =20 log [500 x 100]- 33=20 log [50,000]- 33=61 dB.

Note that we could have obtained the same result from (i) above by using the fact that doubling the
surface mass gives a 6 dB increase in TL. Thus, a 4 in. thick concrete panel would give a TL of (49 +
6)=55 dB, and an 8 in. thick panel would give a TL of (55 + 6) =61 dB.

5.5 SINGLE-LEAF The mass law is based on the assumption that each part of the panel
oscillates independent of the other as if it slides past its neighbors
PANEL- THE without affecting them in any way. Indeed this is possible only if
COINCIDENCE EFFECT the panel has zero stiffness, i.e., it is a completely inelastic (or a
limp) element. That is why the mass law is also referred to as the
limp mass law.
A real panel is not limp. The individual parts of a real panel are
by no means independent of each other, but are coupled to each other
by elastic forces. Therefore, the bending of one part of the panel
leads to the propagation of a bending wave in the panel. In other
words, the stiffness of the panel resulting from its elastic
interconnections among parts affects its transmission loss.
The stiffness of the panel reduces its TL, i.e., the stiffer the panel,
the smaller its TL. For example, Figure 5.8 shows the TL-frequency
relationship of two plywood panels with equal surface mass. The
grooved panel, because of its lower stiffness, gives a higher TL.
For most materials, surface mass and stiffness are related
quantities. For example, if the surface mass of a panel of a given
material is increased by increasing its thickness, its stiffness also
Frequency
increases. Since the increases in surface mass and stiffness have
opposite effects on a panel's TL, the mass law is not completely
valid fo1 a real panel. Deviation from the mass law is further
5.8 Approximate TL-frequency accentuated by the energy-absorbing property (damping) of a panel.
relationships of solid and grooved In other words, the transmission loss of a single leaf panel is not
plywood panels of equal surface mass. simply a function of the mass of the panel alone, but of the following
Adapted from Reference 5.1. three factors:
94 Chapter 5

Mass

Stiffness

Damping

However, it must be noted that mass is the most important factor.


Consequently, the TL of a real panel4 increases on an average by
nearly 5 dB per doubling of frequency (i.e., per octave), or per
doubling of mass - not 6 dB - as indicated by the mass law.

5.5.1 Coincidence Frequency

Laboratory measurements indicate that the TL-frequency relationship


of a single-leaf panel is generally of the form shown in Figure 5.9.
This relationship may be divided in two frequency regions5. In
Region 1, the TL of the panel is determined by its surface mass. It is
in this region that the mass law holds and the TL of a panel increases
by nearly 6 dB per octave. Therefore, this region is referred to as the
mass controlled region.
In Region 2, the TL is governed by the stiffness of the panel.
Hence, Region 2 is known as the stiffness controlled region. This
region is characterized by a major dip in TL (an acoustical hole),
referred to as the coincidence dip. This dip may be as much as 15
dB for stiff materials, and is the result of the resonance produced
when the wavelength of bending waves in the panel coincides with
the wavelength of sound incident on the panel. Thus, coincidence is
a special type of resonance. Bending waves in the panel are similar
to the waves produced in a rope that is shaken at one end.

Low damping

--

EKtent>lon of ___
-- Medium damping
the mat>e; law _ _ .. --- ::
.
Large damping
----- __

Ill
Ill
..9
s:::
0
'iii
Ill
.E
Ill
s:::

REGION I REGION 11
MASS STIFFNESS
CONTROLLED CONTROLLED

5.9 TL-frequency relationship Frequency Coincidence freuency


of a single-leaf panel.

4 The transmission loss through a panel is also a function of the angle of incidence
of sound. The discussion presented here is for diffuse incidence since it is this
that is relevant in buildings.
5 At extremely low frequencies, which are usually not encountered in building
noise problems, the TL increases with decreasing frequency . This region is to the
left of Region 1 in Figure 5.9. For a more comprehensive treatment of TL
frequency relationship, see Reference 5.2.
principles of Airborne Sound Insulation 95

The frequency at which the coincidence dip occurs is called the


coincidence frequency. The coincidence frequency of a panel is a
function of three factors: (i) the density of material, (ii) the modulus
of elasticity of material and (iii) the panel's thicknessf5 3l.
The density and modulus of elasticity are constants for a given
material. Therefore for a given material, the coincidence frequency
is a function of the panel's thickness only- in fact, it is inversely
proportional to the panel's thickness. That is:

1
Coincidence frequency oc
Panel thickness

5.5.2 Thickness of Panel and Coincidence Effect

Since the coincidence dip affects the TL of a panel adversely, one of


the ways of minimizing its effect is to push the coincidence frequency
out of the frequency region that is of interest to us. Since the
coincidence frequency depends on the thickness of the panel, it can
be increased or decreased simply by changing the panel's thickness.
Table 5.1 Approximate Coincidence For instance, a 3 mm eI a !n.) thick aluminum sheet has a coincidence
(requencies of Selected Materials frequency of nearly 4 kHz, Table 5.1. If its thickness is reduced to
1.5 mm e/ 6 in.), its coincidence frequency becomes 8 kHz, which
1
is above the frequency range of most noise in buildings. The
Material Thickness Coincidence
coincidence frequencies of commonly used materials are given in
(mm) ( in.) frequency Table 5.1.
(Hz)
Decreasing panel thickness, however, reduces the surface mass
of the panel, and hence its TL. Thus, care must be exercised in
Steel 3 118 4,000 increasing the coincidence frequency by decreasing the panel's
Aluminum 3 118 4,000 thickness.
Concrete 200 8 110 For concrete and masonry, which are normally used in large
Brick 200 8 115 thickness, the coincidence frequency is low, usually around 100 Hz.
Glass 3 118 5,000 Frequencies below 100 Hz are seldom of concern in most building
Gypsum 13 1/2 2,500 noise problems.
Lead 3 1/8 17,000 In practice, the coincidence frequency is of concern only in
Plywood 13 1/2 1,700 relatively thin panels, such as gypsum board, plywood, glass, etc.,
Vinyl 3 118 10,000 because, as we see from Table 5.1, the coincidence frequency for
these panels lies between 1,000 to 3,000 Hz. This is a critical region
as considerable speech information is present here, and the ear's
sensitivity is also high in this region.

5.5.3 Panel Damping and Coincidence Effect

Another way of manipulating the effect of coincidence dip is to


increase the damping in the panel. The greater the damping, the
flatter the coincidence dip, see Figure 5.9.
Steel and aluminum have the least internal damping among
building materials, and sheet lead, the highest. Sheet lead has the
additional advantage of a large surface mass and a low stiffness- a
perfect combination of properties to provide a high transmission loss.
However, because of its cost and environmental concerns, the use of
lead is rare in contemporary building construction.
Damping can also be added to panels through the use of
viscoelastic materials. Various types of visco-elastic materials are
available. Usually, these are mastics that can be sprayed or rolled on
the panel. Another form of viscoelastic material is a polymeric film
96 Chapter 5

adhered to the panel, such as an acrylic sheet 3 mm etg in.) thick.


Some films are manufactured in peel-and-stick format, so that they
are easily laminated to the panel.
Both mastics and films are usually sandwiched between two
sheets of the same or dissimilar materials. For example, a
manufacturer makes plywood and waferboard structural panels with
a sandwiched damping layer, Figure 5.10. Damped plywood panels
provide a higher TL than regular plywood or waferboard panels of
the same thickness.

Plywood ---.:=...
Damping laminate
Plywood

5.10 Two sheets of plywood


bonded to gether with an
intervening visco-elastic
(damping) laminate. Sample
courtesy of Greenwood Forest
Products. Photo by Madan
Mehta.

The increase in (airborne) transmission loss due to the use of a


damping layer in plywood is small because a plywood panel has
adequate internal damping of its own. However, the damping layer
gives substantial improvement in structure-borne transmission loss
of a plywood panel [ 5 .4J.
Note that damping has little or no effect in the mass controlled
region of the TL-frequency relationship, except for the added weight
of the damping material. It is effective only in the stiffness controlled
region. Damping absorbs bending and vibrational energy at
- ,---
resonance by converting the energy into heat. Thus, added damping
f-- Glaee eheet is particularly effective with stiff materials with little internal
.-Jo---+-- Glaee ehe damping, such as steel, aluminum, and glass.
,._+--- Plaetlc Laminated glass provides about 3 dB higher TL than an annealed
lnterlayer glass of the same thickness, because the plastic interlayer, as shown
in Figure 5.11, serves as damping. That is why laminated glass is
commonly specified for glazing in noisy areas, such as downtown
offices, airport hotels, etc.
A new class of damped material that is being increasingly used
is loaded vinyl. Loaded vinyl consists of a polymer mixed with a
heavy weight inorganic material, such as barium sulfate or calcium
5.11 The intermediate plastic carbonate, to obtain a heavy but limp material that can be hung like
layer in a laminated glass a curtain off a frame. It is normally used to enclose noisy machines
functions as the damping layer. in manufacturing spaces, Figure 5.12.
principles of Airborne Sound Insulation 97

5.12 A loaded vinyl curtain surround


ing noisy equipment. An STC of nearly
20 is achievable with the help of a
loaded vinyl curtain (see Section 5.8for
STC explanation).

5.6 TRANSMISSION Since the average increase in TL of a single-leaf panel is only about
5 dB per doubling of panel thickness, the law of diminishing returns
LOSS OF A TWO-LEAF applies here. For instance, the average TL of a 100 mm (4 in.) thick
PANEL brick wall is 40 dB. To obtain a 5 dB increase in TL, we need a 200
mm (8 in.) thick brick wall. A further increase of 5 dB requires a
400 mm (16 in.) thick wall. Since weight and thickness of elements
are important considerations in contemporary construction, the TL

-
of a single-leaf panel reaches a practical limit fairly rapidly.
If, however, a panel is constructed of two separate leaves with

n_"
an intervening cavity, its TL is higher than that provided by a single

40 ,.
100mm ( 4 In.)
"'"""
200 mm (8 In.)
leaf panel of equivalent weight. For example a brick wall, with two
100 mm (4 in.) thick leaves separated by a cavity, has a higher TL
than a 200 mm (8 in.) thick wall although their weights are the same.
THICK BRICK WALL THICK BRICK WALL

5.6.1 The Importance of Decoupling the Leaves of a Panel

Ideally, the two leaves must not be connected together. If we could


somehow separate the two leaves of a cavity wall completely with
no interconnection between them of any kind, the TL of such a wall
BRICK CAVITY WALL WITH TWO 100 mm would simply be the sum of the transmission losses of each individual
( 4In.) THICK LEAVES leaf. Thus, if a 100 mm (4 in.) thick brick wall has a TL of 40 dB, a
TL of thl6 cavity wall can be 80 dB provided cavity wall with two 100 mm (4 in.) thick completely uncoupled
the two brick ieave6 are completely brick leaves will have a TL of (40 + 40) = 80 dB, while a 200 mm (8
uncoupled - a virtual lmpo661blllty in.) thick brick wall has a TL of only 45 dB, Figure 5.13.
However, it is virtually impossible to completely decouple two
5.13 Approximate TL values of a 4 in. leaves of a panel. For instance, even if there are no intermediate
connecting elements, the two leaves of a cavity wall are connected
thick brick wall, 8 in. thick wall, and a
at the bottom and the top by the supporting floor and ceiling. In
brick cavity wall with two 4 in. thick practice, a cavity wall has several intermediate connecting elements
leaves. other than those at the top and bottom.
98 Chapter 5

In addition to the above mechanical connections, the cavity air


functions as a major coupler between leaves, transferring vibrational
energy from one leaf to the other. Remember from Section 4.5 that
the cavity air functions as a spring. Thus, a two-leaf panel behaves
as two masses joined together by a spring in which the leaves
represent the masses, and the cavity air represents the spring.
As a result of the couplings, the actual TL of a two-leaf panel is
much smaller than that obtained by adding their individual
transmission losses. However, a two-leaf panel provides a better
alternative to a single-leaf panel of the same total weight.

5.6.2 The Importance of Dissimilarity of Leaves

Does a two-leaf panel experience coincidence dip similar to a single


leaf panel? The answer is yes. In fact, the coincidence phenomenon
is more complex in a two-leaf panel than in a single-leaf panel.
However, the coincidence effect can be decreased by ensuring that
the two leaves of the panel have different coincidence frequencies.
If the two leaves are identical, i.e., if they are of the same
thickness and material, and are fastened in the same way (i.e., by the
same size and spacing of screws or nails), they will have the same
coincidence frequency. Consequently, the coincidence effect is
accentuated, which increases the coincidence dip. Using leaves of
either two dissimilar materials or of unequal thickness reduces the
severity of this dip. Using two different thicknesses is usually more
practical than using dissimilar materials.

5.6.3 The Importance of a Porous Absorber in a Cavity

There is an additional dip to be considered in the TL-frequency


relationship of a two-leaf panel. This dip is due to the resonance of
the cavity air. Since the cavity air behaves as a spring, it has its own
resonant frequency. The smaller the cavity depth, the greater the
stiffness of the cavity, and hence the resonance occurs at a higher
frequency. As the cavity space is increased, the resonance occurs at
a progressively lower frequency.
There are two ways to reduce the effect of the dip caused by
cavity resonance. The first way is to introduce a porous absorber
such as fiberglass (or an open cell plastic foam) into the cavity. A
porous absorber has the same effect in reducing air resonance dip as
a viscoelastic material has in reducing the coincidence dip in a single
leaf panel.
To achieve full benefit, fiberglass must cover the entire length
and width of the cavity. The thickness of fiberglass need not exceed
three-quarters the depth of cavity, as any further increase provides
an insignificant improvement in TL. In fact, fiberglass must not be
packed into the entire depth of the cavity, as that will enhance the
coupling between the two leaves of the panel, reducing its TL.

5.6.4 The Importance of a Large Cavity Depth

The second method of reducing the effect of cavity resonance is to


increase the depth of the cavity. The resonant frequency of cavity
air is inversely proportional to cavity depth6. If we increase the
nciples of Airborne Sound Insulation 99

depth of cavity sufficiently, it may be possible to push the cavity


resonance dip to a frequency lower than that of interest to us.
Therefore, cavity depth should be increased as much as is practical.
Glaee In fact, a narrow cavity depth in a two-leaf panel may yield worse
Glaee --+--+--1 TL performance than a single-leaf panel because of the cavity
resonance dip. This explains why a conventional insulating glass
unit with its narrow (typically 13 mm, i.e., 1/2 in.) air space has a
worse low frequency TL than a single glazed window of equivalent
glass thickness.
This is shown in Figure 5.14 which gives the TL-frequency
INSULATING GLASS UNIT relationships of a 13 mm (11 in.) thick monolithic glass and an
insulating glass unit (IGU) with two 6 mm (1/4 in.) thick glass sheets
with a 6 mm (112 in.) air space. The TL dip a1 low frequency in an
IGU is due to cavity resonance. Included also in this diagram is the
relationship for an IGU with two 6 mm (1/4 in.) thick laminated glass
sheets with a 6 mm (1/ 2 in.) air space, confirming the superior
performance of laminated glass.

IGU with two 6 mm (114


50 In.) laminated glaee and
12 mm (112 In.) cavity

Ill IGU with two 6 mm


Ill
.2 (114 In.) glaee and 12
s::: mm (112 In.) cavity
0
'iii
Ill
E
Ill
s:::

1-

10 '---'-----"----'---_L____L_---'---"---'-'---.J
125 250 500 1.000 2,000 4,000
Frequency (Hz)

5.14 TL-frequency relationships of 13 mm (112 in.) thick monolithic glass, an 1GU


with two 6 mm ( 114 in.) thick glass sheets, and an I GU with two 6 mm (1/4 in.) thick
laminated glass sheets. Adapted from Reference 5.6.

6 The resonant frequency of cavity air in a two-leaf panel is given by the


following approximate relationship:
fres k[(m1 + m2)/dm1m2]05 where m1 and m2 are the surface masses of the
=
2
panels in kg/m , d is the depth of cavity in meters, and k 60 for an empty
=

cavity, 40 for a cavity with a porous material [Reference 5.5].


100 Chapter 5

5.6.5 TL of a Two-leaf Panel - Summary of Factors

We are now in a position to summarize the factors that help maximize


the TL of a two-leaf panel. These are:


Surface mass: This should be maximized as much as
possible.


Stiffness: Although it is usually difficult to control the
stiffness of panels it is important to remember that materials
of low stiffness provide a higher TL.


Decoupling of leaves: This is the single most important factor
affecting the TL of a two-leaf panel. In a wood or metal
stud assembly, decoupling is achieved by using two separate
stud walls or a resilient channel, or both; see Section 6.5 for
details. In brick cavity walls, the ties should be as lightweight
and flexible as structurally permitted.

Cavity depth: This should be maximized as much as possible.


To achieve a significant improvement in low frequency TL,
a cavity space of at least 100 mm (4 in.) is recommended.


Porous Absorber in Cavity: A layer of porous absorber in
the cavity improves TL, provided that the leaves are
adequately decoupled.


Dissimilar leaves: Using leaves of either different material
or different thickness improves the TL of the panel.

5.7 PANELS OF THREE The reasons that account for the higher TL of a two-leaf panel over
a single-leaf panel are the very reasons that account for the
OR MORE LEAVES improvement of a three-leaf panel over a two-leaf panel. In fact,
one can go on adding more and more leaves to improve the TL of a
panel. Cost, weight and space are usually the limiting factors.
Additionally, double cavity resonances are hard to predict. That is
why the use of a three-leaf panel (two cavity spaces) is rare except
in glazed windows7.

7 The primary reason for using a triple glazed window is to obtain a high R
value, not a high sound transmission loss. The R-value of a construction
assembly is an index of its thermal performance; refer to a book on building
construction, e.g., Madan Mehta: Principles ofBuilding Construction, Prentice
Hall, 1997.
,nnciples of Airborne Sound Insulation 101

5.8 SOUND For most critical sound insulation problems, a detailed TL-frequency
relationship of a panel is required. However, a TL-frequency
TRANSMISSION CLASS relationship is not easily amenable to comparing the insulating
effectiveness of various panels. For this purpose, a single number
rating of a panel's TL is desirable.
Obviously, one way to obtain a single number TL rating is to
average the TL values of a panel at various frequencies. But an
average value is usually misleading since it ignores the high and low
8
points in a relationship . A single number TL rating that addresses
the above concerns is called the sound transmission class (STC).
To obtain the STC of a panel, we first obtain the measured TL
frequency data of the panel. The TL of a panel is measured in a set
up consisting of two reverberation chambers (see Figure 3 . 28),
sharing a common wall, as per ASTM standard E 90[571. The
common wall has an opening in which the test panel is installed,
Figure 5. 15, and fully sealed along its perimeter. One of the two
rooms is the source room, and the other the receiving room. The
construction of the two room set-up is such that the sound will go
mainly through the test panel, not through other paths. In other words,
the flanking transmission (see Section 6.6) is minimal.
A broad-band noise is generated in the source room with the
help of loudspeakers, and the noise levels in both rooms are
simultaneously measured in one-third octave bands. The difference
between the two corresponding levels is the TL of the panel9.
The TL-frequency relationship of the the panel (in sixteen one
third octave bands from 1 25 Hz to 4 kHz) is now plotted with
frequency on the horizontal axis and the TL on the vertical axis. The
STC of the panel is obtained by comparing the measured TL values
of the panel with a standard TL-frequency relationship, called the

- Teet panel

SOURCE RECEIVING
ROOM ROOM

oudepeaker
5.15 Set-up for measuring the TL of a
panel. D 1!:Meaeurln{!
(\Uipment

8 See section entitled "STC Value From Transmission Loss Data" at the end of
this chapter.
9 The TL of the test panel, when obtained by this procedure, is also a function
of the area of the panel and the amount of absorption present in the receiving
room; see Equation (8.4), Section 8.5.
1 02 Chapter 5

STC contour. The STC contour consists of three straight line parts
with different slopes. The first part has a total rise of 15 dB, the
second part has a rise of 5 dB and the third part is flat, Figure 5.16,
which approximately simulates the sensitivity of human ear.
The STC contour is drawn on a separate transparent sheet. The
divisions on the horizontal and vertical axes of the STC contour must
be the same as those of the TL-frequency relationship. The STC
contour is now laid over the measured TL-frequency relationship of
the panel in such a way that the vertical axes of both the STC and
TL-frequency plot are aligned. The contour is now moved as far up
as possible until the following two conditions are satisfied[5.81:


The sum of deficiencies of TL at 16 frequencies do not exceed
32 dB -an average of 2 dB per one-third octave.

The maximum deficiency at any point should not exceed 8
dB.

Both the above conditions must be satisfied, and when they are,
the STC of the panel is the TL value at 500 Hz on the STC contour.
The STC of a panel is given in a whole number and the unit "dB" is
omitted. Thus, a panel is rated as STC 41, not 41 dB. The greater
the STC value, the more insulating the panel. The method of
determining the STC of a panel is explained with the help of an
example at the end of this chapter.

5.8.1 Limitations of the STC Rating

Like all single number ratings, STC has its limitations. First, it does
not give any idea as to the magnitudes and locations of dips in the
TL of the panel. Second, it is limited to the 125 Hz to 4 kHz region
- the frequency region of speech. There are several indoor noise
situations that extend beyond this frequency range (e.g., music) to
which STC is not fully applicable. Despite its limitations, STC is a
convenient measure to rank order the airborne TL of panels. The
TL-frequency data of commonly used panels and their STC ratings
are given in Appendix J.

1 dB per one
third octave
m :3 dB per one- 5dB
- third octave
Ill
Ill
_Q
<::
0
ii'i 15dB
Ill
e:
Ill
<::
<1:1
.....
1-

125 200 :315 500 800 1,250 2,000 :3,150


16 0 250 400 6 :30 1,000 1,6 00 2,500 4,000
Frequency (Hz)

5.16 Standard STC contour.


Principles of Airborne Sound Insulation 103

5.9 OVERALL Most panels are an assemblage of two or more panels with different
TL (or STC) values. For example, a wall assembly usually consists
TRANSMISSION LOSS of openings - doors and windows. The openings generally do not
OF AN ASSEMBLY have the same TL as the wall. Similarly, a roof may have skylights
of a lower TL value than the roof. How do we determine the overall
TL of an assembly, which consists of two or more panels of different
TL values?
The problem can be solved using Table 5 .2, which is based on
the mathematical formulation of Appendix F. Table 5.2 assumes
that an assembly consists of two panels, the main wall and an opening.
5m To use the table, use the following three steps.

1 Oponlne ._
Openlng'e n.
=

=
10%
20 dB

Calculate the opening area as a percentage of the total wall
area. For example, if the wall has overall dimensions of 5 m
x 3 m, of which the opening measures 1 .5 m x 1 m, the
opening area is 1 0% of the total wall area ..


Determine the difference between the TL of the wall and the
opening. Assume that the wall ' s TL = 30 dB and the
opening's TL = 20 dB , as shown in Figure 5 . 1 7 . The
Wall'e TL 30 dB
=
difference between the two TL values is (30 - 20) = 1 0 dB .

5.17 Refers to the text.


Entering the TL difference and the opening percentage in
Table 5.2, we find that the reduction in the TL of the wall is
3 dB . In other words, the overall TL of the wall-opening
assembly is (30 - 3) = 27 dB .

Table 5.2 Reduction in the TL of a Wall by an Opening

REDUCTION IN THE TRANSMISSION LOSS OF WALL BY AN OPENING (in dB)

Opening area as a percentage of total wall area

100% 50% 20% 10% 5% 2% 1% 0.5% 0.1% 0.01%

0 0 0 0 0 0 0 0 0 0 0
1 1 0.5 0 0 0 0 0 0 0 0

2 2 1.0 0.5 0 0 0 0 0 0 0
3 3 2.0 1.0 0.5 0 0 0 0 0 0
4 4 2.5 1.0 0.5 0.5 0 0 0 0 0
5 5 3.0 1.5 1.0 0.5 0 0 0 0 0
..

l 6
7
6
7
4.0
5.0
2.0
2.5
1.0
1.5
0.5
1.0
0
0.5
0
0
0
0
0
0
0
0
8 8 6.0 3.0 2.0 1.0 0.5 0 0 0 0
9 9 6.5 4.0 2.5 1.0 0.5 0.5 0 0 0
I

=i 10 10 7.5 4.5 3.0 2.0 1.0 0.5 0 0 0



..J
E-t
15 15 12.0 8.5 6.0 4.0 2.0 1.0 1.0 0 0
20 20 17.0 13.0 10.5 8.0 5.0 3.0 2.0 0.5 0
30 30 27.0 23.0 20.0 17.0 13.0 10.0 8.0 3.0 0.5
40 40 37.0 33.0 30.0 27.0 23.0 20.0 17.0 10.5 3.0
50 50 47.0 43.0 40.0 37.0 33.0 30.0 27.0 20.0 10.5
60 60 57.0 53.0 50.0 47.0 43.0 40.0 37.0 30.0 20.0
1 04 Chapter 5

If we raise the TL of the wall of Figure 5.17 to 40 dB, keeping


the opening's TL at 20 dB, we see from Table 5.2 that the reduction
in TL is 10.5 dB, giving an overall TL of the assembly as (40 - 10.5)
= 29.5 dB. If we increase the TL of the wall further to 50 dB, keeping

the opening's TL at 20 dB, the overall TL of the assembly becomes


30 dB. If we increase the TL of the wall to 60 dB or 70 dB, the
overall TL of the wall does not increase, but remains constant at 30
dB. These facts are collected in Figures 5.18(a) to (d).
This example illustrates that although 90% of the wall has a
fairly high TL, the overall TL of the entire assembly is closer to that
of the opening's TL. This is an extremely important result, which
may be stated as follows:

If a wall has an acoustically "weak" element, its overall TL


is closer to that of the weak element.


The overall TL does not change significantly, or not at all, if
we make the acoustically "strong" part of the wall stronger.

In general, the overall TL of an assembly follows the general


pattern shown in the graph of Figure 5.18(e), which has been derived
from Figures .18(a) to (d). Table 5.2 can also be used to determine
the overall transmission loss of an assembly with more than two
components, as illustrated in Example 5.4. For more accurate results,
however, the mathematical formulation of Appendix F should be
used.

11=201 ] I
[:] 11 1
TL = 20 = 20
B dB


Lw.u n. = 30dB Lw.u n. = 40dB Lw.u TL =50dB Wall TL =6 0dB

(a) OVERALL TL = 27 dB (") OVERALL TL = 29.5 dB (c) OVERALL TL = 30 dB (d) OVERALL TL = 30 dB

,-- 0v<'rsll n. of
1
aee<'m"ly
iOw
I--

20 L_ l________ ._ ---------------
___
n. of op<'nlne


20 30 40 50 60
TL of wall (dB)

(e)

5.18 Overall TL values of a wall assembly with different TL values of wall TL, but the same TL of opening.
Principles of Airborne Sound Insulation 105

Example 5.4 Overall TL/STC of a Wall Assembly

A wall measuring 10 m x 3 m includes a window measuring 2 m x 1 .5 m and a door measuring 1 .7 m x


2. 1 m, as shown in the following figure. If the TL of the window and the door are 20 dB and 25 dB
respectively and the wall's TL is 55 dB, determine the overall TL of the entire assembly.

Solution: Since Table 5.2 can be used only for a two-component assembly, we must first determine the
effect of the window, and subsequently that of the door. The difference between the wall's TL and the
window's TL = 35 dB . The window occupies 10% of the total wall area. Reading under the 10%
column in Table 5 .2, the reduction in TL is 25 dB (obtained by approximate interpolation). Thus, the
overall TL of the wall and the window is (55 - 25) = 30 dB.
Now combine this TL of 30 dB with the door's TL of 25 dB . The difference between the two TL
values is 5 dB. The area of door is 1 2% of the total wall area. Hence, the reduction at 1 2% of the
opening area is approximately 1 dB . Thus, the overall TL of the entire wall is (30 - 1 ) = 29 dB .

10 m

Window Door Wall


TL = 20 dB TL = 25 TL = 55 dB
dB

OVERALL TL OF A55EM BLY= 29 dB

5.10 SOUND LEAKS AND A sound leak, or simply a leak, is an unintentional gap in the panel.
Since sound will pass through a leak uninterrupted, the TL of a leak
FLANKING is zero. Even a small leak area can reduce the transmission loss of a
TRANSMISSION panel substantially, because of diffraction effects (see Section 3 .4. 1 ) .
For example, assume that a 1 m x 2 . 1 m (3.3 ft x 7 ft door) with a TL
of 30 dB , has a 1 0 mm e18 in.) gap at the bottom of the door. The
percentage area of this gap is approximately 0.5%, obtained by
dividing the area of the gap by the area of the door. How much will
this tiny gap reduce the TL of the door?
The answer is found from Table 5.2. Since the TL of the gap is
[Q zero, the difference between the TL of the door and the gap is 30 dB .
30 Reading under the opening area of 0.5% in Table 5 .2, we find that
....
0
0
the door's TL has been reduced by 8 dB , i.e., the effective TL of the
'3
door is (30 - 8) = 22 dB . If we increase the TL of the door to 40 dB ,
......-
.....
0
20 keeping the same gap size, will it improve the effective TL of the
_J
1-
V
0 door? Very little, because, from Table 5.2, the reduction in TL with
:>
.,i:i a 40 dB door is 1 7 dB , which gives the effective TL of the door as 23
u
0 10
dB . If the door 's TL is increased further to 50 dB, we find that the
.....
..... 20 30 40 50
UJ effective TL of the door remains constant at 23 dB, Figure 5 . 1 9 .
TL of door without leak (dB) In summary therefore, a sound insulating construction must be
as airtight as possible. The sealing of gaps and penetrations is more
important in the case of a panel with a high TL (or STC) because the
5.19 Effective TL of a door with a 0. 5% reduction (in TL) increases with increasing TL of the panel, as
gap at the bottom of the door. indicated by the example just given.
106 Chapter 5

If there are any unintentional gaps such as at the top, bottom and
sides of a barrier, they should be sealed with a sealant over the entire
perimeter. Electrical and telephone outlets, switches, wall fixtures,
etc. - in fact all penetrations through the panel - must be sealed.
Doors and windows must be gasketed. Doors should be provided
with a threshold to seal the space under the door. Masonry mortar
joints must be completely filled, and if possible masonry walls should
be painted. Painting is particularly important in the case of concrete
masonry because of the porous structure of the units.
The extent of required sealing and the resulting benefits are
shown in Figure 5.20. Note that in the case of a light frame wall, all
that is needed is one seal at each face, provided each seal is installed
properly. However, two seals on each side of the wall will provide
redundancy, which is often desirable. Practical aspects of sealing
gaps in walls, doors, and windows are discussed in Chapter 6.

I J

Unsealed One eeslln!! 11esd Two eeslln!! 11esde Four eeslln!! 11esde

5TC 29 5TC 49 5TC 53 5TC 53

5.20 Vertical section through a metal stud wall assembly showing the effect of sealing the peripheral gap on
the STC of a partition. Adapted from Reference 5. 9.

5.10.1 Acoustical Sealant

An acoustical sealant is any sealant that is nonhardening and non


drying - one that remains indefinitely adhesive and flexible. The
properties required of an acoustical sealant are far less stringent than
those of sealants used in building expansion joints, control joint.
and structural glazing. An acoustical sealant is usually not subjecte
to severe weathering, expansion or contraction, structural stress, etc.
A number of manufacturers market acoustical sealants either as
gunnable material, Figure 5.21, or as peel-and-stick preformed foams.
Gunnable acoustical sealants are preferable over preforme
foams because of their ability to fill gaps of all profiles . Preforme d
sealants, on the other hand, require a constant minimum pressure to
produce a good seal. Preformed sealants, however, have rove ,
effective in sealing electrical, telephone and similar outletsf 91 .
If a sealant is to be painted over, its compatibility with the paint
must be checked with the sealant manufacturer. If the acoustical
sealant is used in a fire-rated assembly, the sealant must be fire
resistive. Fire-resistive acoustical sealants are available.
rrinciples of Airborne Sound Insulation 1 07

5.21 Gunnable acoustical sealant. Samples supplied by


Pecora Corporation, Harleysville, Pennsylvania. Photo
by Madan Mehta.

5.10.2 The Importance of Reducing Flanking Transmission

Just as it is important to plug leaks in a panel, it is also important to


eliminate or reduce sound transmission by indirect routes . An indirect
route is called a flanking path. The sound traveling along a flanking
path is referred to as flanking transmission.
The concept of flanking transmission is illustrated in Figure 5 .22.
As shown in this figure, the panel separating Room 1 from Room 2
is the partition wall. In going from Room 1 to Room 2, the sound
will not only travel through the partition, but also along several other
routes. It is these other routes that are referred to as the flanking
paths. If the sound traveling along flanking paths is not sufficiently
attenuated, increasing the STC of the partition is of little or no
consequence. Means to reduce flanking transmission vary with the
type of construction. Some are discussed in Chapter 6.

Floor

ROOM 1 ROOM 2 ROOM 3

Floor -

5.22 A section through a building showing direct and flanking


's ound transmission paths.
108 Chapter 5

5.11 LABORATORY STC The STC data of various construction types, given in Appendix J,
has been obtained from laboratory measurements. In laboratory STC
VERSUS FIELD STC measurements, all leaks around the test specimen are carefully sealed
and flanking transmis sion virtually eliminated. In an actual
construction, however, the sealing is not as good as in the laboratory,
leading to the presence of a certain degree of flanking transmission.
Consequently, laboratory STC values are not usually realized in
the field. Measurements of STC of panels in the field (called the
field STC, abbreviated as FSTC) have been reported to be much lower
than the laboratory STC values. In some situations, FSTC may be
as much as 1 5 dB lower than the STC. However, with a good quality
control at the construction site to ensure that all leaks have been
sealed and flanking transmission reduced, FSTC is usually 5 to 8
FSTC "' STC - (5 to 8) points lower than STC. That is, if the STC of a panel is 50, the
FSTC will be between 42 and 45.
The FSTC of a panel is obtained in the same way as the STC
"' implies "approximately
except that in the case of FSTC, the TL-frequency data of the panel
equal to".
is obtained from field mea surements , instead of laboratory
measurements, as is the case for STC. The FSTC of a panel is
5
measured according to the ASTM E 336 procedure l 10l .

5.12 CEILING SOUND Most commercial buildings have an uninterrupted plenum space
(space between the floor and the ceiling) to allow air conditioning
TRANSMISSION CLASS ducts and other building services to run through continuously.
Lightweight partitions simply terminate flush with the ceiling, or
penetrate the ceiling by a few centimeters, as shown in Figure 5.22.
In rare cases, such as the walls enclosing a toilet, a mechanical room,
or a room where a high degree of acoustical privacy is required, the
walls extend up to the floor or the roof above.
Where partitions do not extend up to the floor or the roof, the
transmission loss of the ceiling becomes an important factor in
providing sound insulation between adjacent rooms. As discussed
in Section 4.9, the ceiling in commercial buildings is generally made
of a porous material to provide needed sou n d abs orption.
Consequently, the ceiling 's TL is low.
To improve the ceiling 's TL, several manufacturers make ceiling
tiles with a back coating of an impervious laminate, such as an
aluminum foil bonded to the back of the tiles, to increase their
transmission loss. One manufacturer makes a ceiling tile consisting
of fiberglass bonded to a backing of a gypsum board. The impervious
layer improves the ceiling 's TL at the cost of a small reduction in its
absorption.
A single number index of ceiling transmission loss is called
ceiling sound transmission class (CSTC). The CSTC of a given
ceiling is measured by a two-room method in which one room is the
source room and the other, a receiver room. The measurements are
made at the same frequencies as the STC. However, unlike STC,
CSTC is expressed in groups, where the group width is 5 dB, such
as 24 to 29, 30 to 34, 35 to 39, etc. A ceiling whose CSTC is 36 is
classified as CSTC 35. A ceiling whose CSTC is 39 is also classified
as CSTC 35. Thus, CSTC 35 means that the ceiling 's STC lies
between 35 to 39. The higher the CSTC, the more sound insulating
the ceiling is.
A ceiling with a high NRC (0.95 to 1 .05) has a fairly low CSTC
(nearly 25). The maximum CSTC of commercially available
Principles of Airborne Sound Insulation 109

abso tive ceilings is nearly 45, with an NRC of about 0.6. A 1 3



m m ( /2 in.) thick gyQsum board gives a CSTC of 55 in the two
room test procedureT5 . 1 l l .

5.13 OUTDOOR-INDOOR Another single number index used i n comparing the T L of panels is
called the outdoor-indoor transmission class (OITC). The purpose
TRANSMISSION CLASS of OITC is to compare the airborne sound insulation of building
facades (including walls, doors, windows and their combinations)
with respect to outdoor noise. The outdoor noise spectrum used in
determining OITC is the average of three transportation noises -
aircraft, railroad and automobile.
The determination of OITCl51 21 of a component or assembly
follows a simple procedure and is similar to the calculation of dBA
levels from dB levels (see Example 2. 1). OITC is determined using
the same TL data as that used for STC, except that OITC includes
data that extends from 80 Hz to 4 kHz, while the data for STC extends
from 1 25 Hz to 4 kHz. Both STC and OITC use one-third frequency
band data. The greater the value of OITC, the more insulating the
assembly.
The reason for extending the data down to 80 Hz in the case of
OITC is that the transportation noise is rich in the low frequency
range. Therefore, OITC is recognized as a better single number
rating of building envelope elements than STC. STC is a good
index for speech, radio, television, and similar sources of noise,
commonly obtained in residential and commercial buildings. For
other noises, such as machinery, manufacturing processes, bowling
alleys, music, etc., the detailed (one-third octave) TL data should be
examined.
1 10 Chapter 5

STC VALUE FROM TL DATA - The Graphical P rocedure

This section illustrates the procedure of determining the STC of a panel from its measured TL data with
the help of the following example.
Assume that the measured TL values of a panel at sixteen one-third octave bands are as follows:

Frequency (Hz) TL (dB) Frequency (Hz) TL (dB)

1 25 22 800 28
1 60 25 1 ,000 33
200 26 1 ,250 38
250 28 1 ,600 42
315 30 2,000 45
400 32 2,500 48
500 34 3,200 51
630 35 4,000 54

To determine the STC of the panel, we first plot the above frequency-TL relationship to a suitable
scale as shown in Figure A (facing page). We next draw the STC contour on a separate transparent
sheet to the same scale as the frequency-TL relationship, and overlay it on the above relationship.
The STC contour is now moved as far up as possible until it satisfies the two conditions described
in Section 5.8, which are reproduced here for convenience.
The sum of deficiencies of TL at 16 frequencies should not exceed 32 dB, and

The maximum deficiency at any point should not exceed 8 dB.
The location of STC contour shown in Figure A satisfies these two conditions. We note that the
deficiencies occur at the frequencies shown in the following table. A deficiency is defined as any point
on the TL-frequency relationship that is below the STC contour. Any point above the contour is not a
deficiency, and therefore a value of zero has been entered in the following table against those frequencies.

Frequency (Hz) Deficiency (dB) Frequency (Hz) Deficiency (dB)

1 25 0 800 8
1 60 0 1 ,000 4
200 0 1 ,250 0 Sum of deficiencies = 1 3 dB
250 0 1 ,600 0
315 0 2,000 0
400 1 2,500 0
500 0 3,200 0
630 0 4,000 0

Reading the STC value at 500 Hz, we observe that it is 34. Hence, this panel is rated as STC
34.
The location of the STC contour in this example has been governed by an 8 dB deficiency
at 800 Hz - not by the sum of deficiencies - since the sum ( 1 3 dB) is well below the maximum
permissible value of 32 dB . In a TL-frequency relationship that does not have a major dip, the
sum of deficiencies governs the location of the STC contour, as shown in Figure B . Note that
the average TL of the panels of Figures A and B are nearly the same, yet the two panels have
different STC values.
Principles of Airborne Sound Insulation 111

:30

3
IS)

]
c:
c
FIGURE A ii\ 20
IS)
The placement of the 5TC contour .E
i:ly 8 dB
IS)
hae Peen eovenned c:
(IS
deficiency at a frequency (000 ....
1-
Hz).
10 L-----

125 200 :315 500 000 1.250 2,000 :3,150


160 250 400 630 1,000 1,600 2,500 4,000

Frequency (Hz)

30

3
IS)
IS)
.2
c:
FIGURE B
c
'ii\ 20
IS)
The placement of the 5TC contour .E
hae Peen eovenned i:ly the eum of IS)


c:
deflclenclee at all frequenclee.
1-

125 200 :315 500 000 1,250 2,000 :3,150


160 250 400 6:30 1,000 1,600 2,500 4,000

Frequency (Hz)
112 Chapter 5

STC VALUE FROM TL DATA -The Numerical Procedure

The numerical procedure for determining the STC is easier, and is amenable to computerization.
However, unlike the graphical procedure, it does not give a clear insight into what is being achieved.
The procedure consists of the following steps and is explained with reference to the same data that was
used for the graphical procedure.

Step 1 : Record the TL at each one-third octave frequency in Column 2.


Step 2: Write the adjustment factor against each frequency in column 3. These factors are based
on the STC contour, assuming the adjustment as zero at 500 Hz.
Step 3: Add the TL and adjustment factors in column 4. This is the adjusted TL (TLa ) .
dj
Step 4: Note the least TLadj value and circle it, as shown i n column 4. In this example, the least
TL is 26.
as.lj
Step 5: Add eight to the least TLad This is the first trial STC (STCtrial). In this example the first
j
STCtrial is 34. Write this value in Column 5, as shown.
Step 6: Subtract STC ttial from TLad i.e., determine (TL STCtrial). If this value is positive,
j adi
-

enter zero in column 5; if negative enter the actual value. Th1s value is called the deficiency.
Step 7: Add all the deficiencies. If the sum of deficiencies is less than or equal to 32, stop here
and our STC trial is the STC of the component. If the sum is greater than 32, repeat step 6
by choosing TCtrial as one below the previously chosen value.

In this example, the sum of deficiencies is 13, which is less than 32. Therefore, the STC of this component
is 34. However, if the sum of deficiencies was greater than 32, we would have then chosen STCtrial =
33, and repeated step 6, until the sum of deficiencies was less than or equal to 32.

Frequency (Hz) TL (dB) Adjustment Adjusted TL Trial STCs and Deficiencies


factor (Tladj ) ( STCtrial 3=
4)

1 2 3 4 5

1 25 22 16 38 0
160 25 13 38 0
200 26 10 36 0
250 28 7 35 0
315 30 4 34 0
400 32 1 33 -1
500 34 0 34 0
630 35 -1 34 0
800 28 -2 @ -8
1,000 33 -3 30 -4
1,250 38 -4 34 0
1,600 42 -4 38 0
2,000 45 -4 41 0
2,500 48 -4 44 0
3,200 51 -4 47 0
4,000 54 -4 50 0

Sum of deficiencies 13
Jrinciples of Airborne Sound Insulation 113

The graphical equivalence of this procedure i s to flatten the STC contour into a horizontal line using the
500 Hz point on the contour as the hinge point. For example, since the 1 25 Hz point is 1 6 dB below the
500 Hz point, the adjustment factor for 1 25 Hz is + 1 6 dB . Similarly, the adjustment factor for 1 ,000 Hz
is 3 dB because, the 1 .000 Hz is 3 dB above the 500 Hz point, see Figure 5 . 1 6. The flattening of the
-

STC contour is to be accompanied by corresponding adjustment of the TL curve. Thus, the TL value at
1 25 Hz is increased by 1 6 dB, and that for 1 .000 Hz is decreased by 3 dB Once the STC contour is a
horizontal line, it becomes easier to move the (horizontalized) contour vertically until it satisfies both
conditions, repeated below for convenience.

The sum of deficiencies of TL at 1 6 frequencies should not exceed 32 dB, and

The maximum deficiency at any point should not exceed 8 dB.

5.1 Egan, D. M.: Architectural Acoustics, McGraw Hill Inc., New York,
REFERENCES
1988, p. 1 80.
5.2 Beranek, L. L.: Noise Reduction, Robert E. Krieger Publishing Company,
Huntington, New York, 1980, p. 287.
5.3 Faulkner, L. L.: Handbook of Industrial Noise Control, Industrial Press
Inc., New York, 1976, p. 173.
5.4 Harris, D. A.: Noise Control Manual, Van Nostrand Reinhold, New York,
1991, p. 40.
5.5 Wilson, C. E.: Noise Control -- Measurement, Analysis, and Control of
Sound and Vibration, Harper & Row, New York, 1989, p. 249.
5.6 Flat Glass Marketing Association: Glazing Manual, Flat Glass Marketing
Association, Topeka, Kansas, 1990, p. 36.
5.7 American Society for Testing and Materials, Philadelphia: "Standard Test
Method for Laboratory Measurement of Airborne Sound Transmission
Loss of Building Partitions", ASTM Standard E 90-90.
5.8 American Society for Testing and Materials, Philadelphia: "Classification
for Rating Sound Insulation", ASTM Standard E 413-87(94).
5.9 American Society for Testing and Materials, Philadelphia: "Standard
Practice for Use of Sealants in Acoustical Applications", ASTM Standard
E 91 9-84(92)
5.10 American Society for Testing and Materials, Philadelphia: "Standard Test
Method for Measurement of Airborne Sound Insulation in Buildings",
ASTM Standard E 336-90.
5.11 Egan, D. M.: Architectural Acoustics, McGraw Hill Inc., 1988, p. 2 1 8.
5.12 American Society for Testing and Materials, Philadelphia: "Standard
Classification for Determining Outdoor-Indoor Transmission Class",
ASTM Standard E 1 332-90.
Airborne Sound
Insulation
Practice

This chapter is a continuation of the previous chapter and deals with


airborne sound transmission loss characteristics of commonly used
assemblies, such as concrete and masonry components, light-frame
construction in wood and steel, and doors and windows. Construction
details and practical issues that affect airborne sound insulation, in
terms of the STC values, are discussed.
It is important to note that good detailing is fundamental to
achieving a high STC. For example, it is futile to use a door or a
window with a high STC if its perimeter is not fully sealed. Similarly,
a partition with a high STC is superfluous if flanking paths are not
minimized. On-site quality control of construction is equally
important since the STC value is highly sensitive to lapses by
construction workers.
Before proceeding to discuss the above issues, it is important to
understand the subjective perception of an STC value. In other words,
what does STC 20, or for that matter, STC 30 mean? Will an STC
30 partition block all noise from one space to the other? Table 6.1
attempts to answer these questions and relates an STC value with
the subjective perception of airborne sound insulation.

115
116 Chapter6

Table 6.1 Subjective Perception of STC Values*

5TC FSTC 5ul:ljectlve deecrlptlon

This is just
This is just
wonderful
wonderful

~
30 22-25 Moet sentences clearly understood.

~
This is just
Thi .....

wonderful
wonderful
Speech can l:le heard with some effort.
40 32-35 Individual worde and occasional

~
phraeee heard.

That is ' Tha....


'
absolutely crazy absolute.. crazy

Loud epeech can l:le heard with some

~
50 42-45 effort. Muelc eaelly heard.

That is '

absolutely crazy

~
Loud speech essentially lnaudll:lle.
60 52-55
Mueic heard faintly; !:lass note
dleturl:llng.

n
'
Loud mueic heard faintly, which could


70 62-65 l:le a prol:llem If the adjoining apace le

~
highly eeneitlve to eound intruelon,
such ae a recording etudlo, concert
:
hall, etc.

~
75 Most nolees effectively l:llocked.
and
a!:love
'

* This table assumes a reasonably quiet background noise level in the receiving room - NC 35 or less. See Chapter 8 for
NC values.
Airborne Sound Insulation Practice 117

6.1 CONCRETE WALLS As indicated in Table 5.1, the coincidence dip of a 200 mm (8 in.)
thick concrete panel occurs around 100 Hz; for a 400 mm (16 in.)
AND SLA BS thick panel, the dip is around 50 Hz. For a 100 mm (4 in.) thick
panel, the dip occurs at nearly 200 Hz (see Section 5.5). Since, the
thickness of a commonly used concrete wall or slab is greater than
150 mm (6 in.), its coincidence dip usually falls below the frequency
region of interest to us, which in most situations is 125 Hz to 4kHz.
Consequently, the coincidence dip does not influence the general
shape of TL-frequency relationship of a concrete wall/slab,
particularly that of a thick one. In other words, the TL of a concrete
wall/slab increases almost uniformly within 125 Hz to 4kHz region,
as suggested by the mass law.
However, due to its stiffness, the TL values of a concrete wall/
slab are lower thari those given by the mass law. Additionally, the
TL of a concrete wall/slab increases at the rat of nearly 5 dB per
octave, instead of the mass law's prediction of 6 dB per octave. These
observations are illustrated in Figure 6.1, which gives the TL
frequency relationships of a 200 mm (8 in.) and a 150 mm (8 in.)
thick concrete wall/slab.
Although measured TL-frequency data should be obtained for
use in any critical sound insulation application, the following
empirical relationship may be used to give a rough estimate of the
TL of a concrete wall/slab. Thus, if f is the frequency of sound in
Hz, and m is the surface mass of wall/slab in kg/m2 (lb/ft2), then its
approximate TL is given by:

TL
(dB)

-+--
5 dB/octave fit to !
meaeured data
Meaeure TL-frequency
Meaeured TL-frequency
t---+---:>-<tr--+- relatlonehlp for a 200 mm
h.S+=+ relatlonehlp for a 150 mm
(6 In.) thick concrete panel
(8 In.) thick concrete panel
I
40 L__L--_J__L_
125 250 500 1,000 2,000 4,000 250 500 1,000 2,000 4,000
Frequency Frequency
(a) 200 mm (8 In.) thick concrete panel: (b) 150 mm (6 In.) thick concrete panel:
weiht 475 /m2 (97 lb/ft;2) welht 350 /m2 (71 lb/ft;2)
STC58 STC56

6.1 TL-frequency relationships of (a) 200 mm (8 in.) thick concrete panel, and (b) 150 mm (6 in. ) thick concrete
wall or slab.
1 18 Chapter 6

TL = 16.5 (f m) - 35, if m is in kg/m2


(6. 1)
TL = 16.5 (f m) 24, if m is in lb/ft2
-

Equation (6.1) should be used for a normal-weight concrete walV


slab, which is at least 150 mm (6 in.) thick. For a wall/slab less than
150 mm thick, the coincidence frequency lies in 125 Hz to 4kHz
region, and therefore the actual TL values are less than those obtained
from the above equation. For a 100 mm (4 in.) thick wall/slab, a 2
dB reduction from the values of Equation (6.1) may be used.
From Equation (6.1), we can obtain the following empirical
relationship for the STC of a concrete wall/slab - at least 150 mm
thick.

STC = 16.5 log (m)+ 14, if m is in kgtm2


(6.2)
STC = 16.5 log (m)+ 25, if m is in lb/ft2

6.2 SINGLE-LEA F Although the TL-frequency relationship of a masonry wall followS!


the general principles of the mass law, it cannot be estimated by 31
MASONRY WALLS simple relationship similar to Equation (6.1). The reason is that unlike
concrete, which is relatively monolithic, masonry is a composite of
masonry units and mortar. Being of lower strength than the units,
the mortar reduces the stiffness of masonry, thereby affecting the TL
of the wall. The presence of air pockets - core holes in clay masonry
(bricks) and cells in concrete masonry units (concrete blocks)
also influences the wall's TL.
The most important factor that distinguishes a masonry wall from
a monolithic concrete wall/slab is the air leakage through masonry
mortar joints. Usually the horizontal mortar (bed) joints are more
BLOCK airtight than the vertical (head) joints because of the pressure exerted
by the weight of the wall on bed joints. Similar pressure is not
available on head joints, unless the masons consistently press each
head joint to ensure the wall's airtightness.
Obviously, the air leakage reduces the TL of a masonry wall.
This is particularly true of a block wall, since most blocks are made
of a lightweight, and hence a porous, concrete. Additionally, block
walls are usually face-shell-mortared, implying that the mortar join
are only as deep as the thickness of face shells - typically 35 to 40
1
mm (nearly 1 / in.). If the workmanship is substandard, a block
2
wall's airtightness is seriously compromised.
Mortar joints in brick walls, on the other hand, extend the full
depth of units. Therefore a brick wall, of a comparable thickness
and weight, is more airtight and hence, more insulating than a block
MORTARING OF A wall.
BRICK WALL
Figure 6.2 shows the TL-frequency relationships of brick walls
of three thicknesses - 100 mm (4 in.) nominal, 200 mm (8 in.)
Airborne Sound Insulation Practice 119

nominal and 300 mm (12 in.) nominal. Notice, once again, the virtual
absence of coincidence dips, and an almost uniform increase of TL
with frequency and thickness.
It is because of the absence of a coincidence dip within the
frequency range of interest that a thick concrete or masonry wall is
an excellent sound insulating barrier, and is recommended for use
where its weight is not a deterrent.

70 300 mm (12 In.) nominal brick wall,


570 ketm (117 lb/ft2)
2
r---,---,----.--r-----.--,---,

200 mm (8 In.) nominal brick wall,


-+--+--+--+-!--+-*""'-+----'---,;> 405 ketm2 (83 lbtft2)

60 100 mm (4 In.) nominal brick wall,


-+-----+--+-r'-f----+-A---+--+-r'-l
190 ke/m2 (39 lbtft2)

TL
(dB) 50 f----+--+-+--i'-+----A--+c--+
I
I
I

5TC52
I
-
i
!
--
I
!'

30 '----'---'----"--'---'--"
125 250 500 1,000 2,000 4,000
Frequency (Hz)

6.2 TL-frequency relationship of a 100 mm (4 in.) thick brick wall, a 200 mm (8 in.) thick brick
wall, and a 300 mm (12 in.) thick brick wall. Adapted from Reference 6.1.
Note: The actual thickness of a masonry wall is 10 mm (3/8 in.) less than its nominal thickness.

6.2.1 Estimating the STC of a Brick Wall

The TL-frequency relationship of a thick multi-wythe brick wall,


such as a three wythe thick - 300 mm (12 in.) nominal or thicker
- is similar to that of a concrete wall. Bricks are usually dense and
since the mortar joints in each wythe are staggered, a thick brick
A ONE-WYTHE BRICK WALL
wall is reasonably airtight. For such a wall, Equation (6.2) may be
used to estimate of the wall's STC.
A one or two wythe brick wall is less airtight. Therefore, its
STC value is lower than that predicted by Equation 6.2. As a rough
guide, for a two-wythe wall, STC given by Equation (6.2) should be
reduced by nearly 4 points; for a single-wythe wall, a 6 point
reduction should be used.

For example, assume that a 100 mm (4 in nominal brick wall
has a surface weight (m) of 170 kg/m2 (35 lb/ft ). When the value of
m is substituted in Equation (6.2), we get the calculated STC as 51,
A lWO-WYTHE BRICK WALL which when reduced by 6 points gives us the actual STC of 45.
120 Chapter 6

6.2.2 Estimating the STC of a Concrete Masonry Wall

Based on several tests, the National Concrete Masonry Association


(NCMA) has developed the relationship of Figure 6.3 to estimate
the STC of a block wall. The relationship holds only if the wall is
either painted, plastered, or otherwise sealed, so that the sound
transmission through it is due to its vibrations, rather than the passage
of air through the wall1. Additionally, a concrete block wall must be
at least 100 mm (4 in.) thick for the relationship of Figure 6.3 to
apply.

54 .......,
5TC 16.75 (m)0.2, If m 15 in ke/m2
/
=

50 5TC 23 (m)0.2, If m 15 In 11:1/ft2


/
=

STC
46 /
/
42 V
100 150 200 250 300 350 400 (ke/m2)
20 30 40 50 60 70 80 (11:1/ft2)
Surface ma55

6.3 Approximate STC value of a (sealed) concrete masonry wall. Adapted from Reference 6.2.

6.3 MASONRY CAVITY As stated in Section 5.6, if the cavity depth is small, the TL of a
cavity wall at low frequencies may be lower than that of a single
WALLS leaf wall of the same surface weight due to cavity resonance.
However, the TL values of most commonly used masonry cavity
walls are higher than those of equal-weight single-leaf walls. The
increase in sound insulation resulting from the air space is a function
Fiberglass or similar porous material
in cavity
of the extent of decoupling between leaves.
Maximum insulation is obtained when the two leaves are
structurally independent of each other with no connection between
them, as shown in Figure 6.4. The use of a porous material in the
cavity and increasing the width of cavity space increases decoupling
between leaves (see Section 5.6).
The two leaves of a (commonly used) conventional masonry
cavity wall are, however, not independent. They share a common
support at the top and the bottom, and are connected together by
metal ties for structural reasons. Therefore, the sound insulation
provided by a conventional cavity wall is much lower than the wall
of Figure 6.4.
6 .4 A cavity w a l l with two 1 To minimize porosity leaks in a concrete masonry wall, the wall may be plastered,
structurally independent leaves. parged, or sealed with a high solids block filler, or two coats of heavy enamel.
Airborne Sound Insulation Practice 121

However, improvement in insulation can be obtained by


increasing the width of cavity. Apart from increasing the coupling
between leaves, a small cavity is prone to bridging of the two leaves
caused by squeezed mortar joints. Ideally, the cavity space should
be at least 100 mm (4 in.). The conventional cavity wall with a 50
mm (2 in.) cavity provides only a marginal improvement over a
single-leaf wall of the same weight.
It is also important to ensure that the cavity space is not clogged
by mortar droppings due to poor workmanship, since a mortar
clogged cavity bridges the two leaves. Specifying that a porous
absorber, such as a medium density fiberglass, be placed in the cavity,
not only adds to the damping of the cavity, but also ensures that the
masons will be careful to keep the cavity clean.
Additional decoupling of leaves in a conventional cavity wall is
provided by using a vibration-damping material, such as a high
density fiberglass board, under the first course of one of the leaves
of the wall, Figure 6.5. A gap should be left between the top of the
wall and the structural frame. The gap should be finished with backer
rod and an acoustical sealant on both sides. The lateral stability to
the wall is provided by embedded steel angle, spaced nearly 1.2 m
(4ft) apart. Using two-part ties, Figure 6.6, in place of one-part ties
further enhances decoupling of leaves.

A TWO-PART TIE FOR A


BRICK CAVITY WALL

High denelty fll1erglae;e;


.
. ,i. :v
. .
.
... .. . .
.

.. .. . Eml1edded etllel angle


<;)<:>.o'. F111erglae;e; lnflll
:

.

c;=tl-tt:f;- Acouetlcal eealant


and 11acker rod
F111erglae;e; lnflll

Mlltal tiee

6.5 A conventional masonry cavity wall detail.

6.6 Two-part metal ties for clay masonry and


concrete masonry walls.
122 Chapter 6

6.4 FURRED MASONRY A gypsum board lining to a masonry wall, either with metal furring
sections or wood furring sections (such as 50 mm x 50 mm, i.e., 2 x
WALLS 2 nominal) improves the sound insulation provided by a masonry
wall. Further increase in insulation is obtained by using fiberglass
in the cavity and resilient channels (see Figure 6.9 for an illustration
of a resilient channel).
For example, a painted 200 mm (8 in.) nominal concrete masonry
wall gives an STC of nearly 48, Figure 6.7(b). If gypsum board is
fastened to resilient channels, which in turn are fastened to 50 mm x
50 mm (2 x 2) nominal wood furring sections, and the cavity space
is filled with fiberglass, the STC of the assembly is nearly 56 - an
increase of 8 points, Figure 6.7(e).
If the gypsum board lining and fiberglass filling in cavity is used
on both sides of the wall, the STC of the assembly is nearly 59,
Figure 6.7(f). Note that it is adequate to use the resilient channel on
one side of the assembly. Using it on both sides provides an
insignificant increase in STC.
If the gypsum board lining is fastened to a metal or wood stud
wall, which is not attached to the masonry wall (i.e., a structurally
independent stud wall), the resulting assembly gives a high STC.
Figure 6.7 (g). Two independent gypsum stud walls, one on each
side of a masonry wall, yield an assembly with a STC of more than
70, Figure 6.7(h).
In fact, gypsum board lining of masonry walls with furring
sections is a commonly used method to improve the TL of an existing
masonry wall.
Airborne Sound Insulation Practice 123

APPROX. 5TC
44
1 11 1 t " " '''""" 200 mm c ,,_) m'' " ""'m.,..
" ''"h-ht ""

(a)

52
]D D[C ;::;:: :: (c)
, .rrr,htw<r,ht ,oo-. m""' 11

16 mm (5/8 In.) gypeum j,oard

]tiD1E
Rt:elllt:nt channt:le
53 Hat et:ctlone
Palnttld 200 mm (8 In.) nominal llghtwt:lght concrt:ttl maeonry wall
(d) SEE FIGURE 6.9 FOR
AN ILLUSTRATION OF
16 mm (5/8 In.) gypeum j,oard RESILIENT CHANNEL
Rt:elllt:nt chanMie
56 50 x 50 (2 x 2) nominal wood furring et:ctlone
Palnttlt.:l 200 mm (8 In.) nominal llghtwt:lght concrtlttl maeonry wall

(") SEE APPENDIX J FOR


DETAILED TL DATA OF
16 mm (5/8 In.) gypeum j,oard THESE AND OTHER
. . ... . ' . .
. ..
Rt:elllt:nt channt:le ASSEMBLIES

Joo:c
T-- '- - T-. -:
50 x 50 (2 x 2) nominal wood furring et:ctlone
59 '37 mm (11/2 In.) thick flj,t:rglaee
Palnttld 200 mm (8 In.) nominal llghtwt:lght concrtlttl maeonry wall
,... ::.. .0:-: .. .....
'37 mm (11/2 In.) thick flj,t:rglaee
..:.. ........ . .....:.: ... .-:: ...(. :.... .:: 50 x 50 (2 x 2) nominal wood furring et:ctlone
(f) 16 mm (5/8 In.) gypeum j,oard

. .. :t.. .:.. ...:- ._


.> :.:r. 16 mm (5/8 In.) gypeum j,oard

INCREASING
STC
65
J':oo (e)
c
75 mm ('3 In.) dt:t:p lndt:pt:ndt:nt mt:tal etud wall
6'3 mm (21/2 In.) thick flj,t:rglaee
Palnttld 200 mm (8 In.) nominal llghtwt:lght concrtlttl maeonry wall

16 mm (5/8 In.) gypeum j,oard


75 mm ('3 In.) dt:t:p lndt:pt:ndt:nt mt:tal etud wall
6'3 mm (21/2 In.) thick flj,t:rglaee
72 Palnttld 200 mm (8 In.) nominal llghtwt:lght concrtlttl maeonry wall
6'3 mm (21/2 In.) thick flj,t:rglaee
75 mm ('3 In.) dt:t:p lndt:pt:ndt:nt mtltal etud wall
16 mm (5/8 In.) gypeum j,oard.

6.7 Increase in sound insulation provided by various ways of furring a masonry wall.
124 Chapter 6

6.5 LIGHTWEIGHT In principle, a stud wall is similar to a masonry cavity wall, in


which the two gypsum board leaves are separated by an air space.
GYPSU M BOARD The leaves are tied together by studs, whose function is similar
ASSEMBLIES to that of metal ties in a masonry cavity wall. The TL-frequency
relationship of a stud wall is governed by the principles discussed
in Section 5.6, and it is suggested that the reader review this
section at this juncture.
Figure 6.8 shows the TL-frequency characteristics of the
following stud wall assemblies:
A wall with a layer of gypsum board on each side of
studs.
A wall with a layer of gypsum board on each side of
METAL STUD WALL
studs with fiberglass in cavity space.



A wall with a two-layer gypsum board on each side of
. . .
studs.
Observe that in all the three relationships there is a large
dip at nearly 2 or 2.5 kHz frequency. This is the coincidence
WOOD STUD WALL dip. In fact, the coincidence effect in all gypsum board
assemblies occurs in the 2 to 2.5 kHz region (see Table 5.1).
The cavity resonance dip usually occurs around 100 to 200 Hz
region Hz (see footnote 6 in Chapter 5). Therefore, in most
gypsum board assemblies, the cavity resonance dip is not seen
in 125 Hz to 4 kHz region.

(a) Two lay<'re of 16 mm (5/8 In.) eypeum


board, 63 mm (21/2 In.) ete<'l etude, and two
lay<'re of 16 mm (5/8 In.) eypeum board.
53 kg/m2 (10.8 lb/ft2)

TL
(dB)
50 \
:
: :!
j/ : (b) 16 mm (5/8 In.) gypsum board, 63
mm (21/2 In.) ete<'l etude, flb<'rglaee, and
, ,
..
-+-J--_,1,._ ._'-+!--/.'+----:t ),__ 1


16 mm (5/8 In.) gypeum board.
31 ketm2 (6.3 lb/ft2)
i225E
i ' . . /
-+--\-l--+--"""-1 (c) 161 mm (5/8 In.) eype;um board, 63
J
mm (2 /2 In.) ete<'l etude, and 16 mm (5/8
STC 49 In.) gypeum board.
l----i-----l+---,f+--"'-..+/'--J'--_,_---+_+--+---+-+--+"""""If--1 27 lcg/m2 (5.5 lbtft2)
(c;)
.] :
. .. . . .
STC '37

20 ----__,___-+--
125 250 500 1,000 2,000 4,000

Freliuency (Hz)

6.8 TL-frequency relationships of gypsum board and steel stud assemblies.


Airborne Sound Insulation Practice 125

6.5.1 Use of Porous Absorber and Resilient Channels

From Figure 6.8, observe the marked improvement in TL due to the


use of fiberglass (or similar porous absorber) in the cavity. In fact,
the increase is almost of the same order as that obtained by doubling
the layers of gypsum board. Further increase in TL of a gypsum
board assembly is obtained through the use of a resilient channel.
A resilient channel is a means of decoupling the gypsum board
from the studs. It consists of a lightweight steel section, whose profile
is shown in Figure 6.9. Resilient channels are fastened to studs
(typically) at 600 mm (24 in.) on center, and the gypsum board is
fastened to the resilient channels, Figure 6.10.
The free end of the channel must be oriented toward the top so
that when the gypsum board is fastened, the free end is pulled away
from the stud by a small amount. Thus, it will be incorrect to fasten
the channel with its free end at the bottom. It is also important to use
6.9 Resilient channel. the correct size of screws. Longer screws will increase the mechanical
connection (coupling) between the gypsum board and the studs, if
the screws bridge to the stud.
A resilient channel is effective only if a porous absorber, such as
fiberglass or mineral wool, is used in the spaces between studs.
Without the porous absorber, the improvement in sound insulation
obtained through the use of a resilient channel is marginal.

Free end of
reelllent channel

Reelllent channel
(approx. one-third
full elze)

Note the emall elu


of ecrewe here.

6.10 (Part) vertical section through a gypsum


Gypeum Poroue abeorber
board wall assembly with resilient channels. board In etud cavity
126 Chapter6

The perimeter of a resiliently mounted gypsum board wall must


not make contact with the floor, ceiling or the adj acent wall. A small
gap should be left between the wall and the above surfaces, which
should be caulked with an acoustical sealant, Figure 6.11. If two
layers of gypsum board are used on the same face of the wall, the
joints between the layers should be staggered.
The use of a resilient channel on both si9es of an assembly is
indeed better than on one side only, but the improvement so gained
is marginal, Table 6.2. Therefore, the resilient channel is usually
used on one side only. Resilient channel is equally effective with a
metal stud or a wood stud assembly.
With metal studs, which are typically less stiff than wood studs,
it is important to specify 20 gauge or heavier studs.

Porou& a11&orl1er &uch


as fl11erela&&

Re&lllent channel Caulk

=======* . . . . . . . . . .
Caulk

Reelllent
channel
DETAIL Q
Firet layer of eYp&um
11oar.:l 5econ.:l layer of
RESIL IENTLY SUPPORTED SURFACE SHOULD
eypeum 11oard
NOT CONTACT A DJACENT SURFACE EXCEPT
THROUGH CAULKING

W I DTH OF A CAULKED JOI NT: 6 mm (1/4 In.)


Poroue al1eorl1er euch
a& fll1erela&&

Two-layer eYp&um 11oard Reelllent


with &taeeered Joint& channel

Note from Fleure 5.20 that only one 11ead of &ealant


F111erl1oar.:l le reulred alone the wall'e perimeter. Thle 11ead le
. : .: Caulk 11eet placed In the outer layer In a multi-layer
/
DETAIL p eYp&um 11oard a&&em111y, where lte pre&ence le
ea&lly verified durlne lnepectlon.

6.11 Typical details of a gypsum board wall assembly with resilient channels.
Note: Tests by U.S. Gypsum indicate improved STC when multiple layers of gypsum board
are laminated together with a rubber-like acoustical sealant, with base layer screwed to studs.
The damping provided by the sealant layer (referred to as constrained-layer damping) provides
additional energy loss in transmission.
Airborne Sound Insulation Practice 127

Table 6.2 Approximate STC Values of Wood Frame Walls with 13 mm (1/2 in.) Thick
Gypsum Board

Without cavity absorption With cavity absorption

Layers of gypsum board Layers of gypsum board

Gypsum board support system 1+ 1 1+2 2+2 1+ 1 1+2 2+2

38 mm x 90 mm (2 x 4) wood studs 37 40 43 40 43 46

38 mm x 90 mm (2 x 4) wood studs, 40 45 49 50 53 57
resilient channel on one side

38 mm x 90 mm (2 x 4) wood studs, 41 46 51 49 53 58
resilient channel on both sides

Staggered 38 mm x 90 mm 41 48 52 50 54 58
(2 x 4) wood stud wall

Double stud wall with 25 mm 46 53 57 57 61 63


(1 in.) space between studs

6.5.2 Staggered Stud and Double Stud Wall Assemblies

An alternative to the use of resilient channels is a staggered stud


assembly. A staggered stud wall consists of studs that are alternately
displaced, but are fastened to common top and bottom plates. The
Staggered etu.:l wall STC value of a staggered stud wall is almost the same as that obtained
Approx. 5TC 55 from a wall with resilient channels, Table 6.2.
Where a high STC value is required, a double stud wall assembly
is recommended. A double stud wall assembly consists of two
independent walls, with separate top and bottom plates, and an air
space between walls. Together with fiberglass in the cavity, and
multiple layers of gypsum board, a double stud assembly can give
an STC value of greater than 60, Figure 6.12.
By contrast, the commonly used partition in a single family home,
with 38 mm x 90 mm (2 x 4) wood studs and a 13 mm (1/2 in.) thick
Double etu.:l wall aeeembly gypsum board on each side, gives an STC of only 37 if the studs are
Approx. 5TC 63 spaced 600 mm (24 in.) on center. If the studs are spaced 400 mm
(16 in.) on center, the STC of the same assembly is only 34.
6.12 Staggered stud and double stud
rzssemblies.

STAGGERED STUD WALL DOUBLE STUD WALL


128 Chapter6

6.5.3 Effect of Stud Size and Spacing

In general, a larger stud will give a slightly higher sound insulation,


particularly at low frequencies, because of the deeper cavity space
and the increased decoupling it produces between leaves. For the
same reason, closer spacing of studs will reduce insulation. Thus, as
far as possible, stud spacing should not be less than 600 mm (24 in)
on center, and the spacing of screws should be as recommended by
the manufacturer of assembly. More-than-necessary screws will
increase coupling, and hence decrease TL.

6.5.4 Floor-Ceiling Assemblies

Sound insulation principles that apply to lightweight walls also apply


to lightweight floor-ceiling assemblies. The use of resilient channels
and fiberglass in cavity spaces between floor joists improves sound
insulation of floor-ceiling assemblies. A tongue-and-groove subfloor
is better than a (regular) plywood subfloor. If plywood subfloor is
used, the edges must be blocked to provide airtightness. An additional
layer of nearly 6 mm (1/4 in.) thick plywood panels, laid over a
[ Plywood eulriloor structural subfloor with staggered joints, provides neces sary
redundancy against airtightness.
To achieve a higher ST C , gypsum concrete or lightweight
portland cement concrete, 37 to 50 mm (1 1 o 2 in.) thick, laid over
a plywood subfloor, is recommended. An :STC value of nearly 60 is
obtainable with a concrete topped floor-ceiling assembly and a
resiliently supported ceiling, Figure 6.13. Usually, the concrete
topping is laid over a compressed fiberglass board or a similar impact
Block joint between absorbing surface to improve the structure-borne sound insulation
eubfloor panele of the assembly (see Chapter 7).

Carpet and pad


Plywood eulriloor
Carpet and pad
Plywood eubfloor
Carpet and
Lightweight
concrete
ad

High denelty .:.ao... . :a:;:



flberglaee --e

Flberglaee or
Plywood
Floor jolet mineral wool
eubfloor

Gypeum board Flberglaee or


ceiling Reelllent channel mineral wool

Reelllent
channel .......____.,

-::.E
;
-

Approx. 5TC 42 Approx. 5TC 48 Two-layer


gypeum board
Approx. 5TC 60

6.13 Commonly used wood light-frame floor-ceiling assemblies.


Airborne Sound Insulation Practice 129

6.6 SOUND LEAKS AND As stated in Chapter 5, sound insulating construction is highly
sensitive to air leakage. Therefore, assemblies should be as airtight
THE CONTROL OF as possible. Joints between two overlapping layers should be
FLANKING PATHS staggered.
The entire back surface of the electrical, telephone and other
outlet boxes in a sound insulating wall should be sealed to render
them fully airtight. Preformed sealant tape is best for this purpose.
Outlets in the front and the back of a wall should preferably not
be placed in the same stud cavity, Figure 6.14(a). They should be
staggered, with at least one intervening stud between them -
preferably more - Figure 6.14(b). Stud cavities containing outlets
should be lined with fiberglass or mineral wool.
Similarly, cabinets should not be provided in a sound insulating
single-stud wall. Cabinets in a double-stud wall should be staggered
and backed by gypsum board on all concealed faces, Figure 6.14(c).

Seal the !:Jack of Seal the !:Jack of


outlet I:Jox outlet I:Jox

(a) (11)

Gypsum I:Joard
lining to cai:Jinet
(c)

6.14 Detailing of electrical outlets and cabinets in a lightweight wall. Adapted from Reference 6.3.
130 Chapter6

6.6.1 Flanking Paths in Commercial Interiors

The importance of reducing flanking transmission was discussed in


Chapter 5. If flanking paths are not carefully blocked, a severe
degradation of sound insulation is obvious. In commercial buildings,
the most common cause of flanking transmission is through the
plenum, where the partition extends only up to the ceiling, Figur
6.15. To realize the sound insulation of the partition, flankin g
transmission through the plenum must be controlled.

0
0

Roof or floor /
PLENUM

Sound al1eorl11ne
c:elflne

6.15 Flanking transmission through a


plenum.

The most effective way to control flanking transmission is to


extend the partition all the way to the underside of the roof or the
floor above, Figure 6.16. In most situations, however, this is
impractical, since an an uninterrupted plenum is required to house
various building service equipment.
In the case of an uninterrupted plenum, the two alternatives given
in Figure 6.17(a) and (b)- neither as effective as that of Figure
6.16- may be used. In both these alternatives, the partition should
extend a little (at least 150 mm, i.e., 6 in.) above the ceiling.


Provide ceiling tiles with a high CSTC value (see Section
5.12), i.e., tiles with nonporous backing such as aluminum
foil, gypsum board, etc. In addition, lay a minimum of 150
mm (6 in) thick medium density (15 kg/m2, i.e., 3 lb/ft2)
fiberglass or mineral fiberboard layer over the ceiling, Figure
6.17(a). This layer should extend at least 1.2 m (4 ft) on
each side of the partition.

Provide a loaded vinyl barrier (see Figure 5.12) over the
partition, Figure 6.17(b). The vinyl barrier and fiberglass/
mineral wool in ceiling board layer may be used in
conjunction for greater effectiveness.
If the plenum is used for return air, provide a sheet metal
"boot" or chimney lined internally with fiberglass above each
room. The chimney should extend at least 600 mm (24 in.)
above the ceiling, Figure 6.17(c).
Airborne Sound Insulation Practice 131

0
0

Roof or floor

Extend partition up
to the roof or floor

PLENUM

6.16 Control offlanking transmission through


a plenum by extending the partition up to the
(a)
fioor or roof.

, o
o_ :.

Roof or floor
PLENUM
Loaded vinyl
Flt.>erelsee or curtain
mineral wool
Flt.>erelsee or mlnt:rsl wool
1.2 m (4ft;) minimum PLENUM

Sound st1eort11ne Sound st1eort11ne


celllne ceiling

(a)

(11)

:or
--o

PLENUM
Return-sir t.>oot lined
Internally with
flt.>erelsee, minimum
600 mm (24 In.) hleh

(c)

6.17 Control offlanking transmission with an uninterr upted plenum.


132 Chapter 6

Sound transmission via the curtain wall is another most common


flanking path in a modem commercial building. To counteract this
type of flanking, it is important to synchronize the location of the
partition and vertical frame member of the curtain wall (mullion), s
that the end of the partition abuts against the mullion. The joi
between the mullion and the partition should be filled with mediUDI
density fiberglass and caulked on both sides, Figure 6.18. Mullion
void should be filled with fiberglass or a plastic foam.
Air conditioning ducts are other efficient sources of sound
transmission - functioning as sound tubes - between rooms. A
careful layout of ducts in sound sensitive spaces can reduce this source
of flanking (see Chapter 9).

Curtain wall glazing

Flberglaaa or plastic
foam In mullion cavity

Flberboard
Realllent channel
Gypaum board
Flberglaaa/mlneral wool --t-+'ff'"="" _.,..
_

Partition

6.18 Joint between a sound insulating partition and curtain wall.

6.6.2 Flanking Paths in Wood Frame Interiors

Figure 6.19(a) shows a poorly detailed party wall between two


dwelling units in a multi-family residential building. Although the
party wall has been provided with resilient channels and fiberglass
in stud cavities, its fully potential is not being realized since some
sound will flank through the floor and attic spaces, as shown. For
satisfactory results, floor and attic spaces must be blocked, as shown
in a suggested detail in Figure 6.19(b).
Where a high degree of sound insulation between two units is
required, the party wall should be a double stud wall, and detailed to
provide discontinuity in the floor, Figure 6.20, and preferably, in the
roof and the foundation.
For a comprehensive coverage of detailing a light frame building
for sound insulation, the reader is referred to ASTM Standard E
497[6.31.
Airborne Sound Insulation Practice 133

6.19(a) Section through a party wall poorly


I i
detailedfor the control offlanking transmission.
ATII C

Flanking through
attic

....... ..

ATII C

:\ Reellient channel
Gypeum uoard
Caulk
Flanking through Jolet epace


o:.
'-tO
:.!to:.
.
. .

6.20 A party wall between two dwelling units to


achieve a high STC .

cantileverd here
6.19(b) Party wall [of Figure 6.18(a)] detailed
for the control offlanking transmission.
134 Chapter6

6.7 SOUND INSULATING Windows with a high transmission loss (sound insulating windows)
are required in buildings situated in high outdoor noise environments.
WINDOWS Houses, apartments, hotels, and hospitals in a busy urban center or
near an airport are usually required to have sound insulating windows.
A sound insulating window is one whose transmission loss
performance is better than that of a conventional single glass window,
or a window with an insulating glass unit. A typical single or double
hung window with a 3 mm e18 in.) thick glass has an STC of nearly
g_I
26. If the 3 mm thick ass is replaced by an insulating glass unit
(19 mm or 25 mm, i.e., I or 1 in.) thick, the STC value is slightly
lower (see Section 5.6). Commercially available sound insulating
windows, on the other hand, provide an STC of up to 55.

6.7.1 Air Leakage Control

Window technology has undergone a major evolution in the past


couple of decades, particularly in the production of energy-efficient
windows. An important part of an energy-efficient window is the
control of air infiltration or exfiltration (air leakage) through the
window. The air leakage control not only increases the energy
efficiency of the window but also its sound insulation. A strong
inverse correlation between air leakage and the STC of a window is
obvious.
Therefore, where a high sound insulation value is needed, a fixed
window should be specified. An operable window of the same
construction as the fixed window lowers the STC by 3 to 5 points,
depending on the STC of the window. A fixed window with a high
STC value suffers a greater reduction if made operable. Sliding
windows or sliding glass doors are particularly difficult to seal
effectively.
Air leakage is not only a function of the window itself, but also
of its installation in the wall. Therefore, a sound insulating window
should be installed strictly according to the manufacturer's
recommendations.

6.7.2 Thermal Break

Another important aspect of a thermally efficient window is a thermal


break, which is a plastic connector (usually high-strength
polyurethane) between two parts of the aluminum frame. Thus, a
thermal break separates the aluminum window frame in two parts:
one exposed to the outside and the other exposed to the inside. Being
of much lower conductivity than alurninum, a thermal break retards
the flow of heat through the frame.
The acoustical benefit of a thermal break is minimal, if any. By
providing discontinuity between the parts of the frame, it may increase
the acoustical efficiency of a window, but there is no data to support
Thermal l:lreak the assumption. Besides, the thermal break material is too stiff to be
THERMALLY BROKEN ALUMINUM FRAME an effective "break", acoustically. Figure6.21 shows a cross-sectional
detail of a sound insulating aluminum window. Notice that the
thermal break separates the window into two separate windows, each
with its own frame and glass panes.
Airborne Sound Insulation Practice 135

Thermal ureak

6.21 A sound insulating window. Sample courtesy


ofWindow Technologies, Inc., Temecula, California.
Photo by Madan Mehta.

Table 6.3 Approximate STC 6.7.3 Glass Thickness and the Use of Laminated Glass
Values of Fixed Windows With
Single Glass
Another important aspect of a sound insulating window is the
thickness of glass. Greater glass thickness means a greater mass,
and hence a higher sound insulation. However, the coincidence dip
Glass thickness STC
limits the increase in ST C of a thicker glass to a much smaller value
mm in
than that predicted by the mass law.
1
For instance, the coincidence frequency of a 3 mm ( /8 in.) thick
3 1/8 29 glass occurs at 5 kHz (see Table 5 .1). This is outside the frequency
6 1/4 31 range within which ST C is determined 125 Hz to 4 kHz. For a 6
9 3/8 32
-

mm (1/4 in.) thick glass, the coincidence frequency is 2.5 kHz, and
13 1/2 33 1
for a 13 mm ( /2 in.) thick glass, it is 1.25 kHz, both within the STC
25 1 37 frequency range.
Table 6.3 gives the approximate STC values of fixed single glazed
Note: Increase the above values by 3 windows with different glass thickness. Note that a window with a
for laminated glass. 3 mm (1/8 in.) thick glass has an ST C of 29. According to the mass
law, a window with a 6 mm (1/4 in.) thick glass should have an STC
of 35, a window with a 13 mm et2 in.) thick glass should give an
STC of 41, and so on. However, the actual ST C value for a 6 mm
thick glass window is only 31, and that of a 13 mm (1/2 in.) thick
glass window, 33.
As stated in S ection 5.5 , a laminated glass dampens the
coincidence effect. In general, a laminated glass increases ST C by
3, as compared with a monolithic glass of the same thickness.
However, under extremely cold climates, the resilience of the plastic
interlayer in a laminated glass is adversely affected, reducing the
benefit of lamination.
136 Chapter 6

6.7.4 Depth of Air Space

Using two glass panes with an intervening air space increases the
STC of window. However, a small cavity (less than 25 mm, i.e., 1
in.) in a double glazed window can in fact give worse TL than a
single glazed window (see Section 5.6). This is due to the cavity
resonance dip occurring within the frequency range of interest.
By increasing the depth of air space and the thickness of glass,
the cavity resonance can be lowered to below the frequency range of
interest. Usually, adding a storm window over an existing single
glass window with 75 mm (3 in.) air space can increase STC by 10
to 15 points.
If an air space is provided between the two glass panes of a
fixed window, the panes must be removable for cleaning purposes,
or a fully sealed unit should be used.

Table 6.4 Approximate STC Values 6.7.5 Lining of Window Reveal With a Porous Absorber
of Fixed Windows With Two Glass
Panes With an Air Space An improvement in TL performance is obtained by lining the interior
perimeter of window frame (window reveal) with a porous absorber
- fiberglass or mineral wool (nearly 65 kg/m3, i.e. 4 lb/ft3 density).
Glass thickness Air space STC Usually, the porous absorber in the reveal is covered with a perforated
mm in. mm in. metal or vinyl sheet, Figure 6.22. The lining of the reveal with a
porous absorber increases the STC of window by nearly 3 points.
Table 6.4 gives the approximate STC values of commercially
3 1/8 50 2 42 available sound insulating windows with two fixed glass panes with
6 1/4 100 4 44 different depths of air space. Note that using two 13 mm (11.2 in.)
6 1/4 200 8 46 thick glass panes with a 200 mm (8 in.) deep air space, an STC of
9 3/8 200 8 50 nearly 52 is achieved. If the reveals of this window are lined with a
13 1/2 200 8 52 porous absorber, the approximate STC of the window will be 55.
This is almost the same value as that given by a 200 mm (8 in.) thick
Note: Increase above values by 3 for windows concrete wall (see Section 6.1). It must, however, be noted that a
with reveals lined with a porous absorber. 200 mm thick concrete wall is far superior than the above mentioned
window at low frequencies.

6.7.6 OITC Versus STC as Sound Insulation Rating of Windows

In the above discussion, we have referred to the STC values of


windows. As indicated in Section 5.13, STC is good index of sound
insulation of a component against speech-dominated noise, while
OITC is a better index against transportation noise. Since windows
are meant primarily to insulate against outdoor noise, OITC is a more
appropriate index to compare the effectiveness of one window with
another. Some manufacturers of sound insulating windows provide
both STC and OITC values[6.41.
r\irborne Sound Insul ation Practice 137

Gypeum
eheathlne

Air epace
Reelllent channel

Flaehlne et--- Gypeum 1'1oard


Fli'1er1'1oard
Brick veneer

SECTION TROUGH A FIELD-ASSEMBLED SOUND INSULATING WINDOW

Perforatlld cover over a


poroue al'1eorl'1er

A PREF ABRICATED SOUND


INSULATING WINDOW.
Window eample courteey of
Waueau Metals, Waueau,
Wleconeln. Photo j,y Madan
Mehta.

6.22 Sound insulating sliding windows.


138 Chapter6

6.8 SOUND INSULATING Doors are more difficult to treat acoustically than windows. A typical
residential hollow core wood door has an STC of only 17; a typical
DOORS commercial solid core wood door has an STC of only 20. Therefore,
doors should be avoided as far as possible in walls with high sound
insulation.
The insulation of doors is based on the same principles as sound
insulating windows. Because they are operable, the most important
factor that influences the transmission loss of a door is the leakage
of air through gaps between the door and its frame. A typical door
\
has nearly 6 mm < in.) wide gaps at the top and bottom edges, and
.
3 mm (1/8 in.) on sides.
Unless these gaps are properly sealed, the STC of a door cannot
exceed 20, regardless of the increase in the weight of the door or
improvements in its construction, Table 6.5. Rubber and neoprene
gaskets pressed between the door and the frame are quite effective,
but may need periodic replacement to maintain the efficiency of the
seal.

Table 6.5 Approximate STC Values of Conventional Doors

Door type Surface mass STC STC


kgtm2 (lbfft2) (Unsealed) (Sealed)

Hollow core wood door 7.0 (1 .5) 17 20

Solid core wood door 20.0 (4. 1) 20 28

Hollow core steel 25 (5. 1 ) 20 30

Two solid core wood or 20.0 (4. 1) 28 40


two hollow core steel doors each
with 70 mm (2.8 in.) space

Two solid core wood or 20.0 (4. 1) 42 50


two hollow core steel doors each
with 230 mm (9 in.) space

Rubber or neoprene seals work well at the bottom edge of the


door if a threshold is used. If a threshold is not used and a scraper
seal attached to the door bottom is used, it will mark the (concrete)
floor, or wear the carpet, Figure 6.23. Sound retardant door
manufacturers provide various sophisticated sealing mechanisms.
An automatic drop seal (also called an automatic threshold
closer) at the bottom of the door is a better alternative to the scraper
seal. An automatic drop seal drops on the floor when the door is
shut. Two types of automatic drop seals are used:

attached to the middle of the door, Figure 6.24(a), and

attached to the exterior surface of the door, Figure 6.24(b).
In addition to the drop seals, the door can be hung on cam-lift
hinges. Cam-lift hinges drop the door and bring its fixed bottom
seal into contact with the threshold when the door is in closed position,
Figure 6.25.
Airborne Sound Insulation Practice 139

Scraper
&ea I

DO NOT RELY ON FELT, VINYL OR METAL WEATHERSTRIPPING FOR GOOD ACOUSTICAL SEALS. USE
WEATHERSTRIPPING MADE OF CLOSED-CELL SPONGE RU 66ERS OR HOLLOW EXTRUSIONS OF SOLID
ELASTOMERIC MATERIALS FOR GOOD ACOUSTICAL SEALS. MAGNETIC SEALS ARE ALSO AVAILABLE.

6.23 Perimeter door seals.

(a) In middle of door (a) At the exterior of door

6.24 Automatic drop seal.

(a) Hlf16e when door I& open (17) Hlnee when door I& c;loeed

6.25 A cam lift hinge: (a) door open position, (b) door closed
position.
140 Chapter 6

Where higher STC values are needed, two separate doors, with
a sound lock space should be used, Figure 6.26. A two-door assembly
in which the doors are connected together can be used in situation
where space for a sound lock is not available, Figure 6.27.
Doors with glass vision panels must have the glass assemb1)1
meet the acoustical requirements of the door. Sound rated door
manufacturers u sually limit the glazed area to meet STC
specifications.

EXECUTIVE SECRETARY

CORRIDOR

6.26 Sound lock with two doors.

6.27 Double door for the anechoic


chamber of the Callier Hearing and
Speech Center, Dallas, Texas. Photo by
Madan Mehta.

6.8.1 Hollow Door Frames

Hollow metal door frames provide a convenient route for sound to


go through. This is particularly so if the door frame is not properly
sealed at the jambs and the lintel. It is important that the hollow
metal door frames be densely packed with fiberglass or mineral wool
in case the door is installed in a lightweight wall. If the door is
installed in a masonry wall, the frame should be solidly grouted with
a portland cement based grout, Figure 6.28.
Airborne Sound Insulation Practice 141

Fll:lerglaee filling
lnelde frame

. .

6.28 Treatment of hollow metal doorframes in sound retardant doors.

6.1 Brick Institute of America, Reston, Virginia: "Sound Insulation - Clay


REFERENCES
Masonry Walls", Technical Note 5A, 1986.
6.2 National Concrete Masonry Association, Hemdon, Virginia: "Estimating
Sound Transmission Class of Concrete Masonry", NCMA -TEK 9, 1972.
6.3 American Society for Testing and Materials, Philadelphia: "Standard
Practice for Installing Sound-Isolating Lightweight Partitions", ASTM
Standard E 497-89(94).
6.4 American Society for Testing and Materials, Philadelphia: "Standard
Practice for Determining the Acoustical Performance of Exterior
Windows and Doors", ASTM Standard E 1425-9 1 .
Structure-borne
Sound Insulation
(Impact Isolation)

As discussed in Section 5.1, any vibration or impact causing object


that is rigidly attached to a building element will cause the element
to vibrate. It is the sound radiated by this element that we refer to as
structure-borne sound. The greater the radiated sound, the lower the
element's structure-borne sound insulation. A lightweight element
will vibrate more vigorously, giving a lower insulation. Mass
(weight) is, therefore, an important determinant of structure-borne
sound insulation of a barrier.
The sound radiated by a vibrating barrier reaches the receiver as
airborne sound. Therefore, the principles of airborne sound insulation
apply to structure-borne sound as well. For instance, providing a
two-leaf barrier provides a higher structure-borne sound insulation
than a single-leaf barrier.
That is why, a resiliently supported ceiling below a floor with a
layer of fiberglass in the intervening cavity improves structure-borne
sound insulation also. However, it is important to appreciate that
the ceiling be resiliently supported to improve structure-borne sound
insulation. In fact, structural isolation, also referred to as structural
discontinuity, provided by resilient (i.e . , impact or vibration
absorbing) attachments is the single most important factor in
improving structure-borne sound insulation.

143
144 Chapter 7

Impact and Vibration

Impact and vibration are the two most common sources of structure
borne sounds. Impact is the result of a force that occurs for a short
duration. Though an impact force may be repetitive, its repetition is
usually not periodic in nature. Vibration, on the other hand, is periodic
and continuous.
Walking, jogging and dancing are obvious examples of impact
sounds. Other impact sources are playing basketball, bowlingl
wheeling equipment and furniture, slamming of a door, etc. Vibration
is usually produced by machinery and equipment mounted on floors.
such as air conditioning equipment, fans, pumps etc. VibratioQ
control is best achieved by mounting the equipment on vibratiod
isolators. Since this is a specialized subject, it is covered separatel1
in Chapter 15.
In this chapter, we shall deal only with impact sound insulation.
Although airborne sound insulation is required of all barriers - wallsi
floor-ceiling assemblies and roofs - impact sound insulation is
primarily required of floors, because most impact-producing sourcet
rest on floors. Therefore, this chapter is limited to the sound insulati
of floor-ceiling assemblies only.

Insulation and Isolation

In most architectural acoustics literature, the terms "sound insulation1


and "sound isolation" are used synonymously, although, there is a
subtle difference between the two terms. Sound insulation is similllll
to thermal insulation, and is the reduction of sound energy as the
sound passes through an element from one side to the other.
The reduction in sound energy caused by isolating the sound
source from the receiver is referred to as sound isolation. Thus, a
reduction in the transmission of sound energy obtained througt
structural discontinuity or break is referred to as sound isolation.
Similarly, enclosing a sound source in an enclosure is also a form of
sound isolation.
Since the most important factor that affects the transmission of
impact sound from one side of the floor to the other is structural
isolation, the term "impact isolation" is used interchangeably with
"impact insulation".

7.1 IMPACT INSULATION The structure-borne sound insulation of a floor-ceiling assembly


measured in a two-room set-up, one room above the other. The fl(}(j
CLASS (IIC) between the two rooms has an opening in which the floor-ceiliDI
assembly, to be tested, is tightly fitted, Figure 7 . 1 . A standard tappin
machine, which has five equally spaced hammers, is placed on the
test assembly to produce impact at a constant rate. The tappinl
machine noise transmitted to the lower receiving room is measurel
in sixteen one-third octave bands, from 100Hz to 3,150 Hz. The
greater the noise level in the receiving room, the lower the sound
insulation of floor-ceiling assembly.
Using the above noise level data, a single number rating of
structure-borne sound insulation of the assembly is obtained by
comparing it with a standard contour, Figure 7 .2. The rating st
obtained is called the impact insulation class (IIC), and the standruf
contour is referred to as the IIC contour[?.ll.
Structure-borne Sound Insulation 145

SOURCE
ROOM

7.1 Experimental set-up for RECEIVING


measuring structure-borne sound ROOM

insulat ion of a floor-ceiling


assembly.

The procedure to determine the ne value is similar to that of


determining the STe value of an assembly. The measured noise
levels are plotted on a graph paper. Next we overlay the ne contour
on this plot and move the contour vertically as far down as possible
until the following two conditions are met.

The sum of deficiencies at 16 one-third octave bands does
not exceed 32 dB .
The maximum deficiency at any single one-third octave does
not exceed 8 dB .
A deficiency is a measurement that lies above the ne contour,
not below the contour (unlike the STe measurement). When both
conditions are met, the noise level corresponding to 500 Hz is
subtracted from 110 dB . The resulting value is the ne of the
assembly, as further explained at the end of this chapter.
In stating the ne value, the unit dB is omitted. Therefore, ne is
simply a number, just like the STC. The greater the ne value, the
higher the structure-borne sound insulation of the assembly.

7.2 Standard IIC contour.


146 Chapter 7

Note that the shape of ne contour is reverse of STC contour


(see Figure 5.16). The reason is that STC is determined from the
transmission loss data- the difference in levels between the source
room and the receiving room. The greater the transmission loss, the
higher the STC. ne, on the other hand, is determined from the noise
levels in receiving room. The greater these noise levels, the smaller
the ne.
A major criticism of ne rating is that it does not correlate well
with the ear's perception of insulation. It is highly skewed in favor
of low frequencies. Consequently, a lightweight floor (e.g., a plywood
subfloor on wood floor joists) whose structure-borne sound insulation
is worse than a heavy concrete floor, particularly at low frequencies,
may have a higher ne rating than a concrete floor. Despite the
criticism, no better single number rating procedure has yet been
agreed upon.
Another criticism ofnC rating is that the massive impacts (people
jogging and dropping weights etc.) in modem exercise facilities are
not represented by the standard low-mass hammers of the tapping
machine.

7.2 STRATEGIES TO In general, there are four basic strategies available to increase the
structure-borne sound insulation of a floor-ceiling assembly, as listed
INCREASE IMPACT below and discussed in the following sections.
INSULATION Soft or resilient floor covering
Resiliently supported floor- floating floor
Resiliently supported ceiling
Structural discontinuity in floor and ceiling - reducing
flanking transmission through the structure

7.3 SOFT OR RESILIENT The best means of insulating a floor against structure-borne sound is
to weaken the impact on the floor at the source - before the impact
FLOOR COVERING becomes structure-borne. Thus, a soft floor covering, such as a carpet
backed by a foam underlayment (pad), is an excellent way of
improving the structure-borne sound insulation of a floor.
For example, a 6 in. thick bare concrete slab has an STC rating
of nearly 55, but its ne rating is only25. The same slab when covered
with a pad and a carpet gives an IIC rating of nearly 85 (an
improvement of 60 points) , but its STC rating remains unchanged at
150 mm (6 In.)
55, Figure 7.3.
thick concrete
elal:> The increase in structure-borne sound insulation due to a carpet
is far greater for a hard inflexible floor such as concrete than for a
relatively flexible wood floor. For example, a typical residential
floor with a plywood subfloor and gypsum board ceiling attached
directly to floor joists gives an IIC of 34 and an STC of 38.
If the same floor is covered with a pad and a carpet, its IIC
IIC25 IIC 85 increases to 55 (an improvement of21 points) and the STC increases
5TC55 5TC55 to 39, Figure 7 .4. The small increase in STC is partially due to the
(airborne sound) absorption provided by the carpet and partially due
to the covering of joints of the floor by the carpet.
7.3 Increase in IIC of a concrete slab
Although a carpet is the best way to improve the structure-borne
by the addition of a pad and a carpet. insulation of a floor, resilient floor coverings such as cork, rubber
Adapted primarily from Reference 7.2. and vinyl also provide some improvement, Table 7.1.
ltructure-borne Sound Insulation 147

Table 7.1 Approximate Improvement It is important to emphasize that a soft or a resilient floor covering
in IIC for Some Floor Coverings has virtually no effect on airborne sound insulation, except that a
carpet, because of its absorption at high frequencies may slightly
increase the airborne sound insulation at these frequencies.
,...

Floor Improvement
in ne

Pad and carpet on:


Wood floor 20
Concrete floor 60 16 mm (5/8 In.)
Pad and carpet
thick plywood
Vinyl, rubber, etc., on:
Wood floor 5
Concrete floor 7

50 X :300 (2 X
fdapted from Reference 7.3. 12) floor jolet

1:3 mm (112 In.)


IIC34 IIC55
thick eypeum
board celllne 5TC38 5TC39

7.4 Improvement in 11C of a conventional re sidential woodfloor by


the addition of a pad and a carpet.

7.4 FLOATING FLOOR Although a soft floor covering improves the structure-borne sound
insulation of a floor, in many situations a hard concrete or wood
surface is required. In such a situation, a floating floor is the answer.
Unlike a carpet or a resilient floor covering, a floating floor also
increases the airborne sound insulation. Thus, a floating floor is
used where high values of both STC and IIC are required.
A floating floor is an additional layer of floor (concrete or wood)
supported on a structural floor (concrete or wood) through resilient
mounts. To be effective, the floating floor must be isolated at all
sides from walls or other building components, so that the impact or
vibration from the floor does not flank to other parts of the building
through the wall. This isolation is provided by a perimeter isolation
board (fiberglass board or a plastic foam), Figure 7.5.

Wall
Caulk here
Perimeter leolatlon board

Floatlne floor
R"elllent layer
Structural floor
7.5 Essential elements of a floating floor.
148 Chapter?

7.4.1 Floating Floor on Conventional Wood Floor

A simple plywood floating floor is shown in Figure 7 .6. It consists


of plywood panels glued and nailed to 50 x 100 (2 x 4) wood sleeper s,_
The sleepers are laid over 25 to 40 mm (1 to l1 I in.) thick compress
f
fiberglass boards, placed over a conventional p ywood subfloor. Note
that the sleepers are simply laid over fiberglass boards with no
attachment to the structural floor.
Because of its low cost and simple construction, this floor is
commonly used for homes and apartments. With a gypsum board
ceiling attached to floor joists through resilient channels, this floor
ceiling assembly gives an STC of nearly 55 and an ne of nearly 50.
With a carpet and pad, an ne of nearly 70 may be achieved.

Plywood floatlne
floor
50 X 100 (2 X 4)
wood elpr

Compreeed
fll:lerelaee
Plywood
eul:lfloor ------.=" lb=::--oo="""",_.,

7.6 Plywood floating floor on


IIC50 IIC70
conventional wood floor.
5TC55 5TC56

A major disadvantage of such a floor is that, due to its light


weight, it transmits low frequency impact noise, which the lower
floor occupants perceive as thumps or rattling sound as people walk
on the floor above. This fact is not obvious in ne values since, as
stated in Section 7.1, the ne contour is skewed in favor of low
frequencies, which overrates lightweight floors.
Another disadvantage of a lightweight floor is that it creates more
noise within its own space. Tapping on a lightweight wall versus a
heavy concrete or masonry wall makes this fact at once obvious.
A layer of portland cement or gypsum concrete in place of
plywood provides the necessary weight and improves low frequencj
1
insulation, Figure 7.7. In practice, nearly 40 to 50 mm (1 /2 to 2 in.)
thick lightly reinforced cement (or gypsum) concrete layer is used.
A polyethylene sheet between compressed fiberglass and concrete
provides necessary waterproofing. This assembly gives an ne of
nearly 58 and an STC of nearly 60. With a pad and carpet, an IIC of
up to 80 is achieved.
Structure-borne Sound Insulation 149

Portland cement concrete


or eypeum concrete
Polyethylene eheet

Plywood
eul:rl'loor

channel
Gyp eum
l:>oard T
IIC 58 IIC80
7. 7 Concrete floating floor on 5TC60 5TC60
conventional wood floor.

One manufacturer of floating floor system uses a honeycomb


floorboard in place of compressed fiberglass. This floor board
consists of a thin layer of fiberglass laminated to both sides of a
cellulosic honeycomb core, Figure 7.8. With a total thickness of
only 16 mm (5/8 in.), it makes an excellent cost-effective alternative
to compressed fiberglass board, and is particularly suitable with
concrete-topped floating floors.

7.8 Honeycomb resilient floor board. Sample


courtesy of Kinetics Noise Control Inc., Dublin,
Ohio. Photo by Madan Mehta.
150 Chapter 7

Table 7.2 Approximate IIC and STC 7.4.2 Wood Floating Floor over Concrete Structural Floor
Values for Some Floating Floors
A wood floating floor over a concrete structural floor is ideal for
aerobic exercise halls, gymnasiums, dance floors, high-rise
floating floor IIC STC apartments, etc., particularly over suspended concrete slabs.
Although there are different versions, a typical wood floating floor
over a concrete structural floor is shown in Figure 7.9. Impact
Wood floating floor on: absorption is provided by high-density fiberglass blocks 50 mm x
Wood structural floor 52 58 50 mm x 50 mm (2 in. x 2 in. x 2 in.).
Concrete structural floor 64 62 The fiberglass blocks are bonded to 50 x 100 (2 x 4) wood
sleepers at nearly 300 mm (12 in.) on centers. Depending on the
Concrete floating floor on: load on the floor, the sleepers are simply laid (not attached) on the
Wood structural floor 58 60 concrete floor at 300 to 400 mm (12 to 16 in.) on centers.
Concrete structural floor 74 62 The space between sleepers is filled with low-density fiberglass.
Next a layer of plywood panels is nailed to the sleepers. Finally, a
These values are approximate and are provided second layer of plywood is adhesively bonded and nailed to the lower
to compare one type of floating floor with the plywood, with staggered joints. The STC and IIC values depend on
other. The actual values depend a great deal on the thickness of the structural floor. A floor covering, such as a
the thickness of materials, depth of air cavity, hardwood floor, carpet, etc., provides the floor finish. Table 7.2
the presence or absence of fiberglass in the cavity, gives some representative STC and IIC values of floating floors.
etc. The values represent bare floors with no
carpeting.

Plywood ---..-'< Perimeter


Adheelve

..- 50 x 100 ( 2 x 4)
wood eleepere
- Low-denelty
fli:lerglaee i:llanket
Wood eluper

7.9 Wood floating floor on a reinforced concrete structural floor- a system supplied by Kinetics Noise
Control Inc., Dublin, Ohio.
ltructure-borne Sound Insulation 151

7.4.3 Floating Concrete Slab over Concrete Structural Floor

A concrete topped floating floor is similar to a wood floating floor


Plae;tlclzlld described previously, and is commonly used in suspended floors for
p rotllctlvll mechanical rooms, squash and racquetball courts, exercise rooms,
e;urfacll gymnasiums, etc. A typical concrete floating floor consists of nearly
100 mm (4 in.) thick reinforced concrete slab supported on high
density fiberglass blocks, placed at nearly 300 mm (12 in.) on center
Flbllr!!lae;e; over the concrete structural floor, Figure 7 . 1 0.
CUTAWAY SECTION THROUGH A
HIGH-DENSITY FIBERGLASS BLOCK

Concrlltll floatln!!
floor

Plywood

L Low-dllnelty flbllr!!laee blanlcllt with


hl!!h-dllnelty flbuglae;e blocke bondlld
to a polyllthylllnll e;hllllt In a roll form

7.10 Concrete floating floor assembly supplied by Kinetics Noise Control Inc. , Dublin, Ohio.

Plywood panels over fiberglass blocks function as a permanent


form for the floating concrete slab. The air space between the floating
concrete slab and the structural floor is filled with low density
fiberglass. In fact, one manufacturer supplies the blocks and low
density fiberglass blanket bonded to a plastic sheet, all packaged in
rolls, Figure 7 . 1 1 .

7.11 50 mm x 50 mm x 50 mm(2 in. x 2 in. x 2 in. ) high


densityfiberglass blocks and low-densityfiberglass blanket
bonded to a plastic sheet- by Kinetics Noise Control Inc.,
Dublin, Ohio. Photo by Madan Mehta.
152 Chapter 7

Thus, the low-density fiberglass blanket with high-density


fiberglass blocks bonded to it is unrolled over the structural floor
and covered with plywood panels,. Figure 7 . 1 2. The panels are
connected together at joints with steel plates, Figure 7 . 13, and a
polyethylene sheet placed over them. Reinforcement is now laid
and concrete. poured.

Perimeter Isolation

7.12 Plywood form and high-density fiberglass blocks bonded 7.13 Connecting plywood panels together.
to low-density fiberglass blanket. Photo by Madan Mehta. Photo by Madan Mehta.

7.4.4 Jack-up Floating Concrete Slab

Lifting l:>olt A jack-up concrete floating floor has the same finished appearance
as the floor described previously. The difference between the two
floors is in the processes of construction and the impact-absorbing
mounts. Each mount consists of a neoprene block enclosed in a
cast-iron housing, Figure 7. 14. The housing is supported on a lifting
bolt, which in turn rests on the neoprene block. The cast iron housing
has two cantilevered brackets to support reinforcing bars.
The process of construction is shown in Figure 7. 15(a). First, a
plastic sheet is laid over the structural concrete floor. This sheet
works as a bond breaker between the structural floor and the floating
slab. Depending on the load on the floor, the mounts are then placed
at 600 to 1 ,200 mm (2 to 4ft) on centers each way. Reinforcing bars
are now placed over the brackets of the mounts. Additional
reinforcement is now laid over the previously laid reinforcement,
and concrete is poured.
7.14 Cast-iron housing and neoprene
mount.
lructure-borne Sound Insulation 153

After the slab has cured and attained the necessary strength, it is
jacked up with the help of jack screws, Figure 7 . 15(b). One or two
people can lift a large floor little by little, ensuring a uniform lift at
all points. The total lift of the floor need be only 25 mm (1 in.), but
a greater lift may be specified for a higher insulation. A jack-up
floor is particularly suitable for heavily loaded floors, or floors that
have an irregular shape.

Perlmr
le;olatlon

(a)
Reinforcement

(!?)

7.15 Jack-up floating concrete floor.


154 Chapter 7

7.5 RESILIENTLY Apart from a soft floor covering and/or a floating floor, a continuous
resiliently supported ceiling of a nonporous material can augment
ATTACHED CEILING the structure-borne sound insulation of a floor, just as it augments
the airborne sound insulation. For a conventional wood floor, resilient
channels are used (see Figure 6.9).
For a suspended ceiling, the use of ceiling isolation hangers,
available from several manufacturers, is recommended, Figure 7 . 16.
In addition, ensure that there are no sound leaks in the ceiling and
that the entire ceiling is airtight, including light fixtures.

7.16 Suspended ceiling and


resilient hangers.

7.6 DISCONTINUIT Y IN A major point of difference between airborne and structure-borne


sound is that while airborne sound decreases markedly with distance,
FLOOR AND CEILING structure-borne sound (vibration) can travel through the structure
with very little decrease along its path. This is because of limited
internal damping in most building materials such concrete, steel,
masonry, etc. It is because of the lack of damping in steel rails that
one can hear an approaching train, although unable to see the train.
Structure-borne Sound Insulation 155

Due to its low attenuation, structure-borne sound can present


itself in a faraway space, particularly if a lightweight element is
attached to the structure there. The only way to attenuate structure
borne sound transmission along its path is to provide a discontinuity
or break in the floor and ceiling. Thus, the joint between a resiliently
supported ceiling and wall must be isolated with the help of an impact
absorbing material, Figure 7.17.
Similarly, a floating floor must be isolated from the adjoining
structure with a perimeter isolation board. (In fact, as the reader
may have realized, the resilient mounts under a floating floor also
provide structural discontinuity). A wall separating the floated floor
room from other spaces must be isolated with the help of impact
absorbing material, Figure 7 . 1 8. All pipes and other penetrations
through the floors and ceilings must be similarly isolated.

Llehtweleht
- Joint fillet.! with wall
lmpa al1eorl11ne
material

lmpa
al1eorl11ne
material

Joint fillet.! with


lmpa al1eorblne
material

. . . . ... ..
7.17 Joint between suspended ceiling and wall. . . 0. 0

: D 0 '
u
"""'I'-
lmpa
al1eorl11 ne
material


,
Impact Impact
al1eorl11ne al1eorl11ne
material material

'o . . - ..:.f. .
. . .

7.18 Jonts between wall and floor.


156 Chapter 7

IIC VALUE FROM MEASURED DATA

This section illustrates the procedure of determining the ne of a floor-ceiling assembly through the
following example.
Let the measured sound pressure levels in the receiving room of a two-room set-up for a conventional
wood floor, whose cross-section is shown, be:
Frequency (Hz) TL(dB) Frequency (Hz) TL (dB)

100 72 630 66
125 69 800 61
160 70 1,000 56 SECTION THROUGH
200 73 1,250 52 FLOOR ASSEMBLY
250 71 1,600 52
315 69 2,000 53
400 68 2,500 54
500 67 3,150 51

To determine the ne of the assembly, first plot the above data to a suitable scale as shown in Figure
A. Next draw the IIC contour on a separate transparent sheet to the same scale, and overlay it on the
above plot.
Now move the ne contour as far down as possible until it satisfies the two conditions described in
Section 7. 1, summarized here for convenience.
The sum of deficiencies at 16 frequencies should not exceed 32 dB, and,

The maximum deficiency at any point should not exceed 8 dB.
A deficiency is defined as any measured value lying above the contour. The location of the IIC
contour shown in the diagram satisfies the above conditions, and its placement is governed by the first
condition. Reading the SPL value at 500Hz, we observe that it is it is approximately 65. Hence
IIC = 110- 65 = 45, which can also be obtained directly from the right hand vertical axis (note
reversed values).
The second curve in Figure A shows the measured SPL values for the same assembly, but
with a 40 mm (1 1/2 in.) thick cement concrete topping, highlighting the improvement obtained
at low frequencies through the use of a concrete topping.

r--.--,----. 35
IIC contour le fitted to thie curve

FIGURE A

SPL IIC
(dB)
-L-50 ----L-r--+-L-+-'

1-q--L
.
_ -4--+--
J:.
' ... 1' . .. . . J
..

l"f
-H--+--
. . ... .... . ........

55

Data for the eame floor


-- -- -------+
i -+ - 60
--

with 40 mm (11/2 In.)


thick concrete topping
160 250 400 630 1,000 1,600 2,500
125 200 315 500 800 1,250 2,000 3.150

Frer:tuency (Hz)
Structure-borne Sound Insulation 157

REFERENCES 7.1 American Society for Testing and Materials, Philadelphia: "Standard
Classification for Determination of Impact Isolation Class (IIC)", ASTM
Standard E 989-89(94).
7.2 Harris, C. M.: Noise Control in Buildings, McGraw Hill, Inc., New York,
1994, p. 6.4.
7.3 Harris, C. M.: Noise Control in Buildings, McGraw Hill, Inc., New York,
1994, p. 6.9.
Noise Control in
Buildings

The ill effects of noise are well documented. Apart from general
annoyance, high noise levels cause interference with task
performance, speech communication and sleep. Noise also affects
the health and general well-being of humans. Exposure to sufficiently
high noise levels for extended periods can cause stress, hypertension,
and hearing loss.
This chapter deals with various means of reducing noise inside
buildings to acceptable levels. Note the word "acceptable" here,
since a complete elimination of noise is neither possible, nor even
desirable. A certain amount of noise is necessary for our well-being.
Anyone who has spent some time in an anechoic chamber knows
that a totally silent space can be as disturbing as a noisy space.
The noise level in a building is due to a combination of two
sources: interior noise - that produced inside the building - and
exterior noise. The primary source of interior noise is the use and
occupancy of the building. For example, in an industrial building,
most noise is due to the manufacturing process.
In shopping centers, educational establishments and offices,
interior noise is due to human conversation in addition to task-related
activities. In restaurants, clubs, and similar commercial spaces, the
interior noise is due to human conversation, music, and eating-related
activities. A noise source common to almost all interiors is heating
and air conditioning noise.

159
160 Chapter 8

Outdoor noise consists primarily of road traffic and outdoor air


conditioning equipment. For neighborhoods close to a railroad or
an airport, the outdoor noise is also due to railway trains and/or
aircraft.

8 8
Path All noise, whether interior or exterior, is subject to a three-way
control mechanism - at the source, along the path, and at the
receiver. In fact, the source-path-receiver principle is so fundamental
that it is not only applicable to building noise, but to all types of
noise control.

Noise Control at the Source

The most effective control takes place at the source. For example,
road traffic noise is best controlled by making the automobiles as
well as the road surfaces quieter. Federal and state regulations control
noise emissions from various sources, particularly motor vehicles.
For instance, the California Vehicle Code requires that the noise level
from a motor vehicle, measured at a distance of 15 m from it, should
not exceed 85 dBA if the vehicle speed is less than 35 miles per
hour. If the speed is greater than 35 miles )Jer hour, the corresponding
noise level should not exceed 90 dBA[s.n.
Most noise reduction at the source is achieved during the design
and manufacturing phases of the product. It is only when further
noise reduction at the source is not practical that the control must
take place along the path and/or at the receiver.
Noise Control Along the Path

Controlling noise along the path of its transmission is at the heart of


most noise control measures, and it is this that is discussed in this
chapter. Three means are available to control the transmission of
airborne sound along its path:
Separating noisy areas from noise-sensitive areas through
architectural design and site planning.
Erecting sound insulating barrier/s between noisy spaces and
quiet areas.
Providing sound absorptive treatment in source room and/
or receiving room.
Although the first two measures apply to both interior and exterior
noise control, the last measure is applicable only to interior noise.
Recently noise control through active noise cancellation has become
possible (see Chapter 9).
Structure-borne noise control measures are discussed in Chapters
7 and 15.

Noise Control at the Receiver

Once the sound has reached the receiver, very little can be done to
control it, except to muffle it at the listener's ears through ear
protectors. Experience in noisy industrial facilities shows that the
use of ear plugs or hearing defenders is not liked by workers.
Therefore, these measures are recommended only where the noise
cannot be reduced by the other two means- at the source and along
the path.
Noise Control in Buildings 161

8.1 INTERIOR NOISE Before proceeding to discuss noise control measures, a question that
must be addressed at the outset is: what are the maximum acceptable
CRITERIA interior and exterior noise levels? Obviously, the maximum
acceptable interior noise level is a function of the type of occupancy.
The acceptable level for a noise-sensitive area, such as a recording
studio or a concert hall is much lower than that for an office or a
dwelling.
The acceptable interior noise level is generally based on the
degree of interference produced by noise on task performance. The
acceptable exterior noise level, on the other hand, is based on the
degree of annoyance produced by noise in our use of outdoor spaces.
In a noisy manufacturing facility, the acceptable noise level may be
based on hearing damage risk.
Once the acceptable noise levels are known, we can then
determine the required noise reduction. For example, if the acceptable
noise level for an office space is 50 dB, and the noise level in the
adjoining space is 80 dB, then the required reduction is 30 dB.
An acceptable interior noise for given activity cannot be specified
in dB levels because interference or annoyance produced by a noise
is frequency dependent. As discussed in Chapter 2, our ears are not
equally sensitive to all frequencies. Consequently, acceptable noise
levels cannot be specified by a single number, but in terms of a
detailed noise spectrum.
Since a dBA level is more representative of the ear's sensitivity,
efforts have been made to specify acceptable levels in dB A, but even
this has been found unsatisfactory. Two noise environments with
different spectra (and hence different potential for annoyance and
effects on task performance) can have the same dBA value.

8.1.1 Speech Interference Level

Speech communication is the most critical activity in most spaces.


Therefore, an obvious approach on which to base the acceptable
interior noise level is the degree of interference a given noise level
will cause on speech communication.
The bulk of speech energy lies between 300 and 5,000 Hz. In
Section 2.4, it was noted that a sound of given frequency is most
easily masked by a sound of the same frequency. Thus, in determining
speech interference, we need consider only noise lying between 300
to 5,000 Hz.
The arithmetic average of sound pressure levels in four octaves
centered at 500, 1 ,000, 2,000 and 4,000 Hz1 , has been defined as the
speech interference level (SIL). That is:

SPL5oo + SPLt,ooo + SPLz,ooo + SPL4,ooo


SIL =
(8. 1 )
4
Thus, i f the sound pressure levels due to background and
occupancy-related noises in a space in the above four octaves are as
given below, the SIL is 39.5, i.e., 40 dB.
Frequency (Hz) 500 1,000 2,000 4,000 SIL = (46 + 40 + 37 +
SPL (dB) 46 40 37 35 35)/4 = 39.5 dB

1 From Figure 1. 7, we see that the octave band centered at 500 Hz begins at 354
Hz, and that centered at 4kHz ends at 5.6 kHz.
162 Chapter 8

5pealcer-lletener dletance (ft)

0.4
90 0.8 1.6 :3.:3 6.5 1:3 26 52

80 8-
0


I 4'1.
.:'ot.r. 'tf

i
8.1 Speech interference levels as afunction
of(male) speaker-listener distance for just 70
-
reliable speech communication(under free- .V.o,..
SIL
field conditions, i.e. , the speaker is not aided (dB) 60 - /
by reverberation). To determine speech
interference level for female speakers, 50
subtract 4 dB from SILobtained from this
figure, since 4 dB represents the difference
40
between the speech level of an average
female and an average male. For instance,
SILfor a male speaker at 1 m distance, in :30
normal voice, is nearly 55 dB. T he 0.125 0.25 0.5 1.0 2.0 4.0 8.0 16.0
corresponding S/Lfor a female speaker is 5pealcer-lletene;r dletance (m)
51 dB. Adapted from Reference 8.2.

Based on extensive measurements of speech communication


between speakers and listeners under different speech interference
levels, a relationship between SIL and speaker-listener distance for
reliable speech communication has been prepared[82l. This is shown
in Figure 8. 1. Thus, if the SIL of the noise is 40 dB, two males will
be able to (barely) communicate with each other in normal voices
up to a distance of nearly 6 m (20 ft). If they raised their voices, they
would be able to communicate up to distance of nearly 1 1 m (35 ft).

8.1.2 NC Curves

Specifying background noise levels based on speech interference


levels has serious limitations, since it does not consider other effects
of noise such as annoyance, interference with activity, listening tq
music, etc. Based on extensive interviews with people in officesJ
public spaces, and manufacturing facilities, a family of octave band
sound pressure level curves has been developed to specify acceptable
background noise levels. The curves are called noise criterion curves,
abbreviated as NC curves[83l. Each curve is designated by a number,
such as NC 20, NC 30, etc., Figure 8.2.
Note that the shape ofNC curves is similar to the equal loudness
contours discussed in Chapter 2, and highlights the ear's lower
sensitivity to low frequency noise.
The curves were originally developed from measurement studies
of how much heating ventilating and air-conditioning (HVAC) noise
interferes with speech communication, and with listening to music,
radio, or television. However, the curves are now used extensive!
to specify acceptable interior noise levels from all types of sources
including HVAC noise but excluding occupancy-related noise,
provided the noise is of a continuous nature.
Noise Control in Buildings 163

- - - - - - - -f- - - - - - - - - - - - - - J_ _ _ _ _ _ _ l
90.----------
1
______________ ______________ _ ______________

--- - + + + ! + +
- - - - - - - - - - - - - - - - - - - - - - - - - - -r- - - - - - - 1- - - - : : : :::
80 - - -- -------------- ------------- -------------- ------------- ----------- -------------

:: :: :: :: :: ::
: : : : : : :- --
T _ _ __ __ __

NC-70

NC-65

NC-60

NC-55

NC-50
SPL (dB)
NC-45

10 '-----
-- --'----------------
--' _j_ ____ _j_____ ___j_ __j
__ __ NC-15
63 125 250 500 1,000 2,000 4,000 8,000
Frequency (Hz)

NC Value Frequency (Hz)

63 125 250 500 1,000 2,000 4,000 8,000

NC-15 47 36 29 22 17 14 12 11
NC20 51 40 33 26 22 19 17 16
NC-25 54 44 37 31 27 24 22 21
NC-30 57 48 41 35 31 29 28 27
NC-35 60 52 45 40 36 34 33 32
NC-40 64 56 50 45 41 39 38 37
NC-45 67 60 54 49 46 44 43 42
NC-50 71 64 58 54 51 49 48 47
NC-55 74 67 62 58 56 54 53 52
NC-60 77 71 67 63 61 59 58 57
NC-65 80 75 71 68 66 64 63 62
NC-70 83 79 75 73 71 69 68 67

8.2 NC curves and their octave band values. Adapted from Reference 8.3.
164 Chapter 8

To determine the NC value of a given environment, we first


measure the sound pressure levels in each octave band from 63 Hz
to 8 kHz. The measured values are then plotted on the same graph
as the NC curves. The nearest NC curve that is completely above
the noise plot is then shifted downward so that it becomes tangent to
the plot. The rating of the NC-curve minus its downward shift gives
the NC value of the environment, see Example 8.1.
In determining the NC value of an HVAC noise, ensure that sound
pressure level measurements are taken in absence of occupancy
related noise - preferably in an unoccupied building.

Example 8.1 Determining the NC Value of an Environment

Determine the NC values of two environments for which the measured sound pressure levels are:

Frequency (Hz) 32 63 1 25 250 500 1,000 2,000 4,000 8,000

SPL(dB) 50 57 55 49 44 43 43 41 40 ..... Environment 1

65 65 59 52 40 36 33 30 33 . .... Environment 2

Solution: The above octave band values are plotted for both environments in Figure A. Superimposed
on these plots are NC-45 and NC-40 curves. We see that NC-45 is the nearest NC curve that lies
completely above environment 1 plot. Now if we shift NC-45curve down, we see that the downward
shift needs to be only 1 dB, and the NC curve becomes tangent to environment 1 plot at 2,000 Hz.
Hence, environment 1 is (45- 1), i.e., NC-44.
The same procedure is followed with with environment 2. This is also NC-44 environment, since
the NC-45curve becomes tangent to its plot at 125Hz through 1 dB downward shift.

FIGURE A

5PL
(dB)

NC-40

125 250 500 1,000 2,000 4.ooo .ooo


Fre(\uenc;y (Hz)
Noise Control in Buildings 165

8.1.3 RC Curves

From Example 8.1, we observe that two noise environments with


the same NC-rating (NC-44) have vastly different spectra.
Consequently, they will sound very different. In fact, the subjective
perception of environment 1 will be that it is "hissy", because it is
relatively rich in high frequencies. The subjective perception of
environment 2 will be that it is "rumbly" because it is relatively rich
in low frequencies.
However, the NC rating procedure does not distinguish between
the two environments and regards them as identical. Experience
with NC curves indicates that only when the measured values follow
the shape of the chosen NC curve over at least four octaves that the
environment will be judged as "neutral"- neither hissy nor rumbly.
Both hissy and rumbly environments are objected to by the occupants.
In order to address the above limitation, a family of curves known
as room criterion curves, abbreviated as RC curvesl8.4J, has been
developed, Figure 8.3. When the RC curves were first introduced
(in 1981), the American Society of Heating, Refrigeration and Air
Conditioning Engineers (ASHRAE) recommended the use of either
NC or RC curves to specify HVAC noise levels.
Since ASHRAE's 1995 publication, only RC curves are
recommended for use2. As explained later, RC curves provide a
more comprehensive evaluation of noise environments than NC

The: pre:ee:nce: of nolee: In thle region


lndlcate:e a probability of nolee: 80
lnduce:d vlbratlone In lightweight
walla and ceiling, which will be: felt
and/or will be audible: ae a rattling 70
eound. Region A re:pre:ee:nte high
probability and region B, a moderate:
probablty of e;uch vlbratlone;.

SPL
(dB)
RC-45
RC-40
RC-35

32 63 125 250 500 1,000 2,000 4,000


Frequency (Hz)

8.3 RC curves.
2 As of 1997, ASHRAE is still debating whether to use RC or a different
criterion being advanced by some researchers.
166 Chapter 8

curves. RC curves extend from 16 Hz to 4 kHz in place of 63 Hz to


8 kHz for NC curves.
RC curves are not really curves but straight lines with a 5 dB per
octave slope. Each curve is designated by a number, such as RC-30,
RC-35, etc. The number associated with an RC curve is the sound
pressure level at 1 kHz. Thus, an RC curve is easy to draw. For
instance if we wish to draw RC-40 curve, we start with 40 dB point
at 1 kHz and draw a straight line with a slope of 5 dB per octave.

8.1.4 RC Value of an Environment

To determine the RC value of an environment, we measure the octave


band levels from 16 Hz to 4 kHz and plot them on a graph paper.
The arithmetic average of sound pressure levels at 500, 1 ,000 and
2,000 Hz is the RC value of the environment. Starting from this
(average) value at 1000 Hz, we draw a straight line with a slope of 5
dB per octave.
Next we classify the environment as neutral, hissy or rumbly by
following the procedure given in Example 8.2. Thus, a given RC
rating, such as RC-36 will be classified as: RC-36(N), or RC-36(H),
or RC-36(R), where N stands for "neutral", H for "hissy" and R for
"rumbly".
A neutral environment is free from tonal exaggerations and will
be judged as unobtrusive or bland if its spectrum follows the RC
curve closely. More specifically, a neutral environment is one for
which: (i) the measured values at and below 500 Hz do not exceed
the RC curve values by more than 5 dB at any octave, and (ii) the
measured values at 1 ,000 HZ and above do not exceed the RC curve
values by 3 dB at any octave. If the environment meets both above
conditions, the designator (N) is placed after the RC value.
A rumbly environment is one in which the measured values at
and below 500 Hz exceed the RC curve by more than 5 dB at any
octave. If so, the designator (R) is placed after the RC value. If the
measured values at 1 ,000 Hz and above exceed the RC curve values
by more than 3 dB at any octave, the environment will be hissy. The
designator (H) is placed after the RC value of such an environment.
The shaded portion in Figure 8.3 is the region from 16 to 63 Hz
octaves in which perceptible vibrations in walls and ceilings may
occur, particularly in lightweight structures. Such vibrations may
be felt and may cause lightweight walls, cabinet doors, picture frames,
suspended ceiling, etc. , to rattle audibly. If one or more values fall
in the shaded portion, the designator (V) is placed after the RC value.
An environment may be classified as RC-XX(RV) if it is rumbly
and produces feelable vibrations, or RC-XX(HV) if the hissiness is
combined with feelable vibrations.
As far as possible, noise environment should be neutral. If it is
designated as H, R or V, it will generally be judged objectionable.
Apart from giving a more comprehensive description of noise, a
major benefit of RC value is that its numerical rating gives the
approximate speech interference level.

8.1.5 Criteria for Acceptable HVAC and Non-HVAC Noises

To be acceptable, HVAC noise should be low enough so as not to


mask desirable occupancy-related sounds. In offices, the desirable
sound is generally the conversation between people at different work
stations. In classrooms, teacher-pupil and pupil-pupil communication
Noise Control in Buildings 167

Example 8.2 Determining the RC Value of an Environment

Determine the RC values of the two environments of Example 8.1. The measured sound pressure levels
are repeated below for convenience.

Frequency (Hz) 32 63 1 25 250 500 1,000 2,000 4,000 8,000

SPL (dB) 50 57 55 49 44 43 43 41 40 .. . ... Environment l

65 65 59 52 40 36 33 30 33 ...... Environment 2

Solution: The above octave band values are plotted separately for both environments in Figures A and
B. For environment 1, the average of sound pressure levels at 500, 1,000 and 2,000 Hz is (44 + 43 + 41)
= 43.3 i.e., 43 dB. Hence, environment 1 is an RC-43 environment.

Now we must classify this environment as N, H, R or V. Therefore, beginning with the 43 dB point
at 1,000 Hz, draw a line at a slope of 5 dB per octave, to represent RC-43 line. Next, draw two lines
parallel to RC-43, one to the left of 500 Hz and 5 dB above RC-43 line, and the other line to the right of
1,000 Hz, and 3 dB above RC-43 line. These lines are designated as R and H respectively. Observe that
two environment 1 values are above the H line. Therefore, this environment will sound hissy. Hence
environment 1 is rated as RC-43(H).
The average of sound pressure levels at 500, 1,000 and 2,000 Hz for environment 2 is (40 + 36 +
33)/3 = 36.3, i.e., 36 dB. Therefore, draw a line with a slope of 5 dB per octave beginning with 36 dB
at 1,000 Hz. Then draw lines R and H, and observe that some environment 2 values lie above the R line.
Hence environment 2 will sound rumbly, and hence it is classified as RC-36(R). Note from Example
8.1 that both these environments were NC-44 environments.

FIGURE A FIGURE B

80 eo

70

SPL
(aB) 60 60

50

40

:30 :30

16 :32 6:3 125 250 16 :32 6:3 125 250 500 1,000 2,000 4,000

Frequency (Hz) Fre'luency (Hz)

Environment 1 Environment 2
168 Chapter 8

must stand well above HVAC noise. In recital and concert halls
with their unamplified sound, HVAC noise must not mask even the
faintest of performance sounds. Generally, HVAC noise should be
at least 10 dB lower than occupancy-related sounds in all octaves.
Table 8. 1 lists acceptable HVAC background noise levels in terms
of RC values for various occupancies. Conformance to RC values
generally requires a costlier HVAC system than conformance to NC
values. That is why NC values are extensively used even though
RC values provide a more reliable specification. Table 8. 1 also lists
NC values.
The corresponding approximate dB A levels of acceptable HVAC
noise are also listed in Table 8. 1. They may be used only in those
rare situations where instrumentation for octave band measurements
is not available to verify existing noise with that specified; dBA values
should not be used for specification purposes.

Table 8.1 Recommended RC and NC Values for Unoccupied Spaces

Space Recommended Recommended Approximate


RC(N) value NC value dBA value

Private residence, apartment, condominium 25 - 35 25 - 35 33 - 43

Hotels or motels:
Individual rooms, meeting rooms 25 - 35 25 - 35 33 - 43
Halls, corridors, lobbies 35 - 45 35 - 45 43 - 53

Office buildings:
Executive and private offices 25 - 35 25 - 35 33 - 43
Open plan offices 30 - 40 30 - 40 38 - 48
Circulation areas 40 - 45 40 - 45 48 - 53

Hospitals and clinics:


Private rooms and operating rooms 25 - 35 25 - 35 33 - 43
Wards, corridors and public spaces 30 - 40 30 - 40 38 - 48

Performing arts spaces:


Drama theaters, music teaching spaces 25 (max) 25 (max)
Music practice rooms 35 (max) 35 (max)
Concert and recital halls Consult an acoustical consultant

Laboratories (with fume hoods):


Testing/research with minimal speech
communication 45 - 55 45 - 55 53 - 58
Research with extensive telephone use 40 - 50 40 - 50 48 - 58
Group teaching 35 - 45 35 - 45 43 - 53

Churches, mosques and synagogues 25 - 35 25 - 35 33 - 38

Schools:
Classrooms up to 70 m2 (750 ft2) 40 (max) 40 (max)
Classrooms over 70 m2 (750 ft2) 35 (max) 35 (max)

Libraries 30 - 40 30 - 40 38 - 48

Courtrooms:
Unamplified speech 25 - 35 25 - 35 33 - 43
Amplified speech 30 - 40 30 - 30 38 - 48

Indoor stadiums and gymnasiums 40 - 50 40 - 50 48 - 58

Adapted from American Society of Heating, Refrigeration and Air Conditioning Engineers' (Atlanta) Applications
Handbook, 1995, page 43.5.
Noise Control in Buildings 169

Table 8. 1 values may also be used as the upper limit of non


r_level
Outdoor nolee
HVAC noises, such as traffic and equipment noise. For instance, the
/
Indoor background
nolr;e level required transmission loss of an external wall of a building situated
J
--"---
-
I
-1/ External wall
in a high traffic noise area may be determined on the basis that the
wall should reduce traffic noise to a value below that given in Table
Tranr;mltted 1 8. 1 for that occupancy (see Section 8.5.2).
nolr;e level ____j

8.2 INTERIOR NOISE Various interior noise criteria discussed in the previous section are
simply guidelines for good practice. Although generally followed,
LEGISLATION their use is entirely voluntary. The only mandatory interior noise
control in the United States is in situations where a possibility of
hearing damage risk exists.
Hearing damage results from exposure to high noise levels over
long periods of time . The Occupational S afety and Health
Administration (OSHA) of the U.S. Department of Labor regulates
Ta b l e 8.2 OSHA P e r m i s s i b l e
the exposure of workers to workplace noise through federal
Noise Exposure Limits legislation passed in 1970. This legislation sets the upper limit of
exposure to noise in a worklace based on daily noise dose.
Time duration Maximum According to the Act[S. J, the maximum permissible value of
per day (T) in permissible daily noise dose (D) for a worker is 1.0 (or 100%). The value of D is
to be calculated from the following expression.

..
hours exposure (dBA)

I
8 90 ; Ct + Cz + c, + .
D (8. 2)
6 92 Tt Tz T3
4 95
3 97
2 lOO
1.5 102 where, a C value (C , C2, C 3 , etc.) is the total daily exposure time of
J
1 105 a worker to a specittc noise level, and a T value (T1 , T2 , T3, etc.) is
0.5 1 10 the corresponding maximum permissible exposure time tor that noise
0.25 or less 1 15 level. The values of T, as specified by the Act, are given in Table
8.2. Note that there is a 5 dBA exchange in the values, implying that
when the level increases by 5 dBA, the maximum exposure time is
Mathematically, the above values may be halved. For instance, the maximum exposure time for 90 dBA is 8
expressed as: hours, the maximum exposure time for 95 dBA is 4 hours, and so
Maximum permissible exposure in dBA on3 .
= 105 - 1 6.6 log (T) Now consider a worker who is exposed to a noise level of 100
dB A for 1 hour, 90 dB A for 6 hours, and 80 dB A for the remaining 1
hour on a particular day. From Equation (8. 2), the worker's noise
dose for that day is:

1
D = - + __ + _!_ = 1 .25
00
2 8
which is unacceptable since it is greater than 1 .0. Note that T for 80
dBA exposure has been assumed to be infinite, since according to

3 The OSHA Act also provides that if an employee has a standard threshold
shift (STS), that is, if he/she has lost an average of 10 dB of hearing at 2.0, 3 .0
and 4.0 kHz in either ear, then the 90-dBA- 8-hour limit will be replaced by
85-dBA-8-hour limit. The other values in Table 8.2 will also be correspondingly
reduced. Thus, 95-dBA-4-hour limit will be reduced to 90-dBA-4-hour limit,
and so on.
170 Chapter 8

Table 8.2 the permissible exposure time for a noise level lower than
90 dBA is unlimited.
In case D exceeds 0.5, the employer must institute an action
plan for hearing conservation of the worker/s, as specified in the
Act. Employers usually enforce conformance with the Act through
a noise dosimeter, which the workers wear during work time, Figure
8.4. A noise dosimeter continuously records and displays the value
of D.

8.4 A worker wearing a noise dosimeter. Photo


courtesy of CEL Instruments, Milford, New
Hampshire, with permission.

8.3 INTERIOR NOISE The simplest and the most efficient means of controlling interior
noise is through architectural design. Rooms in which noise level
CONTROL THROUGH is expected to be high should be separated from noise-sensitive
ARCHITECTURAL rooms. Thus, in a multistory apartment building, bedrooms and
study rooms should be separated from lobbies, corridors and general
DESIGN circulation areas, Figure 8.5(a). The separation between noisy and
noise-sensitive spaces should be examined both in plan as well as
sections, Figure 8.5(b).
Noise Control in Buildings 171

Noise-sensitive spaces such as auditoriums, assembly halls,


concert halls etc., should be surrounded by ancillary spaces such as
lobbies, foyers, toilets, etc., in order to isolate them from exterior
noise. Mechanical equipment rooms should be separated from rooms
requiring quiet. In as much as possible, the building should be zoned
into noisy and quiet zones.
Open spaces and courtyards should be used, whenever possible,
to separate different or similar occupancies to provide acoustical
isolation, Figure 8.6.


u KIT
Ill I
KIT I--
BED
- =-
BED

J u u
c
J [
LIVING LIVING Alrl:oorne
BED BED sound

(a)
,I (11)
r
8.5 (a) Grouping of relatively noisy and quiet roo ms in a multifamily residential building. (b) Section through
a multifamily re sidential building indicating that structure-borne sound can transmit to a dwelling unit below
as well as to a unit to the side.

Court
Court

8.6 Roof plan of a courtyard building showing that an


111111111111 11111111111

open space can work as a good buffer between a noisy


and a quiet space, or between two occupancie s.
172 Chapter S

8.4 INTERIOR NOISE Another means of controlling interior noise is through the use of
sound absorbing materials. Since sound absorbing materials act by
CONTROL THROUGH reducing the intensity of reflected sound, they are effective in
SOUND ABSORPTIVE reducing reverberant sound only. The addition of sound absorption
has no effect on the level of direct sound.
TREATMENT It can be shown that noise reduction achieved through the use of
sound absorbing materials is given by:

A aft
NR = 10 log ---'- ( 8.3)
Abet

where, NR = noise reduction, Aaft = total room absorption after the


addition of sound absorbing materials, and Abef = total room
absorption before the addition of sound absorbing materials.
As explained in Example 8.3, treating the room with sound
absorbing materials to reduce interior noise level is effective only if
,_.:-,.,.j'-"'}0.:-:.Jf:'
-
'*' ",..';;.-:;., A.:-:"'" the room is relatively "live", i.e., if it does not have much sound
-...-:
-=.
-..:-
::- ----- -=- - --=-
..__
.: _

Ceiling aueorptlon absorption already. In such a case, a reduction of 9 to 10 dB may be


achieved through absorptive treatment. In fact, 10 dB is usually the
upper limit of reduction possible through sound absorptive treatment.
If sufficient absorption is already present in the room, noise
reduction obtained by treating the room is small - usually 2 to 4
dB. Although 2 to 4 dB seems a small improvement, it may be
worthwhile if the noise levels are high. Remember from Table 1 .3
8.7 In a room with a low ceiling, ceiling that a 3 dB reduction is a perceptible reduction.
absorption is effective since it is close
to noise source/s.
8.4.1 Placing Absorption Close to the Source

Note that since absorptive treatment reduces only reverberant sound,


it is beneficial to workers or occupants who are away from the source.
It does not help a worker who is close to the source since he/she gets
......
most of the noise as direct sound4. However, placing absorption
Space -- close to the source/s reduces the reverberant sound level by a greater
al7eorl7er amount than that indicated by Equation 8.3, since the sound gets
"killed" (absorbed) before becoming a part of the reverberant field.
In a small room, sound absorption may be placed on both walls
and the ceiling. In a large room with a low ceiling, the ceiling is
perhaps the best location for sound absorption, since it is the only
surface close to sound source/s, Figure 8.7. In a large room with a
8.8 In a room with a high ceiling, space
high ceiling, space absorbers are commonly recommended, since
absorbers bring absorption close to the they can be hung from the ceiling and brought close to the source/s,
noise source/s. Figure 8.8 (see also Figure 4.8).

4 Close to the source, it is the direct sound that is dominant (free field condition).
Away from the source, it is the reverberant sound that dominates (reverberant field);
see Figure 1.16.
Noise Control in Buildings 173

Example 8.3 Noise Reduction Through Sound Absorptive Treatment

A rectangular room measuring 20 m x 1 0 m x 6 m (high) with highly reflective surfaces (await& = 0.05,
afl r = 0.03, ac il = 0.03) is found to be quite noisy, Figure A. Determine the noise reduction 1f: (i) its
qp is treateo with a sound absorptive material ac l = 0.65) in the first phase, and (ii) if the upper
ceumg ( \li
part of walls, 4 m high, is also treated with sound absoromg material (awaits = 0.57) in the second phase.

Solution: First phase, Figure B:

Abef = Swalls (<lwans) + Sfloor (<lfloor) + Sc eil (<lcei l) =


360(0.05) + 200(0.03) + 200(0.03) = 30 sabins.
a = 0.03
Aaft = Swalls (<lwalls) + Sfloor (<lfloor) + Sceil (Ucen) =
360(0.05) + 200(0.03) + 200(0.65) = 1 54 sabins.
a = 0.05 = 0.05 a = 0.05
From Equation (8.3), NR = 10 log ( 1 54/30) = 7 dB (l

a = 0.03

FIGURE A INITIAL CON DITION

a = 0.65

a = 0.05
a = 0.05
a = 0.05

Second phase, Figure C: a = 0.03

Abef = 1 54 sabins FIGURE B


FIRST PHASE
TREATM ENT
Swalts(<lwans) = 240(0.57) + 1 20(0.05) = 1 43 sabins. Hence,

Aaft = Swa!ts (Uwatts) + Sfloor (<lfloor) + Sceil (<lceit) = 143 + 200(0.03) + 200(0.65) = 279 sabins.

From Equation (8.3), NR = 10 log (27911 54) = 2.6 dB, say 3 dB.

Thus, after the second phase, a total noise reduction of


(7 3) = 10 dB is obtained.
+
a = 0.65

The amount of absorption added in the first phase is


( 1 54 - 30) = 124 sabins. The amount of absorption added
in the second phase is (279 - 1 54) = 1 25 sabins. Although a = 0.57
both amounts are (almost) equal, the noise reduction in the
second phased is much smaller than the first phase. If we
could somehow add another 1 25 sabins in this room a = 0.03
(although practically impossible), we would find from a = 0.05
Equation 8.3 that noise reduction this time is a mere 1 . 6
dB . This i s an important conclusion, which states that SECOND PHASE
reducing interior noise through the use of sound absorptive FIGURE C TREATMENT

treatment follows the law of diminishing return. If there is


already adequate sound absorption in the room, adding some
more will give a small benefit in terms of noise reduction.
174 Chapter 8

A particularly effective means of noise control in a large


manufacturing facility is to use space absorbers coupled with free
standing partial-height sound absorbing barriers close to the sources,
Figure 8.9, provided that the barriers do not interfere with the
manufacturing process. In fact, partial-height sound absorbing
barriers are commonly used in open-plan offices to produce speech
privacy (refer to Chapter 14).

Celling
euepended
. / al:leorl:ler
Sound ""-

5ource In the middle


of the room, I.e.,
radiating In all 8.9 Partial-height sound absorbing barriers and space absorbers.
5ource near a room
directlon5. Q 1 =
5urface, I.e

radiating in half
apace. Q 2
8.4.2 Location of Noise Source in a Room
=

Room eound In a room with reflective walls, it is important to keep noise sources
level = X dB away from the walls. Theoretically, a noise source near a reflective
wall increases the noise level by 3 dB, as compared to a source in
the center of the room. A noise source placed near the edge of a
room increases the level by 6 dB , and a source placed in the corner
5ource near a of a room increases the level by 9 dB8.61.
room'5 edge, I.e., Therefore, in a manufacturing facility, the machines should be
radiating In quarter placed away from the walls, if possible. Thus, the arrangement of
epace, Q 4 =
5ource near a machines in Figure 8 . 1 0(a) is acoustically better than that of Figure
room'e corner, I.e., 8 . 1 0(b).
radiating In one
eighth 5pace, Q 8 =

Room eound level L J


= (X + 6) dB


Room eound level EJ
= (X + 9) dB
(a) D D D D D D
. Reflective wall
Q le called the dlrectlvfty of eourc;e (eee aleo r I]
Section 9.6 and Appendix K).

0 O D D
Machine /


()

JDD"
8.10 Two alternative arrangements of noisy
machines in a room. Less noise will be 0
produced in the room using alternative (a).
Noise Control in Buildings 175

8.5 INTERIOR NOISE A sound insulating (full-height) barrier between a noisy environment
and the receiving room is the most effective means of interior noise
CONTROL THROUGH control. The principles and practical details of sound insulating
BARRIERS construction have already been discussed in Chapters 5 to 7. In this
section, we will discuss the magnitude of sound insulation required
of barriers between occupancies.
In a highly critical occupancy, a detailed analysis of the
transmission loss of the barrier between the source room and the
receiving room must be performed as shown later in Example 8.4.
For normal occupancies, the required sound insulation of the barrier
may be obtained from the empirical data. Table 8.3 gives minimum
STC values for party walls between multifamily dwellings, and Table
8.4 gives recommended STC values for nonresidential occupancies.

Table 8.3 Minimum STC for Party Walls and Floors


Between Multifamily Residential Buildings

HUD Recommendations [ 8. 5 ] Building code requirements

Type of dwelling units


Barrier
Grade I Grade 11 Grade ill

Partition wall STC ;;:: 55 STC ;;:: 52 STC ;;:: 48 STC ;;:: 50

Floor ceiling assembly STC :2: 55 STC :2: 52 STC ;;:: 48 STC ;;:: 50

ne ;;:: 55 ne ;;:: 52 ne ;;:: 48 ne ;;:: 50

The U.S. Department of Housing and Urban Development (HUD) classifies multifamily dwellings into Grade I, 11, and Ill.
Grade I dwellings are those located in suburban or peripheral suburban areas, considered as "quiet" locations, with approximately
35-40 dBA, or lower, nighttime exterior noise levels.

Grade 11 dwellings are those located in urban and suburban areas considered to have "average" exterior noise environment, with
nighttime exterior noise levels of about 40-45 dB A. Grade Ill dwellings are those located in noisy urban areas, with nighttime
exterior noise levels of about 55 dB A or higher.

The building code requirements given in this table are from the Uniform Building Code, 1997. The same values appear in the
draft document of the International Building Code, expected to replace all three U.S. model building codes in the year 2000.
176 Chapter 8

Table 8.4 R ecommended STC Values for Selected Occupancies

Occupancy Room under consideration Adjacent room Recommended STC

Schools Classroom Classroom 45


Laboratory 50
Corridor or lobby 50
Kitchen or dining room 50
Shops 55
Music room 55 +
Mechanical room 60
Toilets 50

Large music room or Adjacent music room 65 +


theater Corridor or lobby 60
Practice room 60
Toilet 60
Mechanical room 60

Executive offices, Office Adjacent office 50


doctors' suites, where General office 45
privacy is important Corridor or lobby 50
Toilets 55
Manufacturing area 55
Mechanical room 60

Rooms for group Conference room Other conference room 50


meetings Adjacent office 50
General office 50
Corridor or lobby 50
Toilets 55
Mechanical room 60

Normal business General office area Corridor and lobby 45


offices Manufacturing area 45
Kitchen or dining 45
Mechanical room 60

Hotels and motels Bedrooms Adjacent bedroom or


living room or bathroom
of different occupancy 55
Corridor or lobby 55
Mechanical room 60

Concert halls, recording studios, TV stations, etc. These are extremely critical
areas. Carry out a detailed
noise analysis, or use values
from Table 6. 1 .
Noise Control in Buildings 177

8.5.1 Analysis of a Barrier's Transmission Loss Requirement

If a critically sensitive space is to be separated from a noisy space


through a barrier, such as a classroom separated from a transformer
room or a noisy manufacturing facility, the required transmission
loss of the barrier should be obtained by analyzing the noise reduction
at each octave from 1 25 Hz to 4 kHz.
Although the noise reduction between two spaces is primarily a
function of the transmission loss of the barrier, it also depends on
the area of the barrier and the amount of sound absorption present in
the receiving room. The smaller the area of the barrier and the greater
the absorption in the receiving room5, the greater the noise reduction.
It can be shown that noise reduction (NR) between two spaces is
given by:

A rec
NR = TL + 10 log (8.4)
s

where, TL is the transmission loss of the barrier, S is its surface area,


and e is the total absorption in the receiving room. If we know
the leve1 of noise reduction required, we can determine the required
transmission loss of the barrier by using the following equation.

Arec
TL = NR - 1 0 log (8.5)
s

For example, let the noise level in the source room at some
frequency be 90 dB . If the receiving room can tolerate a maximum
noise level of only 35 dB at that frequency, the required NR is 55
dB . If the area of the barrier between the source room and receiving
room is 1 5 m2, and the total absorption in the receiving room is 45
sabins, then from Equation (8.5), the required TL of barrier is:
RECEIVING SOURCE
ROOM ROOM
TL = 55 - 10 log(45/1 5) = 50 dB

Nole;e: tranemlt
ted from eou rce: Nolee: le:ve:l In 8.5.2 Required Noise Reduction
room

f3.ackeround The magnitude of noise reduction required between the source room
le:ve:l In and receiving room is a function of the background noise level in the
receiving room. The transmitted noise level should be lower than
the background noise level - generally by at least 5 dB. This ensures
Min. 5 dB
that the sum of transmitted and background noises will not be
significantly higher than the original background noise level, Figure
8. 1 1 .
Tranemltte:d nolee: le:ve:l muet be: be:low
backeround nolee: level in recelvlne
room (minimum 5 dB be:low)

8 . 1 1 Ba ckground noise level in


tceiving room and noise transmitted
source room.
5 To be conservative, for listeners who are close to the transmitting surface (barrier),
no credit should be taken for the receiving room absorption.
178 Chapter 8

Example 8.4 Design for the Transmission Loss of a Barrier

A conference room measuring 10 m x 15 m x 3.5 m (high) is separated from a noisy manufacturing


facility by a barrier, Figure A . The conference room has an acoustical ceiling and a carpet on the floor.
The three walls of the room are of gypsum board on steel studs, and the other wall has fixed glazing.
The absorption coefficients of the room surfaces are:

Frequency 1 25 250 500 1 ,000 2,000 4,000

Ceiling 0.55 0.89 0.73 1 .00 1 .00 1 . 00


Floor carpet 0.08 0.24 0.57 0.69 0.7 1 0.73
Glazing 0. 1 5 0.05 0.04 0.03 0.02 0.02
Walls 0.26 0. 1 0 0.05 0.03 0.04 0.04

The HVAC noise environment specified for the conference room is RC-3 5 . Expected (or measured)
noise environment values in the manufacturing room are :

Frequency 1 25 250 500 1 ,000 2,000 4,000


Manufacturing room
noise 70 78 82 88 86 82

Solution: The barrier surface area = 1 0 x 3.5 = 3 5 m2. We now determine the total absorption in the
conference room for each octave, by multiplying the area of the element with its absorption coefficient.
For instance absorption provided by the ceiling at 500 Hz = 1 50(0.73) = 109.5 sabins. Detailed
calculations of total absorption in the conference room are given below.

Frequency 1 25 250 500 1 ,000 2,000 4,000

Ceiling (150 m2) 82.5 1 33.5 109.5 1 50.0 1 50.0 150.0


2
Floor ( 1 50 m ) 1 2.0 36.0 85.5 1 03.5 1 06.5 1 09.5
2 0. 7
Glazing (35 m ) 5.3 1.8 1 .4 1.1 0.7
2
Walls ( 1 40 m ) 36.4 14.0 7.0 4.2 5.6 5.6

Total absorption (A) 1 36.2 1 85 . 3 203.4 258.8 262.8 265.8 15 m

-
+ i

Since the specified HVAC environment for the conference room is RC- or' Barrier
1.!)
35, the noise after going through the barrier must be at least 5 dB below z 10 m
RC-35 environment, i.e. RC-30. Thus, the required noise reduction in ii2 CONFERENCE
::I ROOM
1-
each octave is given below. (.) Glazlne -
<
u..
::I
-:01
Frequency 1 25 250 500 1 ,000 2,000 4,000 z
<
::t
FIGURE A
Manufacturing room noise 70 78 82 88 86 82
Noise level in receiving room
after transmission (RC-30) 15 20 25 30 35 40

Required NR 55 58 57 58 51 42
Subtract 1 0 log (ArecfS) -6 -7 - 8 -9 -9 -9

Required TL 49 51 49 49 42 33

The barrier must be such that its transmission loss at each frequency must be equal to or greater than the
required TL.
Noise Control in Buildings 179

8.6 EXTERIOR NOISE The most c ommon sources of exterior noise, also termed
environmental noise (or community noise) are traffic noise, aircraft
CRITERIA noise, industrial noise, etc. How do we rate the annoyance potential
of environmental noise? The NC curves and RC curves used for
rating interior HVAC noise cannot be used to rate environmental
noise, because of one fundamental difference between the two types
of noise: the HVAC noise level is constant, while the environmental
noise varies with time.
Because of its temporal variation, environmental noise does not
lend itself to a simple rating criterion. Consequently, an extremely
large number of criteria have been suggested over the years. Although
efforts to arrive at a consensus criterion have yielded some results,
there are still a large number of them in use. In this text, we will
discuss only two commonly used criteria: Le and L A
.

comprehensive coverage may be found in texts deoted sofeiy to


environmental noise[S. 71.

8.6.1 L
eq

It has been found that the annoyance caused by environmental noise


is related to its total sound energy. Therefore, a descriptor that is
most commonly used is Leq - an abbreviation for equivalent steady
(continuous) sound level.
Le is defined as that sound pressure level which, if constant
over agiven period, will contain the same sound energy as the actual
sound that is fluctuating with time over that period. Although any
period tnay be used, Leq measurements are usually made over 1 hour
or 24 hour period; these are denoted by Leg(1) or Leq(24) respectively.
Since Leq is a sound pressure level, 1t can be measured in dB or
dBA. Since dBA measurement correlates better with subjective
perception, Leq(l) or Leq(24) is usually given in dBA.

8.6.2 Ldn

Ldn an abbreviation of day-night equivalent sound level, is Leq24)


in which the sound levels recorded during 1 0 PM to 7 AM are raised
by 1 0 dB over their actual values to account for the greater annoyance
caused by the same sound at night. The Federal Aviation
Administration in the United States considers L better measure
of the annoyance caused by aircraft noise than L . 1herefore, aircraft
e.g
noise contours around an airport are usually mapped in terms of Ldn.
Both Le and Ld are measured by integrating sound level meters,
q n,
which are availaole from several manufacturers. Airport Ldn is
usually computer generated, based on flight pattern.

8.6.3 Noise Impact and Land Use Compatibility

In the United States, cities and several federal agencies, such as the
Department of Housing and Urban Development (HUD), Federal
Aviation Administration (FAA), Federal Transit Administration
(FTA), Federal Highway Administration (FHWA), etc., are involved
with environmental noise issues. The have formulated ordinances
and guidelines to assess the impact of environmental noise.
180 Chapter 8

For instance, a noise impact study of a proposed highway (or


major alteration to an existing highway) on communities affecte4
by the proposal must be carried out where FHWA funds are used. If
the highway-produced noise levels are "excessive", noise abatement
elements, such as noise barriers, earth embankments, or both must
be constructed along with the highway. The definition of "excessive"
is given in FHWA's noise abatement criteria[S.SJ.
Similarly, HUD has established guidelines for environmental
noise assessment and control for HUD-financed housing. HUD's
Noise Regulation, published in July 1979, established an upper limit
on L0n of 45 dB A for interior noise level. Although this establishes
intenor level, the exterior acceptable noise level is obtained bj
assuming that the overall transmission loss of a convention
dwelling's envelope (walls, windows, doors, roof, skylight, etc.) to
be nearly 20 dB.
Table 8.5 HUD's Noise Standards for Thus, 65 dB A is considered to be an acceptable Ldsl value for a


New Housing new residential development. Table 8.5 gives HUD's nmse standards
for evaluating residential sites for financing new housinj
Site Classification Ldn (dBA)
construction[S. 9] .
Thus, for sites that are in the normally acceptable category, w
an Ldn lying between 65 to 70 dBA, the dwelling's envelope mu
have an overall TL of 25 dB (5 dB more than a conventional dwellin
Acceptable 65 which is assumed to give nearly 20 dB A reduction for transportati
Normally acceptable 65 - 75 noise). If Ld is between 70 to 75 dBA, an additional l O dB in
Unacceptable > 75 overall TL of the envelope must be provided, i.e., an overall TL of
30 dB.

8.7 EXTERIOR NOISE Most cities have noise ordinances specifying the maximum noise
levels that can be produced by industries and other noise-producinJ
CONTROL THROUGH occupancies at their property lines. In addition to the noise legislatioJt
SITE PLANNING the following site and town planning principles can be used to control
exterior noise.
Increase distance between the noise source and the received
Heavy foliage coupled with several rows of trees reduce!
noise, Figure 8. 12(a). Although the trees do not absorb much
sound, they diffuse sound so that a part falls on the foliageJ
where it gets absorbed. Heavy foliage absorbs sound to thd
same degree as an interior carpet. One or two rows of trees
with no or little ground foliage will not attenuate any more
sound than that attenuated due to distance, Figure 8. 12(b).
Trees should be evergreen, not deciduous that shed their
leaves during autumn.
Self-protecting building forms can shield noise-sensitiv
parts of the building from the noise source. A few self
protecting building forms are shown in Figure 8. 13.
Buildings housing noise-sensitive spaces should be laid
perpendicular to the street, shielded by buildings that can
tolerate noise, Figure 8.14.
Noise Control in Buildings 181

(a)

(17)

8.12 (a) Several rows of trees with heavy foliage provide attenuation of
sound in addition to that obtained due to the distance effect.
(b) One or two rows of trees with little or no ground foliage give no
additional sound attenuation except that due to the distance.

8.13 A few self-protecting building forms that shield parts of the building sensitive to noise.

Nolee
eeneltlve
epace&

[),______....(]
Street

8.14 O rientation of buildings


containing noise-sensitive spaces.
182 Chapter 8


If tall buildings have overhanging balconies facing a busy
street, the underside of their balconies should be treated with
sound absorbing material to absorb sound before it hits the
building facade, Figure 8 . 1 5 .

Residential districts and other areas where quiet is needed
should be separated from industrial districts, highways,
railways, airports, etc.
Road network should be planned in such a way that traffic is
concentrated on a few streets rather than being distributed
on several of them (see Example 1 .5).

Sound a17eorl71ng
rnatrlal
J

8.15 Sound reflection


from balcony soffits.

8.8 EXTERIOR NOISE When site planning measures do not provide the desired noise
reduction, embankments (berms) or barrier walls or both should be
CONTROL THROUGH built, Figure 8 . 1 6. Barrier walls, or simply barriers, are usually
BARRIERS made of masonry, concrete, metal panels, etc.
As shown in Figure 8. 17, noise attenuation provided by a barrier
is a function of the following three factors:
Sound diffracted over the barrier,
Sound reflected by the ground and diffracted over the barrier,
and
Sound transmitted through the barrier.
The sound transmitted through a commonly used outdoor barrier
is usually much smaller than that diffracted over the barrier, provided
the barrier is without holes or gaps. Therefore, the factors that
determine sound attenuation of a barrier are diffraction and ground
absorption.
Noise Control in Buildings 1 83

Noiee 17arrier
Berm

8.16 Berms and noise barriers.


Note: Trees that are close to a sound barrier wall
should not be higher than the wall, since they will
scatter sound and allow it to penetrate the shadow
zone, compromising the effectiveness of the wall.
For the same reason, trees or shrubs should not be
planted on the top of a berm.
Barrier Diffracted eound
Reflected and
diffracted eound

Source "- --- -- - -- ----.,' - ----


- - - - - - -- - -- n- - - - - - ---- - )
- - - - - - 0 Receiver
Tranemitted
eound
8.17 Sound propagation between a source
and a receiver across a barrier.

The terms commonly used for the noise reduction given by a


barrier are:
Barrier attenuation, and
Barrier insertion loss
Barrier attenuation is defined as the sound pressure level
reduction provided by the barrier under free-field conditions. Hence,
barrier attenuation does not include ground characteristics.
Barrier insertion loss takes into account the modification
introduced by ground absorption on both sides of the barrier. Barrier
insertion loss is, therefore, defined as the difference in sound pressure
levels with and without the barrier, but in the presence of the ground
in both cases. Both barrier attenuation and barrier insertion loss
refer to noise reduction in excess ofthat occurring due to the distance
effect.
1 84 Chapter 8

8.8.1 Semi-infinite Barrier and Point Source

Consider an infinitely long barrier of a finite height (hence the term


semi-infinite) and a point source, Figure 8. 1 8. Barrier attenuatiod
(A) is given by the following approximate relationshiP[8 . 1 0, 8 . 1 1 ].

A = 1 0 log (20 N + 3) (8.6)

Source
where
N = (20 /A), and o = (a + b - d),
Bsrrfer d = length of straight line path between the source
and the receiver, called the line of sight,
(a + b) = shortest path length between the source and
8.18 Barrier and source-receiver receiver over the barrier's edge, and
geometry. A. = wavelength of sound.

8.8.2 Line Source and Semi-infinite Plane Barrier

If the source is a line source (such as highway noise), the attenuation


provided by the barrier is lower than that given by Equation (8.6).
Attenuation by a line source is given by:

A = 10 log (20 N + 3) - (20 N)03 (8.7)

where, N is to be measured in a plane perpendicular to the barrier


0
and passing through the receiver.
.......
-4 '\.
5PL - 8 "' ....... 8.8.3 Semi-infinite Plane Highway Noise Barrier
(.::t B r---r-- Since traffic noise levels are usually expressed in dBA, it is

- 12
convenient to express barrier attenuation in dBA. The dBA values
- 16
of barrier attenuation can be obtained by including the effect of the
spectral distribution of traffic noise and applying A-weightinj
63 250 1,000 4,000 function to attenuation values obtained from Equation (8.7).
Freuency (Hz) The highway noise spectrum is a complicated function of mean
traffic speed, road gradient, type of road surface, composition of
vehicles, etc. However, a typical highway noise spectrum may be
8.19 Approximate spectral composition assumed as shown in Figure 8 . 1 9. With that assumption, the
of highway noise. Lowfrequency noise attenuation of a semi-infinite highway barrier is as shown in Figure
8.20.
is due to automobile engines and the The curve of Figure 8.20 flattens at nearly 1 8 dBA, implying
higherfrequencies are due to tire noise. that to be the maximum achievable attenuation. In practice, the upper
limit of attenuation of a semi-infinite plane highway noise barrier is
lower - nearly 15 dB A- due to ground and meteorological effects.
Example 8.5 illustrates the use of Figure 8.20.
Noise Control in Buildings 185

20

<
ltl ---
15 ------
V

/
1:::
0
43
<1'1
:::3
1:::

<1'1 10

1/
L.
"
1:
L.
<1'1
ltl
5
0 0.5 1.0 1.5 2.0 2.5 3.0

o = (a + 1.> - d) In m

8.20 Attenuation provided by a semi-infinite highway noise barrier as a


function of path length difference.

8.8.4 Finite-length Plane Highway Noise Barrier

In practice, barriers must be of finite length. With the finiteness of


barriers, diffraction of sound occurs at the sides of the barrier, which
reduces barrier attenuation. Attenuation values of finite length
barriers have been developed in the form of graphs[8 . 1 2] . The
following expression may be used as first approximation to obtain
the attenuation provided by a finite-length barrier.

(X

a Receiver Aa = A l SO (8.8)

Barrier
where, Aa = barrier attenuation subtending an angle of a (in degrees)
at the receiver, and A is the attenuation of the same barrier if it were
infinitely long (i.e., a = 1 8o).
186 Chapter 8

Example 8.5 Noise Reduction Due to a Highway Noise Barrier Wall

Determine the noise level at a receiver located behind a long (consider infinitely long) highway noise
barrier with a total height of 4.0 m above the road level, as shown in Figure A. Assume that highway
noise level measured at 20 m from the center of the highway (and at a height of 1 .0 m above road level)
is 85 dBA.
Solution: a = 40. 1 1 m, b = 20.22 m, d = 60 m. Hence, 0 = 40. 1 1 + 20.22 - 60 = 0.33 m. From Figure
8.20, A = 1 2.6 dBA. This is the barrier insertion loss (attenuation in excess of the attenuation due to
distance). From Section 1 .6.3, traffic noise reduces by 3 dB per doubling of distance. Hence at the
location of the receiver (60 m from the center of the highway), noise reduction due to distance is
approximately 4.8 dB. Thus, the total noise reduction (attenuation) = 12.6 + 4.8 = 1 7.4 dBA, say 1 7
dBA.
The above calculation does not include the effect of the ground, which degrades the attenuation.
Usually 3 to 4 dB A degradation due to ground may be expected. Thus, the actual noise reduction will
be nearly 1 3 to 14 dBA, in place of 17 dBA.

3m
C enterllne of
highway
i...::::...----ff- 0
Receiver 1 m

20 m 20 m 20 m
FIGURE A
Noise Control in Buildings 1 87

REFERENCES 8.1 Cowan, James: Handbook of Environmental Acoustics, Van Nostrand


Reinhold, New York, 1 994, p. 204.
8.2 Beranek, Leo: Noise and Vibration Control, Institute of Noise Control
Engineering, Washington D.C., 1988, p. 560.
8.3 American Society of Heating, Refrigerating and Air-Conditioning
Engineers, Atlanta: ASHRAE Handbook ofFundamentals, 1993, Chapter
7.
8.4 American Society of Heating, Refrigerating and Air-Conditioning
Engineers, Atlanta: ASHRAE Handbook ofFundamentals, 1 993, Chapter
7.
8.5 U.S. 92nd Congress: Noise Control Act of 1972, Public Law 92-574,
HR 1 1 021, October 27, 1 972.
8.6 Beranek, Leo: Noise and Vibration Control, Institute of Noise Control
Engineering, Washington D.C., 1 988, p. 231.
8.7 U . S . Department of Housing and Urban Development (HUD),
Washington D.C: The Noise Guide Book, HUD-953-CPD, 1985.
8.8 U.S. Department of Transportation, Federal Highway Administration
(FHWA), Office of Environment and Planning, Washington D.C:
Summary of State Highway Agency Noise Planning Definitions, 1 99 1 .
8.9 U . S . Department of Housing and Urban Development (HUD),
Washington D.C: The Noise Guide Book, HUD-953-CPD, 1 985.
8.10 Moreland , J. B., and Musa, R. S . : "The Performance of Acoustic
Barriers", International Conference on Noise Control Engineering,
Washington D.C., 1 972, pp. 95-104.
8.11 Tatge, R. B.: "Noise Reduction of Barrier Walls", Arden House
Conference, 1 972.
8.12 Mehta, M.: "The Potential for Traffic Noise Reduction by Thin Masonry
Panels", American Society for Testing and Materials (ASTM) Publication
STP 1 063, Masonry: Components to Assemblages, 1 990.
Control of HVAC
Noise

Noise emitted by a heating, ventilating and air conditioning (HVAC)


system is present in almost all modern buildings. The acceptable
HVAC noise levels (NC and RC curves) were described in Chapter
8. In this chapter, the means of achieving the goals established by
the above criteria are discussed.
Although acousticians play only a minor role in the design of an
HVAC system, they are routinely required to review HVAC design
drawings to ensure that the established noise criteria will be met in
the completed building. Calculations of expected noise levels,
particularly in critical spaces, are required at the design stage, and if
the calculations reveal excess noise levels, changes in duct layout,
duct lining, addition of duct silencers etc., are included.

9.1 HVAC SYSTEMS The HVAC systems commonly used in buildip.gs may be classified
as: window air conditioners, fan coil units, roof-top units, packaged
air handling units, and built-up air handling units.

9.1.1 Window Air Conditioner

A window (or through-wall) air conditioning unit copsists of an


integral compressor and a fan assembled together in a box. The unit
uses the refrigerant as the only working fluid, and requires no duct
work. The noise from the unit cannot usually be controlled through
building design, but is a function of the unit's design. Obviously, a
unit that produces a low noise level is preferable. However, the
broad-band masking noise generated by a window air conditioner
may assist in masking exterior noise or noise from adjacent spaces.

189
190 Chapter9

9.1.2 Fan Coil Unit

A freestanding individual fan coil unit, which involves heating or


cooling of mainly the recirculated air, is commonly used in hotel
rooms - one fan coil unit for each room. The unit is placed near an
outer wall (to draw some make-up fresh air if necessary), but may
also be placed in the ceiling void of a hotel corridor. The unit consists
of a fan blowing air over chilled water coils, where the chilled water
is supplied from a central plant . Chilled water coils are similar to
the radiator in a car, consisting of a series of loops of copper piping
with narrow fins past which the air moves.
For heating, hot water or electric resistance heating is used. The
unit uses no duct work, and the noise from the unit is simply that
from the fan and that radiated due to the vibrations of the panels
constituting the unit enclosure.

9.1.3 Roof-top Unit

A roof-top unit is typically a complete system incorporating an


internal compressor and evaporative coils to cool a working fluid,
heat exchanger coils to extract heat from circulating air, filters, and
supply and return fans. Heating is usually accomplished through
electric resistance heating. Duct work is required, as shown in Figure
9.1. The unit is commonly used for single story buildings in which
the ceiling plenum space serves as the return air duct. Noise control
in the system follows the same general principles as those for the
larger systems described in Sections 9.1.4 and 9.1.5.

Roof-top HVAC unit

Exhaue;t

9.1 Roof-top air conditioning unit with distribution duct work.


Control of H VAC Noise 191

9.1.4 Packaged Air Handling Unit

A packaged air handling unit ( AH U) is a factory-prefabricated


"packaged" unit located to serve one or more floors of a building
through connected duct work, Figure 9.2. An AH U typically mixes
fresh air (also referred to as make-up air) with return air, filters it,
cools (or heats) it, adjusts its humidity, and blows it via a supply fan
through the distribution ducts. Vibration isolation of of the fan-motor
assembly may be internal to the unit, or isolation mounts may be
located on the exterior of the unit. Sound attenuators may be integral
to the discharge plenum of the unit.
Chilled water is obtained from a central plant, utilizing a system
of pumps to circulate water between an exterior heat exchanger
(cooling tower or air-cooled chiller) and an interior chiller, which
uses the cooled fluid to extract heat from the water circulating
between the chiller and the air handling unit's coils.
When electrical resistance heating is not used, either steam or
boiler water may be used to add heat to the air via hot water circulating
through a coil in the air stream.
An AH U may have both hot and cold water coils operating
simultaneously for extremely accurate temperature/humidity control.
The most common configuration is the variable air volume ( VAV)
system, which distributes cooled air from an AHU having a variable
speed fan to a series of mixing boxes. A mixing box is so called
because it mixes warmed building air (usually from the above-ceiling
plenum) with freshened, cooled air from the AH U. Each mixing
box is thermostatically controlled.
The AH U's supply fan's speed (hence, the air's volume) is
regulated by the total demand for cool air from the mixing boxes.
Mixing boxes come in a variety of of configurations, including simple
induction boxes and powered induction boxes, with small internal
electric fans.

9.2 Pack aged AHU. Courtesy


ofIndustrial Acoustic Company,
Bronx, New York, with
permission.
192 Chapter9

9.1.5 Built-up Air Handling Unit

A built-up air handling unit is similar to a packaged AHU, except


that it is larger, and hence, its various components are assembled at
the building site. A large built-up AHU may use much of an entire
building floor. It consists of a bank of air filters and coils. It may
have one or more supply fans, which discharge the air through a
silencer bank into a supply plenum and from there to supply ducts,
Figure 9.3.
Very large centrifugal or axial fans, mounted on inertia bases,
are used. If two or more fans discharge into a common chamber,
backdraft dampers may be required to isolate a fan that is not
operating, so that cooled air does not return through the inoperative
fan. This also permits maintenance on the inoperative fan. Sound
pressure levels nearly 2 m away from such large fans exceed 105
dBA. Therefore location of an AHU is critical. Typically, a built-up
AHU is located in the basement or in the penthouse of a building. If
located in the penthouse, the AHU must be supported on a floating
floor as shown in Figure 9.3. A floating floor is usually unnecessary
for a ground-supportedAHU unless there is an acoustically sensitive
space close to the AHU - to dampen ground-carried vibrations.

. :o. "'
o .. . ;.'o .,. ' 41' .0
. o,:c-o
.. . .
..
.
"

. . . .
: .0 6
. . ;
. ':' '0.:
.

0, . , . o. .
.

) m uuoouoouuuQlll\(Y\ J(Wl
j
,.

Sound abeorl:>lng lining


l .
'
Main eupply AHU encloe;ure to provide

l
duct high tranemle;e;lon loee and
high thermal lne;ulatlon D
l 1
__r=D.
001
Dlecharge
plenum 2
--
1:::

- ., - Dr''

"'
pl"m 1

t--- -
._

.!!
iii

.>::
2 :'1
:
I

u:
-----

-.!!-
iii

._
.,
<.l
1:::

.,
Freeh air and
retum-alr In-
take

Floating floor
l
c

Id._
t'----)0
-
Jr
I
jt.o!:;"'
r--- -
.;.? 0 . .
. . .":o .. . .. . "' : . .. . .
,i)
.
.. .
u.,:
. ... .
.
. .. .-t:,;_ . :
. I>
. o: ., :o
, . .. , .
C> _o
.. : o ., _r .. .

9.3 A typical built-up AHU.


Control of HVAC Noise 193

9.2 NOISE The primary source of noise in an HVAC system is the fan. The
noise from the fan is fed into the ducts. As the fan's sound travels
ATTENUATION IN along the duct system, its level decreases through several energy
DUCTS extracting mechanisms. Some of the sound energy is converted to
vibrations of duct walls. The duct radiates this vibrational energy in
the space surrounding the duct as sound, known as duct break-out
noise, Figure 9.4.
The sound energy that is retained in the duct is absorbed by the
internal fiberglass lining (if used), duct bends, junctions, and duct
cross-sectional changes.

9.4 Duct break-out noise.

9.2.1 Duct Lining

To increase its attenuation, as the sound travels in the duct, the duct
is internally treated with fiberglass, referred to as duct lining, Figure
duct 9.5. The use of lined ducts is therefore common in HVAC systems,
since an unlined duct provides negligible attenuation except in the
low frequency region, Figure 9.6. This attenuation is mainly due to
9.5 Fiberglass lining in HVAC ducts. duct break -out.

I I I
1.0 \..
150 mm x 150 mm ( 6 In. x 6 In.) -

duct

300 mm x 300 mm ( 12 In. x 12 -
. .... ::. : :
....
:
In.) duct
300 mm x 600 mm ( 12 In. x 24 - E:
0


... . . . . . . . . .......... . . . .
0.5 0.15

. . . . . . . ... .
E:

.. - . .. . ...... . .

0 ---------L--L--- 0
9.6 Sound attenuation of 63 125 250 500 1.000 2,000 4,000
unlined ducts. Adapted Frettuenc;y (Hz)
from Reference 9.1.
194 Chapter9

Duct lining consists of fiberglass made of bonded fibers with a


neoprene-coated facing to prevent the erosion of glass fibers 1.
Generally, 25 to 50 mm ( 1 to 2 in.) thick lining is used, with a density
of nearly 25 to 50 kg/m3 ( 1 .5 to 3 lb/ft3). The larger thickness provides
greater attenuation. The increased attenuation is mainly in low and
mid frequency regions, with virtually no increase in high frequencies
( 1 kHz and above), Figure 9.7.
While the decrease in sound energy in an unlined duct is
expressed as attenuation, the corresponding decrease in a lined duct
is given in terms of insertion loss. Both attenuation and insertion
loss are measured in dB/meter (or dB/ft). The insertion loss is simply
the additional sound reduction provided by fiberglass lining - over
and above that provided by an unlined duct of the same cross-sectional
dimensions. Thus, the attenuation provided by a lined duct of given
dimensions is simply the sum of its insertion loss plus the attenuation
provided by an unlined duct of the same dimensions.

16 4.9

14 4.:3
... .......... ........... ..
12 ....... :3.6

I
/ ......
'E
iO
50 mm
10 t- llnlne
(2 In.) thick
-r " :3.0
/
""

I
10

,:/

Ill
8 2.4

I 1-
Ill
s::
0
6
i
25
llnlne
mm (1 In.) thick
- 1.8
s::



.!:
4
. . ..
...
...
.



/ 1.2
0)

V
..
.!:

... .
. ............... ...
...

2 -- 0 .6
9.7 Insertion loss of300 mm
x 600 mm (12 in. x 24 in.) 0 0
sheet metal duct with 25 mm 6:3 125 250 500 1,000 2,000 4,000
(1 in.) thick and 50 mm (2 in.) Frequency (Hz)
thick fiberglass lining.
Adapted from Reference 9.1.

1 Recently, the use of fiberglass lining in ducts and plenums has been banned or
severely limited in several educational and medical projects due to the concern
that glass fibers promote microbial growth and may be carcinogenic. Evidence
available so far has failed to show any evidence to endorse this concern (see
Reference 9. 1, page 43.32).
Some manufacturers offer duct lining treated with bactericidal and fungicidal
materials to reduce the contamination effect. In a duct system where no lining is
permitted, system cost increase for the control of noise is inevitable. The increased
cost may be due to increased attenuator lengths, specification of attenuators without
fibrous lining, or with fibrous lining enclosed in thin airtight polyethylene bags, or
the use of computer-controlled active silencers.
Control of H VAC Noise 195

The insertion loss of a duct generally decreases with increasing


duct size, as illustrated in Figure 9.8. For a more comprehensive
data of duct attenuation or insertion loss, the reader is referred to
ASHRAE tables[9.1J .
Note that because of the disparity in low and high frequency
attenuation rates, an internally-lined duct system may exhibit an
unbalance (rumbly) noise spectrum. It may be advisable to
discontinue internal lining beyond a point where a sufficient amount
of high frequency attenuation has been obtained. Duct silencers,
discussed later, suffer from the same problem, i.e., they too are poor
at the low frequency end.
Since duct transitions - change of duct cross-section, elbows,
and duct terminations (diffusers) - also attenuate sound. ASHRAE
tables should be consulted to obtain sound attenuation provided by
such transitions.

16
-------- ----,-- 1 ----,-
-,--- 1 ./ .. . . . . . . . .
.-- ... . --,---------, 9
- . 4

14
,-------,-

:300 :300 --+-+------"..---


--<(
... . .
"-1 -
.
- -----i .
4 :3
(12 In. 12 In.) duct j
mm x mm
\ ..... .

12 :3 6
. .

.
x ..

/ ---+ .... .. --1


.

E' I I -+-/-#---! """"' ----' --....,---... ---- -",-- .


10 r------ :300 600 \ .,,. :3.0
f ......
/(1
iO .
.


(12 In. 24 In.) duct / . .................. """--
mm x mm

x
\


..9 8 I / "- 2 . 4 111

, . / ....... !
1:
0
e 6 .
1.

'""'
600 mm x 600 mm
1:

4 . ->"
..-..-.. +-
...
+- (12 In. x 24 In.) duct - 1.2
-= _

..=
--- -+--_.-.
r-- '
.. :r"'-=-
"-:
2 . 0.6


... ........... ..::: : :..;::;


0 L--------L-- 0
..... _

6:3 125 250 500 1,000 2,000 4,000


Fre'\uency (Hz)

9.8 Insertion loss of three rectangular sheet metal ducts of different cross
sections - all with 25 mm (1 in.) thick fiberglass lining. Adapted from
Reference 9.1.
196 Chapter9

9.2.2 Sound Absorbing Plenums

Where noise attenuation greater than that provided by duct runs and
transitions is required, air plenums and duct silencers are used. A
plenum is routinely provided beyond the outlet section of a fan and
before the main distribution duct to smooth out turbulent airflow in
the vicinity of the fan. A plenum is a large enclosure (usually o
sheet metal), whose walls are lined with fiberglass, Figure 9.9. Tho
surface area of a plenum must at least be 1 0 times the inlet area.
The greater the surface area of a plenum in relation to the inlet area,
the greater its effectiveness.
To achieve a gooq absorption over the entire frequency range, a
1 00 to 150 mm (4 to 6 in.) thick fiberglass lining is required. In
addition, the inlet and outlet openings should be staggered as much
as possible both in the horizontal as well as in the vertical direction.
The effectiveness of a plenum is also expressed as insertion loss.
The insertion loss is simply the additional sound pressure level
reduction obtained by the use of a plenum. Insertion loss of nearly
1 2 dB is obtainable in low frequency region from the use of a lined
plenum.

Inlet duct

9.9 Sound absorbing plenum.

plenum enclosure

9.2.3 Duct Silencers

A duct silencer (also called sound trap, sound attenuator, or simply


a muffler) has a cross-section equal to or greater than that of a duct
in which it is installed. A duct silencer divides the air flow into two
or more fiberglass-lined openings, Figure 9 . 1 0. Sound traps are
available in prefabricated sizes and lengths - 1 m (3 ft) to 3 m ( 1 0
ft) long. The attenuation provided by a sound trap i s also expressed
in terms of insertion loss. Manufacturer ' s data is the best source for
9.10 A typical duct silencer. obtaining the silencer ' s insertion loss.
Control ofHVAC Noise 197

9.2.4 Branching of Ducts


Branch
duct: area On reaching a branch, the sound energy divides itself into two parts.

J
= 0.5A__.J
__
Branch This division is in direct proportion to the individual branch areas.
duct: area Thus, if the cross-sectional area of branch duct is Ab and the area of
= 0.5A
the main duct is Am the reduction in the sound power level in the
,

branch duct (as compared to that in the main duct) is given by the
following expression.
.
area= A Reduction in sound power level = 10 log
A
A
:
9.11 Branching ofducts and noise level For instance, if the main duct divides into two ducts of the same
distribution. area, that is the ratio of the area of branch duct to that of the main
duct is 0.5, the noise in each branch duct is 3 dB less than in the
main duct, since 10 log (0.5) =- 3 dB, Figure 9.11.

9.3 NOISE GENERATED The air flow in ducts should be as smooth as possible in order to
reduce the generation of air turbulence. In an improperly designed
BY AIRFLOW duct system, turbulence increases noise levels, and also decreases
the HVAC system's efficiency. An important factor that affects
turbulence is the velocity of air in the duct.
ASHRAE specifies the maximum air velocity for ducts, which
is a function of the NC (or RC) rating of the space, and whether the
duct is circular or rectangular in cross-section. The greater the NC
(or RC) rating, the greater the maximum air velocity permitted, Table
9.1. A circular duct gives less turbulence. Therefore, a greater air
velocity is permitted in a circular duct.

Table 9.1 Recommended Maximum Air Velocities in Rectangular Ducts

Main duct location Design RC(N) Maximum air flow velocity


mlmin (ft/min)

In shaft or above drywall ceiling 45 1070 (3500)


35 760 (2500)
25 520 (1700)

Above suspended ceiling 45 760 (2500)


35 530 (1750)
25 370 (1200)

Duct located within occupied space 45 610 (2000)


35 440 (1450)
25 290 (950)

Adapted fro m Reference 9.1.


198 Chapter9

Another factor that influences air turbulence in a duct is the shape


of duct transitions - changes in duct cross-section, bends (elbows)
and T-junctions. Smooth transitions are commonly used since they
produce less turbulence than sharp transitions. Thus, a radial elbow
is preferable to a square elbow. In square elbows and right angle
branch take-offs, turning vanes are usually specified to reduce
turbulence, Figure 9 . 1 2.

El"owe in ducte
L ....
....
\..._'- Vanee
L L
Avoid Uee Uee

Air

Right angle "ranch


_jTL Air flow

take-off
Avoid Uee \: VanHJ to direct air
Into 17ranch take-off

'
_)w
:;::moth
n 1 re

Angular "ranch take


off Airflow

Thle anele not to

I
mcceed15
Change in duct Airflow Airflow
croee-eection
Avoid Uee

9.12 Duct transitions to reduce noise generated by air flow.


Control of HVAC Noise 199

9.3.1 Diffusers and Grilles as Noise Generators

Terminal units in ducts that supply and distribute air into the room
are called diffusers. In a return air system, the terminal units remove
air from the room, and are usually referred to as grilles or registers.
Diffusers and grilles generate noise, as they constrict the passage of
air through them. This is unlike the in-duct elements, which attenuate
noise. Diffusers and grilles are therefore noise additive, and because
they are terminal elements, there is no way to attenuate their energy.
Manufacturers of these devices provide sound power level data of
their products.
Diffuser noise tends towards being "hissy", i.e., in mid and high
frequency range, due to the small scale of the turbulent eddies
generated by vanes. Therefore we avoid face dampers to control air
volume, since they increase turbulent flow. Locate volume control
devices several equivalent diameters upstream of diffusers, so that
by the time air reaches the diffusers it is relatively free of turbulence.

9.4 SOUND RA DIATION Duct break-out noise has already been referred to in Section 9.2. It
must be accounted for in calculating the noise in a room. Thus, the
BY DUCT WALL S noise in a room is due to that received from duct outlets and duct
break-out noise.
Duct break-out is more serious in a high velocity system.
However, it is present to some degree in a low velocity systems as
well. Duct break-out can be quite high in the first 6 to 10 m (20 to
30 ft) from the fan. Therefore, the initial part of the duct should not
traverse an acoustically sensitive space. In other words, an important
space, such as a conference room or a board room, should not be
located close to an AHU's main (trunk) duct.
ASHRAE[911 provides a simple procedure to determine duct
break-out noise. If the calculated break -out level is excessive, means
must be employed to reduce it. The simplest means of reducing
duct break-out is to use a circular duct, which because of its inherent
stiffness, is less prone to break-out than a rectangular duct, Figure
9.13. However, increased duct wall thickness (gauge), a duct width
to-height ratio close to 1.0, and the short distance between bracing
flanges improve break-out characteristics of rectangular ducts.

9.13 Break-out noise is lower in a


round duct than in a rectangular duct. nolee
200 Chapter9

Another means of reducing duct break-out is to spot bond


fiberglass boards to the duct exterior and then wrap it with loaded
vinyl, Figure 9.14. In place of loaded vinyl, gypsum boards may be
used (see Section 5.5 for a description of loaded vinyl).
The duct treatment described above is referred to as lagging.
Lagging is particularly important in a duct that enters a quiet room
after passing through a noisy space. Even if the noisy space has no
duct opening, the noise from a noisy space will enter the duct through
its walls and travel along the duct length. This noise will be delivered
into the quiet room through the duct opening and also as break-out
noise, Figure 9.15. This phenomenon, known as duct break-in, is
evidently more pronounced in an unlagged duct.

Sh metal
duct Tape

9.14 A lagged rectangular duct.

.olt, . 6 -
. .. . - . . c - . .. - . .. . ..
..
. . r; . . . ' . . . ', : .
I .o. : ' : .

Ently of nolee In the duct U NCONDITIONED


Duct break-out QUIET CONDITIONED
(Duct 17reak-ln) NOISY SPACE
SPACE

9.15 Duct break-in from a noisy space into a relatively quiet space.
Control of HVAC Noise 201

An important thing to remember here is that the sound travels


equally well along or against the direction of the air flow in a duct.
This is because the velocity with which the sound travels (344 rn/s,
i.e., 1130 ft/s) is much greater than the velocity of air flow in the
duct- usually between 6 to 15 rn/s (20 to 50 ft/s). Thus, if the
velocity of air flow is 6 rn/s, the sound will travel at the rate of 350
rn/s in the direction of air flow and at the rate of 338 rn/s against the
direction of air flow- a negligible difference. Thus, duct break-in
noise will travel equally well in either direction.

9.5 CROSS Cross transmission (or cross talk) is the transmission of noise between
adjacent rooms through a supply or return duct that is common to
TRANSMISSION IN the two rooms, Figure 9.16. A careful examination of cross
DUCTS transmission between rooms is necessary where speech privacy or
noise isolation is important. In general, duct attenuation should be
at least 5 dB more than the TL of the partition between rooms. Thus,
if the TL of the partition is 40 dB, the attenuation of the duct should
be at least 45 dB.

0 s. .
Q

====l[r l-,--::::=:...
iran!lmleelon
through duct
V
iranemleelon
through wall

9.16 The transmission o f sound


prrough a duct should be lower than that
'hrough the wal l dividing the t wo
spaces.

Several means are available to decrease cross transmission.


Indeed, lining the duct is necessary but since the duct length between
adjacent rooms is usually small, the attenuation obtained from duct
lining alone is inadequate. Therefore, these other means are usually
employed to increase attenuation:

Use duct layouts that increase the sound paths (along the
ducts) between rooms, and add one or two elbows in the
supply duct. Thus, the duct layout of Figure 9.17(b) is
preferable to the duct layout of Figure 9.17(a).
If the layout of Figure 9.17(a) is used, a duct silencer should
be added in both the supply and return air ducts. This is
commonly employed to meet privacy requirements in areas
classified as secure compartmented information facilities
(SCIF), where classified information must be controlled.
202 Chapter 9

ROOM ROOM ROOM

Supply duct ---. ! I

Return-air duct __.

-
0 0 0
'----

(a) Duct layout pron to croe;e tranemleelon

Supply duct ---.


r--
Return-air duct __.
I--

c:;-;:::
ROOM ROOM ROOM

JJI UI _lji L...-


.....____

-
g [ g
(1:1) Duct layout to reduce croee tranemleelon

9.17 Cross transmission in duct systems.

9.6 ESTIMATIN G HVAC ASHRAE has developed a straightforward method to estimate the
sound pressure levels in each octave from 63Hz to 8 kHz at a receiver
NOISE LEVELS in a space served byHVAC duct system2. A typical work sheet for a
room is shown in Table 9 .2. We start with the sound power levels of
the fan(s) at the air handling unit- data that is routinely available
from fan and air handler manufacturers.
Proceeding from the fan, we tabulate each element in the duct
system that affects the sound power level in the air stream until the
air is delivered to the served space. Elements include lengths of
straight duct, elbows, sound attenuators, plenums, branches and VAV
mixing boxes. This data is typically available from ASHRAE tables,
except for plenums, which must be calculated, and for attenuators,
whose data is available from manufacturers.

2 Many consultants skip the 8 kHz octave because sound absorption coefficients
are not reported at that frequency, and HVAC noise is seldom a problem at that
frequency.
Control ofHVAC Noise 203

The tabulation continues in terms of sound power level up to the


diffuser. The sound power supplied by the diffuser is converted to
sound pressure level at the receiver from the following formula.

S PL = PWL + 10 log [ 41tr


Q
2 + _!_ ]
R
(9 .1)

where, SPL =sound pressure level in room at a distance r from the


diffuser (in meters),
PWL =sound power level of diffuser,
Q =directivity factor of source (diffuser). Q = 1, if the diffuser
is suspended free in space (radiating in all directions); Q = 2, if the
diffuser is located flush with the ceiling or a wall (i.e., radiating into
half space); Q = 4, if diffuser is located on an edge between two
planes (radiating into quarter space); Q = 8, if the diffuser is in a
corner (radiating into one-eighth space); see also S ection 8.4.2.
R =room constant (in metric sabins), which is equal to:

R=

where r.A =total absorption in room (in metric sabins), and aav =
average absorption coefficient of the room; see Equations (4.3) and
(4.4).
If the customary U. S . system of units is used, in which r is
expressed in feet, and R in ft-sabins, Equation (9.1) becomes:

S PL = PWL + 10 log [. 4 Q1t + _!_] + 10 (9.1 )


'

. r2 R

In practice, we calculate (PWL - SPL) and use this value, known


as the room effect, in the table. For the value of r, select a receiver
location that represents the worst likely case.
If there are several diffusers serving the room, the contribution
from each must be determined at the receiver, and summed on an
energy basis (decibel addition). There can be other continuous
contributions present, which further increase the sound level at the
listener - fan noise entering via the return air path, radiated noise
from mixing boxes above the ceiling, break-out noise from sheet
metal ducts, and sometimes noise transmitted through the wall or
floor or roof separating the receiver from adjacent mechanical
equipment. These, too, are calculated and summed at the receiver.
The resulting octave spectrum at the receiver is assigned an RC
(or NC) value. This value is compared with that desired for this
receiver. If the rating is too high, investigate the likely causes and
make adjustments until the criterion is satisfied. This latter analysis
requires experience hearing and judging noise problems in the field,
knowing when the calculation procedure might be presenting
misleading results, and seeing if the problem frequencies are high,
mid or low, in order to insert appropriate correction measures.

i
204 Chapter9

Table 9.2 Typical Worksheet to Estimate HVAC Noise Levels in a Room

63 125 250 500 1K 2K 4K Element Source

88 88 90 87 82 77 73 Fan Sound power data - 7340 cfm Manufacturer


-1 -1 -1 0 0 0 0 Duct, rectangular, unlined, 30 in. x 24 in. x 5 ft long ASHRAE
0 -1 -4 -6 -4 -4 -4 Elbow, unlined rectangular, turning vanes, 24 in. wide ASHRAE
-1 -1 -1 0 0 0 0 Duct, rectangular, unlined, 30 in. x 24 in. x 5 ft long ASHRAE
-1 -4 -6 -6 -4 -4 -4 Elbow, rectangular, unlined, turning vanes, 30 in. wide ASHRAE
0 0 0 0 0 0 0 Branch take-off ratio 0.925 ASHRAE
-2 -2 -1 0 0 0 0 Duct, rectangular, unlined, 30 in. x 24 in. x 10ft long ASHRAE
0 0 0 0 0 0 0 Branch take-off ratio 0.90 ASHRAE
0 0 0 0 -1 -1 -1 Duct, round, unlined, 30 in. diameter, I 0 ft long ASHRAE
-2 -2 -2 -2 -2 -2 -2 Branch take-off ratio 0.60 ASHRAE
0 0 0 0 0 0 0 Duct, round, unlined, 18 in. diameter, 4 ft long ASHRAE
-3 -5 -10 -15 -17 -16 -9 Duct, flexible, lined, 12 in. diameter, 5 ft long ASHRAE
-7 -8 -17 -21 -14 -13 -9 3 XL silencer Manufacturer

71 64 48 37 40 37 44 Total sound power level - obtained by adding all above values


59 48 41 34 30 27 25 Sound power level generated by diffuser Manufacturer

71 64 49 39 40 37 44 Sound power level in room - dB addition- of above two set of values

-15 -16 -16 -16 -16 -16 -16 (PWL - SPL) from Eq. (9.1'), Q = 2, r = 15 ft- room effect- calculated

56 48 33 23 24 21 28 NC30 RC23(H) TARGET 30 or less

9.7 ACTIVE NOISE Reducing or completely eliminating a sound through another sound,
by destructive interference, is called active noise control, as opposed
CONTROL IN HVAC to the use of fiberglass, which is referred to as passive noise control.
SYSTEMS Although active noise control has been known for a long time- the
first patent was awarded in 1936l921 it became a practical
-

possibility in 1980 when high speed digital signal processors became


available at low cost.
The principle of active noise control is fairly simple. It consists
of producing a secondary sound that is identical to the primary sound
in frequency and level, but is 180 out of phase. Being 180 out of
phase ensures that the pressure maximum of the secondary sound
occurs at the same time and location as the pressure minimum of the
primary source, so that the two sounds add up to a zero pressure,
Figure 9.18.
Theoretically, it should be possible to apply active noise control
to any noisy environment. However, at the present time, active noise
control is limited to reducing low frequency noise in HVAC ducts
apart from its use with several nonbuilding noises. With future
advances in digital signal processing, we may be able to successfully
use active noise control in factories, airport lounges, and other similar
noisy environments.
Control ofHVAC Noise 205

Dletance

Wave 1

9.18 Pressure-distance relationships in two waves that Wave 2


are 1800 (a halfwavelength) out of phase.

Figure 9.19 shows an active noise control system applied to an


HVAC duct. It consists of an input microphone located at the
beginning of the duct. The signal received by the microphone is fed
into a digital controller, which is essentially a computer. The
controller determines the frequency of sound, its amplitude and phase
and finally produces an inverted signal- 1 80 out of phase with the
noise in the duct. The output of the controller is fed into a loudspeaker
through an amplifier, which produces the canceling sound.
The loudspeaker is placed some distance away from the input
microphone, so that as soon as the noise from the duct arrives at the
loudspeaker, it is canceled. An error microphone monitors the
residual noise and provides feedback to the controller to optimize
the system's performance.
The distance between the input microphone and the loudspeaker
is an important factor, since the computer (in the controller) takes a
finite time to digitize and send the inverted signal to the loudspeaker.
Obviously this time must be shorter than the time it takes for the
noise to travel from the microphone to that of the loudspeaker. In
other words, the loudspeaker must be sufficiently far off from the
microphone to provide adequate time for the computer to perform
its functions.

9.19 Pr inciples of active noise


control in ducts. Controller
206 Chapter9

The system works well if the sound to be canceled is a plane


wave. In a plane wave, the sound pressure and phase are constant
on a plane perpendicular to the direction in which the wave travels.
The sound waves in a duct behave as plane waves only if the
wavelength of sound is large as compared with duct dimensions -
at least twice the largest dimension of the duct. Thus, active noise
control is suitable to attenuate frequencies below 250 Hz. This is
rather fortunate since passive means of noise control (duct lining,
plenums and attenuators) are relatively ineffective at low frequencies.
A major difference between passive and active duct silencing is
that the passive duct silencing reduces duct area, which induces a
pressure drop in the air stream. To overcome this loss, the pressure
at the supply fan has to be increased, which increases energy
consumption of the system. By contrast, the energy required for
active noise control system is extremely small - seldom more than
40 watts. Additionally, a passive system introduces resistance to air
flow in the duct, which must be overcome by increasing the air flow
of the fan - increasing the energy consumption at the fan.

REFERENCES 9.1 American Society of Heating, Refrigerating and Air-Conditioning


Engineers, Atlanta: Heating, Ventilating and Air-Conditioninl
Applications, 1995, Chapter 43.
9.2 Gelin, Lawrence: "Active Noise Control: A Tutorial for HVAC
Designers", ASHRAE Journal, August 1997, p. 44.
The Behavior of
Sound in Rooms

One does not need to be an acoustician to appreciate the difference


between the sound produced inside an enclosed space and that
produced outdoors. The primary reason for the difference is that the
sound produced inside a room bounces back and forth from room
surfaces, while a sound produced outdoors travels freely away from
the source. Indeed the characteristics of a room (room geometry,
volume, and the absorption characteristics of its surfaces and
contents) greatly influence both the sound quality as well as its level.
In this chapter, we will examine the behavior of sound in rooms,
a topic known as room acoustics. We will deal with the phenomenon
of the decay of sound in a room and its effect on speech
communication and the perception of music. The next three chapters
are devoted respectively to the acoustical design of rooms meant for
speech, music, and multipurpose spaces.

10.1 IMPULSE If we produce a short "burst" of sound (sound impulse), such as that
generated by pricking an inflated balloon or by a hand clap in a large
RESPONSE OF A ROOM room, we will observe that the sound does not die instantaneously.
The sound of the balloon or the clap persists for a while, decreasing
in level over time. In fact, in a large hall with highly reflective walls,
the persistence of sound after its termination is quite noticeable. In
a reverberation chamber (Figure 3.28), which is specially designed
to have a long decay time, the sound may persist for as long as 10
seconds or more. Some of the old cathedrals in Europe have decay
times close to that value. On the other hand, in an open area or in an
anechoic chamber the sound dies instantaneously.

207
208 Chapter 10

10.1.1 The Reverberation Phenomenon

The persistence of sound in a room after it is turned off is related to


the amount of absorption in the room. In fact, as soon as a sound is
produced, it travels in space in various directions and hits room
surfaces, from which it is reflected and rereflected. At each reflection,
some energy is lost by absorption, and eventually all the sound is
depleted. This is shown in Figure 10.1, in which the path of one
sound ray (out of an infinite number of rays produced by the source)
has been followed over a few reflections.
The phenomenon of successive reflections by room surfaces is
better grasped by considering the source and its mirror images.
Assume a rectangular room and consider once again a sound impulse
generated by a source, S, Figure 10.2. As soon as the source is
energized, images of the source are created by reflections from
various room surfaces.
By following the procedure described in Section 3.9, we note
10.1 The decay of sound intensity in a
that an infinite number of images of the source is established. These
room as represented by the paths of one are referred to as the first-order images, second-order images, third
ray over several reflections. order images, etc. With reference to Figure 10.2, we see that image
I1 is a first-order image and has been produced by reflection from
wall 1. This image, in turn, produces three second-order images I12 ,
I13 and I14 -by reflections from walls 2, 3 and 4 respectively.
Since there are four first-order images (Il' I 2., I3, and I4), a total
of twelve second-order images are produced. However, in a
rectangular room, some second-order images fall at the same location.
For instance, images [1 and I coincide with each other.
2 2
Consequently, only eight second-orer images are seen, although in
reality there are twelve of them.
In fact, the location of various images for a rectangular room is
fairly simple, since each image is contained in an imaginary replica
of the actual room. Only the first- and second-order images and a
few third-order images have been shown. Higher-order images, and
the images produced in the third dimension - by the floor and the
ceiling of the room - are not shown. Thus, the number of images
and their distribution in space is far more complex than Figure 10.2
indicates.
Note that all images of the source come into action at the same
time as the source itself. Additionally, there is no fundamental
difference between the sound received from the source and that
received from any one of its images (the reflected sound), except
that the higher-order images are progressively weaker because 1 of
the partial absorption of sound energy at each reflection . In
recognition of this fact, the second-order images have been depicted
by lighter dots than the first-order images, and the third-order images
by plain dots. For the sake of clarity, the third-order images have
not been designated.

1 Some selective absorption will also occur, as some frequencies will be more
absorbed than others at boundaries and by the air in the room- changing sound's
"timbre".
The Behavior of Sound in Rooms 209

142
q p ,P
\ \ I I
\ I I
\ I
\ I
\
\ I
\
\ I
\ I
I
\ I
\ I I
\ \ I I
\ \ I I

0..------------
1:31

'

/
', '
'
I '

61
'
'o

10.2 Mirror images of the source in a rectangular room and the directions of sound received by a listener. Only
the first-order, second-order and some third-order images are shown - in one plane.
Note that one and only one image lies in each
replica of the actual room. A line and arrow in
the diagram indicate the direction of travel of a
sound wave from the source or its image, as it 10.1.2 Directional Distribution of Reflections
is incident on the listener, L. It does not imply
that sound travels like the "rays" of light; see Now consider the sound received by a listener situated at L. The
Section 10.7. listener receives sound impulses from the source as well as from all
its images. Since the listener is closer to the source than any one of
its images, the first sound impulse to arrive at the listener is from the
source - along direction SL. This impulse is referred to as the
direct sound.
The subsequent impulses come from the images. The directions
from which these impulses arrive at the listener are shown by broken
lines in Figure 10.2. Notice the spread in the directional distribution
of the sound arriving at the listener, implying a good deal of diffusion,
which becomes more obvious if higher-order images are taken into
account.
The directional distribution of sound is particularly important in
halls meant for music. Sound coming from many different directions
creates a sense of "volume" or "envelopment" in the room - an
important requirement for the appreciation of music. More
specifically, as we shall see in Chapter 12, strong early reflections
coming from the lateral directions (as opposed to the overhead
reflections) are particularly favored for concert halls.
210 Chapter 10

10.1.3 Impulse Diagram


Direct eound
Rewerl:lerant While Figure 10.2 gives the directional and spatial distribution, Figure
.---- eound 10.3 gives the temporal distribution of sound, referred to as an impulse
diagram. In an impulse diagram, the vertical axis represents the
sound level and the horizontal axis represents the times of arrival of
impulses. Observe that, in general, each impulse is of a lower level
than the one preceding it, showing a gradual decrease in sound level
5PL over time. This progressive decrease is a consequence of two factors:
(dB) (i) increasingly higher-order images are weaker in power, and (ii)
they are farther away from the listener.
The impulses in a typical impulse diagram are so numerous and
so closely spaced that the listener does not perceive them as distinct
Time
sounds. Instead the listener's perception of this phenomenon, known
Time 0, i.e., the time when
=
as reverberation, is one of an integrated, smooth and continuous
the eource le enerelzed
decay of sound, as shown by the broken line in Figure 10.3.
In practice, therefore, the decay of sound pressure level in a
10.3 Sequence of direct and reflected room is represented by a continuous line, not by an impulse diagram.
The slope of this line gives the rate of sound decay. The greater the
sounds gen erated by an impulsive slope, the faster the decay of sound.
source.

1 0.2 IMPUL SE A smooth decay of sound is obtained only if the sound field in the
room is diffuse. By definition, a perfectly diffuse sound field is one
DI A GR A M A N D SOUND in which the sound reaches every listener in the room in equal strength
DIFFUSION from all directions (see Section 3.7). The impulse diagram in a room
with a perfectly diffuse field is perfectly uniform, implying that
individual impulses decrease uniformly in level with no deviation
from the imaginary broken line, Figure 10.4.
A perfectly diffuse sound field is an ideal condition, seldom
achieved in practice. However, a reasonably diffuse field is obtained
in a room whose average absorption coefficient a is small, and
where the absorption is (almost) uniformly distributed on all surfaces.
As the nonuniformity in the distribution of absorption increases,
fluctuations from the uniform decay increase. For instance, if one
or more surfaces of the room are highly absorptive, significant lack
of diffusion results.
This is shown in Figure 10.5, where one of the four walls (wall
2) of the room is fully absorptive (i.e., a = 1.0). If we construct
mirror images of the source for this room, we will see that there are
no images beyond wall 2, since there are no reflections from this
wall. In fact, the images for this room are located only within a
narrow spatial band. Consequently there is considerable non
uniformity in the directional disposition of sound in this room, as
compared with the room of Figure 10.2. This difference is more
apparent if third- and higher-order images are considered.
5PL
(dB)
10.2.1 Reverberation and Echoes

Lack of diffusion also results if there is focusing of sound, such as


that produced by a large concave surface, e.g., a domical or vaulted
Time ceiling, Figure 10.6(a), or a curved rear wall of an auditorium, Figure
10.6(b). Since focusing concentrates energy within a narrow beam,
other directions are starved of energy, reducing diffusion in the room.
10.4 Impulse diagram of a room with
a perfectly diffuse field.
The Behavior of Sound in Rooms 211

Wall 2

10.5 First- and second-order images in a room in which one wall (wall 2) is fully absorptive (a = 1.0).
Observe the nonuniformity of directional distribution of sound at the listener as compared with the room of
Figure 10.2.

The impulse diagram of such a space is nonuniform because


some reflections become clustered together in time to produce one
or more reflections of a relatively high level, as shown in Figure
10.7. Since our hearing has a limited ability to integrate successive
reflections, the outstanding reflections are perceived as echoes,
particularly if there is a large time gap between the outstanding
reflection and the direct sound (see Section 2.3).
Thus, the difference between reverberation and echoes is fairly
(a) SECTION THROUGH A HALL WITH subtle. Reverberation is a smooth decrease in the energy content of
VAULTED OR DOMICAL CEILING successive reflections, so that the reflections are not individually
perceptible. An echo, on the other hand, consists of one or more
relatively strong reflections within the general reverberation process,
and is heard separately.
An additional difference between reverberation and echo is that
reverberation can be a useful phenomenon which reinforces the direct
sound. An echo, on the other hand, is annoying and creates an
0 unfavorable acoustics. Flutter echo, discussed in Chapter 3, is a
5 special type of echo, in which the relatively strong reflections are
repeated periodically.

(11) PLAN OF A HALL WITH CURVED REAR


WALL
Echo

10.6 A large domical or vaulted


ceiling, or a curved rear wall of a hall
produce focusing of sound.
5PL

'
'
'
'

10.7 Impulse diagram of a room


Time
with sound-focusing elements.
212 Chapter 10

10.2.2 Buildup Versus Decay of Sound

If, instead of an impulsive sound, a sound with a constant level is


produced in a room for some time and later terminated, the level1
time graph of that sound will be represented by Figure 10.8. This
5PL diagram shows that, like the decay process, the buildup of sound is
also gradual, not abrupt. In other words, it takes a finite time for the
sound to build up to a constant level.
However, the rate of buildup of sound is much steeper than its
k- Time when the decay (and also curvilinear). Because of its steepness, the buildup is
eou rce ie
ewitc hed off
perceived by the ear as being instantaneous. The middle part of
Figure 10.8 represents the steady state - a period during which the
sound level is constant, until the source is switched off. The stead
10.8 Relative steepness of the buildup state is made up of the direct and reverberant sounds, and is useful id
and decay processes. determining the room' s effect on the sound levels produced by stead
sound sources, such as air conditioning or machines with constant
noise output.

10.3 REVERBERATION Except sounds created by air conditioning and a few other sources,
most sounds in buildings seldom reach the steady state. They are
TIME transient sounds, identified by start-stop characteristics. A typic
speech, for instance, consists of phrases, which in turn consist d
words. Each word consists of one or more syllables. Silent interval
(pauses) exist not only between words and phrases but also between
syllables. Music is also composed of transients. For example, an
intervening pause is necessary as one piano key is pressed after
another.
Due to the transient nature of most practical sounds, and because
sound decay is far more perceptible than buildup, it is the decay of
sound, not its buildup, that is important in room acoustics. Mord
precisely, it is the rate of decay of sound that affects the acoustics of
lecture and concert halls.
Although the importance of decay rate in room acoustics has
been known for centuries, it was first placed on a quantitative footind
nearly one hundred years ago by Wallace Clement Sabine, a professor
of physics at Harvard University. Sabine identified the decay rate as
reverberation time and defined it as the time taken by a sound to
decrease 60 dB from its value at termination.
90 Thus, if the sound level when the source is terminated is 90 dB,
I
I
and it takes 1.6 seconds for the sound level to decrease to 30 dB,
I
I then the reverberation time of the room is 1.6 seconds, as shown in
5PL I
I
I
60 Figure 10.9.
(d6) I
I
dB By extensive and ingenious measurements in halls of different
I types, Sabine obtained the following relationship betweea
I
I
30 -- -
; - --- - reverberation time and room parameters.
1

lo 1.0 2.0
Ti e (eecorule)

Reverl:leratlon
I RT =
0.16 V
A
(10. 1 )
i time f
where, V is room volume (in m3), and l:.A, the total absorption in the
10.9 Definition of reverberation time room (in metric sabins). The value of RT is in seconds.
- as the time it takes for the sound to Equation (10.1), known as the Sabine equation, shows that the
decay by 60 dB after its termination. reverberation time is a function of only two room parameters: room
The Behavior of Sound in Rooms 213

absorption and room volume. The dependence of reverberation time


on the total absorption in the room is quite obvious, since the greater
the absorption, the more rapid the depletion of sound energy. Room
volume comes into the picture because the greater the volume, the
less often the sound will strike room surfaces where it is absorbed,
and hence the slower the decay of sound, leading to a higher RT.
In fact, the (statistical) average distance between two successive
strikes of sound, as the sound bounces back and forth in a room, is
given by (4V /S), where S is the total surface area of the room, and V
its vdlume. This average distance is known as the mean free path,
and is an important concept in the mathematical derivation of the
Sabine equation (see Appendix G). Mean free path is also important
in understanding some of the factors that influence the acoustics of
music spaces.
The total absorption (:EA) in Equation (10.1) is given by the
following relationship:

where, S1, S S3, ... etc. are the surface areas of a room (in m2), and
' az, <l:3 2. etc., are the corresponding absorption coefficients.
. .

The term, mV, represents air absorption. As discussed in Section


4.11, air absorption is significant only at frequencies equal to or above
2 kHz. For frequencies beiow 2 kHz, the term, mV, is too small to
be considered. The values of m are given in Table 4.2.
If the volume of room is given in ft3 and the surface areas in ft2,
Equation (10.1) becomes:

0.05 V
RT = (10. 1')
:EA

Note that the choice of 60 dB decay2 in the Sabine equation is


entirely arbitrary, ad is embedded in the history of room acoustics.
At the time of Sabine' s investigations, modern sound measuring
equipment was not available. Sabine used organ pipes to produce
the sound, and his ears and a stop watch to measure the time for the
sound to become inaudible. In his experiments, 60 dB represented
the approximate difference between the sound level produced by the
organ pipe, and the level at which the sound became inaudible in the
prevailing background noise level.
The perception of speech or music transients in rooms is based
only on the first few decibels of the decay curve, not the entire 60
dB. Thus, reverberation time should be regarded merely as a measure
of the rate of decay of sound, particularly a comparative measure of
the rates of decay in one room versus the other.
)

2 Since it is usually difficult in practice to produce a noise level that is 60 dB


above the background noise, RT is obtained by measuring the time over 30 dB
decay and doubling it. The 30 dB decay is taken from -5 dB to -35 dB.
214 Chapter 10

10.3.1 Limitations of Sabine's Equation


Area of ceiling
= 36 m2 If we consider a room whose surfaces are completely absorptive (i.e.,
a = 1 .0 for all surfaces), as in an anechoic chamber, we should expect
a zero RT from Equation ( 10. 1 ). However, if we substitute the value
of a = a = a3 . . . = 1 .0 in this equation, we obtain a nonzero value
of RT. ?or example, let us calculate the reverberation time of an
anechoic chamber whose dimensions are 6 m x 6 m x 4 m, Figure
10. 10. Its volume, V = 6 x 6 x 4 = 144 m3 . When the values of S l'
1 0 . 1 0 An anechoic chamber, as S2, S , etc., and a = = a3 . . . = 1 .0 are substituted in Equation
..l
referred to in the text. ( 10. 1 ), we obtain KT = 0 14 sec, as shown below:
.

0. 1 6( 1 44) 0. 1 4 sec
RT = =

3 6( 1 .0) + 36( 1 .0) + 4(24)( 1 .0)

The above calculation shows that Sabine equation has limitations


in its application. In fact, as discussed in Appendix G, the Sabine
equation is based on the assumption that the sound field in the room
is diffuse. The sound field in an anechoic chamber, which is an
acoustically "dead" space, is far from diffuse, since there is only one
direction from which the sound can reach a listener - from the
source.
To have some degree of diffusion, a room must be acoustically
"live", i.e., its surfaces must be fairly reflective. A room is generally
considered live if the average absorption coefficient of its surfaces
is 0.2 or less. In other words, the Sabine equation applies only if aav
0.2. Thus, if all room surfaces are perfectly reflective (i.e . , aav
=
0), the Sabine equation predicts an RT of infinity, just as we would
expect it.
Most lecture rooms, auditoriums, concert halls, etc., belong to
the live or almost live category, for which the Sabine equation is
valid. For a room that is not live, or if the distribution of absorption
in the room is highly nonuniform, the Sabine equation must be used
with caution. For a relatively dead room, other reverberation time
equation s , such as the Eyring equation ( Appendix G), are
recommended.

10.4 SIGNIFICANCE OF It was previously mentioned that reverberation is a useful acoustical


phenomenon since it reinforces the direct sound. In reality, this is
REVERBERATION TIME true only for the early part of reverberation since it is the early part
of the reflected sound that our hearing mechanism integrates with
the direct sound. In Section 2.3, we observed that the integration
period is nearly 50 milliseconds for speech and 80 milliseconds for
music. If the reverberation time of a room is 1 .0 second, a duration
of 50 milliseconds represents a very small fraction of the decay
process. What then is the subjective effect of the later part of the
decaying sound?
The later part of reverberation is not entirely detrimental, since
it adds body and ambience to sound, similar to the contribution of
diffuse light in a photography studio. However, if the reverberation
time is too long, its negative effect is bound to be perceived. The
questions that arise are: (i) what is/are the negative effect/s of
excessive reverberation, and (ii) what is the optimum reverberation
time of a room?
The Behavior of Sound in Rooms 2 15

10.4.1 Reverberation Time and Speech Intelligibility

The simplest way to comprehend the negative effect of an excessively


long reverberation time is to examine it with respect to speech
intelligibility3 . The sound energy that does not reinforce the direct
sound, in effect, becomes a masking sound for speech, reducing
speech intelligibility. The detrimental effect of long reverberation
time on speech intelligibility is well known to anyone who has tried
to speak in a large stair hall, a gymnasium, or a highly reverberant
auditorium.
In a typical speech, as the sound of a syllable decays during the
pause, it tends to mask and obscure the sound of the subsequent
syllable. The masking effect is illustrated in Figure 10. 1 1 with respect
to a three-syllable word "musical" spoken in two rooms, one with
an RT of 2.0 sec, Figure l O. l l (a), and the other with an RT of 0.4
sec, Figure lO. l l (b).
Assuming that each syllable is uttered at the same level (70 dB),
we observe from Figure l O. l l (a) that the syllable "mu" has decayed
to a level of nearly 69 dB before the syllable "si"is spoken. Thus in
this case, the masking level (69 dB) is only 1 dB below the signal
level (70 dB). By comparison, the masking level in Figure l O. l l (b)
is nearly 64 dB. Thus, speech intelligibility will be higher in the
room with a 0.4 second RT, thaJ l the one with an RT of 2.0 sec.
The masking effect also explains why a speaker needs to speak
slowly, providing longer pauses, to increase intelligibility in a room
with a long reverberation time. The effect of slow speech is illustrated
in Figure l O. l l (c) with reference to the word "musical" uttered in a
room with an RT = 2.0. Comparing the profiles of Figures l O. l l (a)
and (c), we observe that longer pauses reduce the blurring of syllable
sequence, giving a clearer and more defined profile. Longer pauses,
however, make the speech appear unnatural, and create an undue
stress on the speaker.

70 ....... 70

5PL 65 5PL 65 -.;


5PL
(dB) .!i ]
60 (dB) 60 .... .!! (dB)
<:
:>< iij
Ill <:
55 55 <11 ....
m

0 0.5 0 0.5 0 0.5


Time (eec) Time (eec) Time (eec)

(a) RT = 2.0 5t::C (b) RT 0.4 51lC


= (c) RT 2.0 51lC
=

NORMAL SPEECH NORMAL SPEECH SLOW SPEECH

10.11 A simplified illustration of the masking effect of reverberation on the intelligibility of the three-syllable
word "musical". (a) RTofroom = 2. 0 sec, (b) RTofroom = 0.4 sec, and (c) word uttered slowly in a room with
RT = 2. 0 sec.

3 See Chapter 14 for a detailed discussion of speech intelligibility.


Chapter 10

The previous discussion implies that the shorter the reverberation


time, the more intelligible the speech. Taken to its logical conclusion,
the statement implies that the speech is most intelligible in a space
where there is no reverberation - an anechoic chamber, or outdoors.
That conclusion is, however, not true, since speech intelligibility
decreases markedly as the listener moves away from the speaker in
an open space. Reverberation increases the distance within which
speech communication is possible. In other words, the absence of,
or inadequate, reverberation restricts the size of room in which speech
communication with good intelligibility can take place.
Thus, the optimum reverberation time for a hall meant for speech
is a compromise between the loss of intelligibility due to excessive
reverberation, and a loss of sound level due to inadequate
reverberation. In addition, a low reverberation time produces sound
that appears "flat".

10.4.2 Reverberation Time and Music

What has been said about the effect of reverberation on speech is


generally valid for music also. Reverberation enhances musical tones.
It also tends to mask minor musical imperfections. As to the required
value of reverberation time, a short reverberation time gives
"definition" or "clarit y " to music, and an excessively long
reverberation time leads to "muddled" music.
However, the effect of reverberation on music cannot be
demonstrated as deterministically as that for speech. Speech
intelligibility in a room is a measurable quantity, and its dependence
on reverberation time is well established.
Music perception, on the other hand, is far more subjective.
Additionally, unlike speech halls whose acoustical quality is
determined primarily by speech intelligibility, the acoustical quality
of music halls depends on a number of descriptors, none of which
can be defined with the same precision as speech intelligibility.
As discussed in Chapter 12, only a few descriptors of the
acoustical quality of concert halls (liveness, fullness of tone and
warmth) are related to the hall's reverberation time. There is no
known objective relationship of these descriptors with reverberation
time. However, numerous studies of existing concert halls and other
music spaces lead us to the requisite conclusion as to the optimum
reverberation time for music spaces.

10.5 REVERBERATION Since the absorption coefficient of surface finishes and the audience
varies considerably with frequency, the reverberation time of a room
TIME CALCULATIONS is a function of frequency. Therefore, the reverberation time of a
room must be calculated at a number of frequencies.
The Behavior of Sound in Rooms 2 17

For a typical auditorium, reverberation time is calculated at three


representative frequencies: 1 25 Hz (representing the low frequency
region and denoted by RT 1 ), 500 Hz (representing the mid
frequency region and denoteazby RT5.o0), and 2 kHz (representing
the high frequency region and denoted by RT2 oool: The calculations
are usually done in a tabular format as shown'm hxamples 10. 1 and
10.2 at the end of the chapter.
For more critical spaces, such as a concert hall, a broadcasting
studio, recording room, etc., RT may be calculated at the center
frequency of each octave band ( 1 25 Hz to 4 kHz). Where RT is
quoted without reference to any frequency, it implies RT500.

10.5.1 Accuracy in Reverberation Time Calculations

As discussed in Section 4.7, the absorption coefficient of a material


is dependent on the mounting condition. Actual and laboratory
conditions of mounting may be somewhat different. Therefore, the
values of absorption coefficients used in calculating RT cannot be
absolutely precise. Additionally, our hearing mechanism is not so
sensitive as to resolve small differences in reverberation times.
Therefore, RT is generally calculated to two decimal points and finally
rounded off to one decimal point. However, if RT is less than 1 .0,
the two decimal point calculation may be retained.
An additional degree of inaccuracy that needs recognition is that
the values of absorption coefficients used in the Sabine equation are
obtained from reverberation chamber measurements where the sound
field is highly diffuse. Real rooms (auditoriums and concert halls)
depart significantly from the diffuse field of reverberation chambers.
Therefore, in practice, some disagreement between calculated and
actual RT values may be expected.

10.6 OPTIMUM Rooms meant for speech require a relatively short reverberation time.
For a small room, a value of 0.5 sec is appropriate. For large lecture
REVERBERATION rooms and auditoriums, a larger value is necessary, because in a large
TIMES room, the audience is farther away from the speaker, and hence a
greater degree of reinforcement is required.
Music requires a longer reverberation time, the actual value
depending on the type of music. Although, this aspect is discussed
further in Chapter 1 2, in general, slow and solemn music, such as
church music, is best perceived in a space with a long reverberation
time. Quick rhythmic music (light concert) requires a shorter
reverberation time. Once again, a larger room is better served by a
longer reverberation time.
Cinema halls require a very short reverberation time, since the
reverberation is built into the sound track of the film. A cinema hall
with a long reverberation time will unduly change the character of
the space in which the plot was filmed, and reduce intelligibility.
218 Chapter 1 0

Figure 10. 12 gives the values of optimum reverberation times


for various types of rooms at 500 Hz. These values must not be
regarded as absolute, rather as suggested values which may be
adjusted by plus or minus 20%. Thus, if the suggested optimum
reverberation time from Figure 10.12 is 1.0, a value lying betweeu
0.8 and 1.2 seconds is acceptable, provided the room has been
designed to satisfy other acoustical criteria, discussed in Chapters
11 to 13.

Room volume (In thou6and cubic ft)


14.0 140
7.0 10.5 17.5 35 70 105 175 350 700 1,050

3.0 .

2.5 -----1-

2.0

RT5oo
(6ec) 1.5 1

1.0 1

0.5 1

0
0.1 0.2 0.3 0.4 0.5 1.0 2.0 3.0 4.0 5.0 10.0 20.0 30.0

Room volume (In thou6and cubic m)

10.12 Suggested optimum reverberation times for various activities at 500 Hz. For cinema halls, and
recording and broadcasting studios, see Chapter 13.
Note that the values in the figure are for people with normal hearing. For hearing-impaired and older individuals, optimum
reverberation times for speech auditoriums should be less than those given above; see also Section 1 1 . 10.
The Behavior of Sound in Rooms 219

10.6.1 Increase in Low Frequency Reverberation

Reverberation time should be (reasonably) constant over the entire


frequency range. However, to provide a greater amount of reflected
sound in the low frequency region, where the ear ' s sensitivity is
low, an increase in low frequency reverberation ( 1 25 and 250 Hz) is
considered desirable. For speech halls, a 30% increase at 1 25 Hz
and a 1 5 % increase at 250 Hz is recommended4 . In other words:

Optimum RT 125 = ( 1 .3) Optimum RT500, and

Optimum RT250 = ( 1 . 15) Optimum RT500

The increase in low frequency reverberation time is particularly


important in halls meant for music. For concert halls, a quantity,
called bass ratio (BR), which is defined by the following relationship,
is used.

[ RT RT250 ]
I 25 +
BR = ( 1 0.2)
[ RT ]
soo + RTl ,OOO

Bass ratio is a measure of the "warmth" in music. As we shall


see in Chapter 1 2, preferred BR values are 1 . 1 to 1 .25 for halls with
a high RT, and 1 . 1 to 1 .45 for halls with RT of 1 .8 sec or less [ l O. IJ . A
hall in which BR is less than 1 .0 appears to lack warmth5 .

4 Several experts do not favor an increase in low frequency reverberation times,


as it increases the masking effect for bass voices.

5 BR greater than 1 . 0 is not favored for music practice rooms, as the music director
and musicians require no such enhancement while working to perfect blend,
projection, ensemble etc.
220 Chapter 10

10.7 COUPL ED ROOMS It is not uncommon for two spaces to be coupled to each other through
an opening between them. Such spaces are referred to as acoustically
coupled rooms. Most commonly occurring coupled rooms are a hall
and the stage coupled through the proscenium opening, or a hall
coupled to a deep balcony.
In coupled rooms, sound energy is exchanged between the rooms
through the opening. If the source is a continuous source and if a
steady state has been established, there is no net energy exchange
between rooms. However, when the source is turned off, the sound
in both rooms decays at their own individual decay rates.
If the reverberation times of the rooms are unequal, there will be
an energy surplus in one room as compared to the other during the
decay process. This will lead to energy transfer from the energy
surplus room to the energy-deficient room, causing a modification
in the reverberation characteristics of rooms. If the reverberation
times of both rooms are the same, there will be no energy exchange
between rooms during the decay process.

10.7.1 Source in Room with a Shorter Reverberation Time

Consider the coupled rooms of Figure 10.13, in which the RT of


room 1 is longer than that of room 2, as indicated by their decay
outlines. These decay outlines refer to the rooms in their uncoupled
situation, i.e., if they were independent rooms with no intervening
opening. The question that arises is: what will the be decay outlines
of these rooms if they are coupled together through an opening as
shown in Figure 10.13?
First assume that the source is situated in room 1, which implies
that the steady-state sound pressure level (i.e., when the source is
switched off) is greater6 in room 1 than room 2. Since the sound
decays at a faster rate in room 1, a time will come when the sound
pressure level in room 1 will begin to drop below that of room 2. At
this time, sound energy will be supplied into room 1 from room 2,
and hence, the sound decay in room 1 will assume the same rate as
that of room 2.
In other words, the decay outline of room 1 will be a two-slope
proftle, shown by heavy lines in Figure 10.13(a). The initial part of
this profile will be steep (room 1 's own decay outline), and the latter
part of the profile will follow room 2's decay outline.
The two-slope profile concept has been used successfully in the
recent design of two concert halls - Eugene McDermott Concert
Hall in Dallas, U.S.A.,
10 and Birmingham Symphony Hall,
Birmingham, U.K 21. As we shall observe in Chapter 12, a steep
decay provides "definition" to music, while a slow and sluggish
decay increases the music's "liveness" and "fullness of tone".
These two opposing requirements (a steep decay and a sluggish
decay) have been provided in the halls just mentioned by designing
them as coupled rooms. Thus, in each of these two examples, the
concert hall is coupled through openings to a large, empty and highly
reverberant space. The purpose of this empty space is to provide a
low-slope "tail" to a relatively steep reverberation of the concert
hall.

6 If the two rooms are separated by a large opening, the difference between
the sound pressure levels between the rooms, at the time when the source is
switched off, will be small. On the other hand, if the opening is small the
above difference will be large.
The Behavior of Sound in Rooms 221

D:cay outllnll of room 2, If lt D:cay outllnll of room 2, If lt w:rll not


wllrll not coupl:d to room 1 coupled to room 1. lt le aleo tnll d:cay
Two-elopll d:cay outllnll outllnll of room 1, coupl:d to room 2.
of room 1
D:cay outllnll of room 1, If lt
SPL - Dllcay outlln: of room 1, If lt SPL w:rll not coupled to room 2
Wllrll not coupl:d to room 2

Time
Time

ROOM 2 ROOM 1 ROOM 2 ROOM 1


Long:r RT Shorter RT Long:r RT Shorter RT

Sourcll E9 EB Sourcll

(a) Source in room 1 (1:1) Source In room 2

10.13 Sound decay in coupled rooms. (a) The sound source is in room 1. (b) The source is in room 2.

10.7.2 Source in Room with a Longer Reverberation Time

If the sound source in the coupled rooms were to lie in room 2, the
decay process of room 1 would be completely masked by that of
room 2, as shown in Figure 10. 1 3(b). In other words, both rooms
would have virtually the same decay rates - that of room 2.

10.7.3 Calculation of Reverberation Times of Coupled Rooms

Unless the reverberation times of the two rooms are (almost) equal,
the reverberation times of coupled rooms should be calculated
separately. This is done by assuming that the absorption coefficient
of the opening (between the two rooms) is the average absorption
coefficient of the other room. Thus, if we are calculating the RT of
room 1 in Figure 1 0. 1 3 , the absorption coefficient of the opening is
assumed as the average absorption coefficient of room 2. In
calculating the RT of room 2, we assume that the absorption
coefficient of the opening is the average absorption coefficient of
room 1 .
222 Chapter 1 0

10.8 BEH AVIOR OF The preceding discussion of the reverberation process was based on
the assumption that sound propagation in space is geometrical- in
SOUND IN A SM ALL the form of sound beams or rays. Additionally, the derivation of the
ROOM reverberation time equation is based on a few statistical assumptions.
For example, it is assumed in the derivation of reverberation time
equation that sound energy has an equal probability of falling on
every unit area of a room surface. It is also assumed that the distance
traveled by each sound ray between two successive reflections is
(4V /S) - the mean free path.
Because of these statistical and geometrical assumptions, the
behavior of sound in a room, represented by the Sabine equation, is
referred to as statistical-geometrical acoustics.
The reality, however, is that sound does not travel as rays or
beams, but in the form of waves. Thus, as the sound is reflected
from room boundaries, standing waves are formed, due to the
superposition of reflected waves with the original waves - a fact
that cannot be accounted for by assuming that sound propagation is
geometrical in nature.
The standing waves are formed only at frequencies whose
wavelength divides the room dimension by a whole number. In other
words, if a room dimension is an integral multiple of the wavelength
of sound, a standing wave is formed at that frequency, referred to as
the resonant frequency or the modal frequency. The room gives a
preferential treatment to a resonant frequency by amplifying the
sound at that frequency. The emphasizing of certain frequencies is
called tonal coloration; it makes voices sound unnatural and music
sound distorted.
Apart from tonal colorations, a standing wave increases the sound
pressure level at some locations and decreases it at other locations,
giving a nonuniform distribution of sound in the room - an
undesirable feature for music and listening spaces. The locations
where the increase and decrease of sound pressure levels occur
depend on the frequency of sound.
The number of modal frequencies in a room increases with the
size of the room, and far more sharply with the frequency of sound.
For an auditorium or a concert hall, the number of modal frequencies
- the frequencies at which the room responds preferentially - is
so large that the room can be assumed to behave nonpreferentially at
all frequencies.
In other words, the statistical-geometrical acoustics applies quite
well to a relatively large room, but not to a small room. For a small
room, the standing wave phenomenon must be taken into account,
particularly at low frequencies. It is precisely because a small room
gives a preferential treatment to a few low frequencies that the
phenomenon of bathroom baritones can be explained. In a small
bathroom with its highly reflective surfaces, the low frequencies,
particularly of a male singer, sound richer and fuller than in a large
room.
In practice, small rooms are: a broadcasting studio, sound
recording room, music practice room, rehearsal room, etc. Large
rooms are: a concert hall, auditorium, opera theater etc.
The standing wave phenomenon is also of less concern in an
irregularly shaped small room, since shape irregularities increase
the number of modal frequencies. Additionally, there are some room
proportions that reduce the standing wave effects in a small room.
For a detailed discussion of the standing wave phenomenon, refer to
Appendix C.
The Behavior of Sound in Rooms 223

Example 10.1 Reverberation Time Calculations of a Lecture Room

Calculate the reverberation times of a rectangular lecture room measuring 10 m x 15 m x 3 .65 m (33 ft
x 49 ft x 12 ft), as shown in Figure A. Interior finishes and the corresponding absorption coefficients
are given below:

Wall : gypsum board on 2 x 6 metal studs {a125 = 0.26, a500 = 0.05, a2,000 = 0.04)
Glazmg: (a 125 = 0.30, a5 00 = 0. 1 0, .ooo = 0.05 )
Flr: concrete (a 125 u.u1 . a50 = O':oz, a2 000 = 0.02)
Cellmg: sspended mmeral f1berboard ( a125 = 0.34, a500 = 0.53 , 000 = 0. 73)
D ,
Doors: sohd core flush door (a 125 = 0. 1 5, a5.oo = 0. 10, a2 000 = 0.05)
Audience: students seated in armchair seats l<X125 = 0.56, a500 = 0.79. a2,000 = 0.86, see Table 4. 1)

Solution: Since the room consists of three different heights , its volume is calculated in three parts.
Thus:
Room volume = (9.0 x 10.0 x 3 .65) + (3.0 x 10.0 x 3 .45) + (3.0 x 10.0 x 3 .25) = 529.5 m 3 . The
remaining calculations are best performed in a tabular format as shown in Table A on the following
page.

FIGURE A

15.0 m

1
(49' 0")
10.0 m
l

:3.65 m
(12'0" )

"'
Step in floor -7

/
V
10.0 m Rear 10.0 m
Front Audience area
(:3:3' 0" ) wall LECTURE ROOM 8.5 m x B m
wall (:3'0")

e.o m J / :3.0 :3.0 m

V "r b!_ 1------' 1-

Door
1.1 m x 2.5 m Door
...=
1.1 m x 2.5 m
Left wall ==

15.0 m
''
(49'0 )
224 Chapter 1 0

Table A

Element Frequency (Hz)


125 500 2,000

1 . Front wall a 0.26 0.05 0.04


2
Area = 36.5 m Sa (sabins) 9.49 1 .83 1 .46

2. Rear wall a 0.26 0.05 0.04


2
Area = 32.5 m Sa (sabins) 8.45 1 .63 1 .30

3. Left wall glazing a 0.30 0. 1 0 0.05


2
Area = 1 8.0 m Sa (sabins) 5.40 1 .80 0.90
Left wall a 0.26 0.05 0.04
2
Area = (52.95 - 1 8.0) = 34.95 m Sa (sabins) 9.09 1 .75 1 .40

4. Right wall doors (two) a 0. 1 5 0. 1 0 0.05


2
Area = 5.5 m Sa (sabins) 0.83 0.55 0.28
Right wall a 0.26 0.05 0.04
2
Area = (52.95 - 5.5) = 47.45 m Sa (sabins) 1 2.34 2.38 1 .90

5. Ceiling a 0.34 0.53 0.73


2
Area = 1 50.0 m Sa (sabins) 5 1 .0 79.5 109.5

6. Audience (ignore absorption provided by concrete floor)


a 0.56 0.79 0.86
2
Area = 8.5 m x 8.0 m = 68.0 m Sa (sabins) 38.06 53.72 58.48

7. Air m 0.00 0.00 0.009


Volume = 529.5 m3 mV (sabins) 0.00 0.00 4.77

Total absorption, EA (sabins)


fully occupied room 134.7 143.2 180.0

RT (sec) = (0.16V)/EA 0.63 0.59 0.47

Optimum RT500 = 0.60 (see Figure 10.12) ;

RT 125 = 1 .3(0.60) = 0 .7 8 0.78 0.60 0.60

0.8

Actual
RT / Optimum
Comparison of the Actual

i
(e>ec:) 0.6
and Optimum Reverl:leratlon
Tlmee
0.4 '--'-----'--.....__
125 500
Frel.\uency (Hz)
2,000
!
The Behavior of Sound in Rooms 225

Example 10.2 Reverberation Time Calculations for a 600-seat Multipurpose Hall

Calculate the reverberation times of a hall, whose plan and longitudinal section are shown in Figure A.
Interior finishes and their corresponding absorption coefficients are:

Rear wall and rear part of two side walls: Sound absorbing concrete block ( a.125 = 1 .07,
500 = 0.6 1 , z..ooo = 0.56)
All other walls: Painted concrete block (125 = 0. 10, a.500 = 0.06, .ooo = 0.09)
Stage floor: Hardwood floor (a.125 = 0.20, a.500 = 0. 15, a., 000 = 0.06)
Stage apron - vertical face between stage floor and hall floor: wood paneling (a.125 = 0.17,
5oo = 0. 10, 2,000 = 0.06)

FIGURE A

6.1 m 17.0 m
( 20'-0") (55'-9")

2b.O m
16.0 m (91'-10")
Door (2.0 m x 2.5 m}
(52'-6")
Sound al>150rl?ing
STAGE I> lock

PLA N

8.0 m
9.0 m
(26'- :3")
(2 9'-6")

SECTION
226 Chapter 1 0

Audience: Seats (upholsterd) - occupied (a 1 25 = 0.76, a500 = 0.88, .ooo = 0.91)


Seats - unoccupted (a 1 25 = 0.68, a500 = 0.82, 000 = 0.86)
Hal. floor (front part and aisles): carpet on concrete (a 125 = O.ul, a5_oo = 0.02, 000 = 0.02)
Cellmg: Sspended two-layer gypsum board (a 1 2 5 = 0. 14, a500 = O.Uo, a2,000 =
Doors: Sohd core flush doors (a 1 25 = 0.15, a500 = 0.10, ooo = 0.05)
O'.U4 )
.

Solution: This problem is basically similar to that of Example 9.1, except that the calculation of volume
and surface areas are more complex in this example. We will ftrst calculate the area of the floor by
dividing it into six component areas. The area of each component is shown within parentheses in
Figure B. Note that the areas of components (2) and (5), which are segmental in shape, have been
worked out by approximating them to triangles. Thus, the component area (2) has been assumed as a
triangle with a base of 16.0 m and a height of 1.5 m.

FIGURE B
Area of stage = Component areas (1) + (2) = 97.6 +
17.0 m 3.0 m 12.0 = 109. 6 m2
Area of hall = Component areas (3) + (4) + (5) + (6)
= 260 + 54 + 42 + 54 = 410.0 m2

Volume of stage = 109.6 x 8.0 = 877 m3

The (approximate) volume of the hall is calculated


by multiplying the area of one of the side walls by the
average width of the hall. The average width of the
hall = 0.5(28.0 + 16.0) = 22.0 m. The elevation of a
side wall is shown in Figure C. The reader may
2
1.5 m @ confirm that the total area of this wall is 126.0 m .
(260.0 m 2) Of this total area, the area of the sound absorbing block
is estimated to be 50.0 m2, the area of the painted
2
block as 71.0 m , and that of the door as 5.0 m2.

FIGURE C

5.0 m f -51!1 f

The volume of the hall = 22.0(126.0) = 2772


50.0 m 2 ........
(Sound al:>sort:>ing
@]

!10
3
t>lock)
m3. Hence,
Total volume = Volume of stage + volume Total area of wall = 126.0 m 2

of hall = V = 2772 + 877 = 3649 m3

l @l
16
(Door)
5.0 2 @ : : . 3
....___,
...._ _...-_ .-; ................................... ...........
4.5 m 12.5 m
f ')5

ELEVATION Of A SIDE WALL Of HALL

The reverberation time calculations are tabulated in Table A on the following page. Since, it is a
multipurpose hall, optimum RT500 is estimated to be 1.0 second from Figure 10.12 (as being in between
speech auditorium and opera theater).
The Behavior of Sound in Rooms 227

Table A

Element Frequency (Hz)


125 500 2,000

1 . Rear wall, stage, painted block a 0. 1 0 0.06 0.09


Area = 1 28.0 m2 Sa (sabins) 12.80 7.68 1 1 .52

2. Side walls, stage, painted block a 0. 1 0 0.06 0.09


Area = 45.8 x 2 = 9 1 .6 m2 Sa (sabins) 9. 1 6 5.50 8.24
Doors a 0.15 0. 10 0.05
Area = 2 x 3.0 = 6.0 m2 Sa (sabins) 0.90 0.60 0. 30

3. Side walls, hall, painted block a 0. 10 0.06 0.09


Area = 71.0 x 2 = 142.0 m2 Sa (sabins) 14.20 8.52 12.78
Side walls sound absorbing block a 1.07 0.61 0.56
Area = 2 x 50 = 100.0 m2 Sa (sabins) 107.00 6 1 .00 56.00
Doors a 0. 1 5 0. 10 0.05
Area = 2 x 5.0 = 10.0 m2 Sa (sabins) 1 . 50 1 .00 0.50

4. Rear wall, hall a 1 .07 0. 61 0.56


Area = 7 1 .8 m2 Sa (sabins) 76. 83 43.80 40.21
Doors a 0.15 0.10 0.05
Area = 3 x 5.0 = 15.0 m2 Sa (sabins) 2.25 1 .50 0.75

5. Ceiling, hall and stage a 0.14 0.06 0.04


Area = (410.0 + 109.6) = 5 1 9.6 m2 Sa (sabins) 72.74 31.18 20.78

6. Floor (stage) a 0.20 0.15 0.06


Area = 109.6 m2 Sa (sabins) 2 1 .92 16.44 6.58

7. Paneling, stage apron a 0.17 0. 1 0 0.06


Area = 16.0 m2 Sa (sabins) 2.72 1 .60 0.96

8. Audience (assume the audience area to include the aisles and the front circulation area)
Fully occupied hall a 0.76 0.88 0.91
Area = 410.0 m 2 Sa (sabins) 3 1 1 .60 360.80 373. 1 0

9 . Air m 0.00 0.00 0.009


Volume = 3649.0 m3 m V (sabins) 0.00 0.00 32.84

Total absorption, !:A (sabins)


fully occupied hall 633.6 539.6 564.5

RT (sec) = (0.16V)JI:A 0.92 1.08 1 .03

Optimum RT500 (Figure 1 0. 1 2) = 1 .0;


RT
125 "' 1 .3 1.3 1.0 1.0
228 Chapter 10

REFERENCES 10.1 Beranek, Leo: Concert and Opera Halls - How They Sound, Acoustical
Society of America, 1996, p. 5 1 3 .
10.2 Beranek, Leo: Concert and Opera Halls - How They Sound, Acoustical
Society of America, 1996, pp. 105 and 275.

Вам также может понравиться