Sound Insulation in Buildings
Sound Insulation in Buildings
Buildings
Sound Insulation in
Buildings
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
This book contains information obtained from authentic and highly regarded sources.
Reasonable efforts have been made to publish reliable data and information, but the author and
publisher cannot assume responsibility for the validity of all materials or the consequences of
their use. The authors and publishers have attempted to trace the copyright holders of all mate-
rial reproduced in this publication and apologize to copyright holders if permission to publish
in this form has not been obtained. If any copyright material has not been acknowledged please
write and let us know so we may rectify in any future reprint.
Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, repro-
duced, transmitted, or utilized in any form by any electronic, mechanical, or other means, now
known or hereafter invented, including photocopying, microfilming, and recording, or in any
information storage or retrieval system, without written permission from the publishers.
For permission to photocopy or use material electronically from this work, please access www.
copyright.com (http://www.copyright.com/) or contact the Copyright Clearance Center, Inc.
(CCC), 222 Rosewood Drive, Danvers, MA 01923, 978-750-8400. CCC is a not-for-profit organi-
zation that provides licenses and registration for a variety of users. For organizations that have
been granted a photocopy license by the CCC, a separate system of payment has been arranged.
Preface xv
About the Author xix
Introduction xxi
2 Mechanical vibrations 19
2.1 A simple mechanical system 19
2.2 Mechanical impedance and mobility 21
2.3 Free vibrations 23
2.4 Forced vibrations 26
2.5 Resonance 30
2.5.1 Resonance frequency 30
2.5.2 Bandwidth of a resonant system 31
2.6 Vibration insulation 33
2.6.1 The power input 33
2.6.2 Force transmitted to the foundation 34
2.6.3 Vibration isolators with hysteretic damping 37
2.6.4 The Q-factor 37
2.6.5 Insertion loss 38
2.6.6 Insulation against structure-borne vibrations 40
2.6.7 Design of vibration isolators 42
2.7 Human response to vibrations 44
2.7.1 Examples of evaluation criteria 46
References 47
12.6 Vibrations 353
12.6.1 Calibration 353
12.6.2 Measurement positions 354
12.6.3 Frequency weighting 354
12.6.4 Integration time 356
12.6.5 Noise floor 356
12.6.6 Averaging of results 356
12.6.7 The vibration dose value 356
12.7 Loss factor 357
12.7.1 Structural reverberation time 357
12.7.2 Upper limits for loss factor 357
12.8 Radiation efficiency 358
12.9 Junction attenuation 359
References 360
xv
xvi Preface
xix
Introduction
xxi
xxii Introduction
among the inhabitants. In recent years, this has led to raised minimum
requirements for sound insulation in many countries. The idea of a sound
classification system for houses, with specified sound classes better than the
minimum requirements, has aroused interest, and an international stand-
ard for sound classification of houses is under preparation (ISO/DIS 19488,
2017). Projects with experimental houses that have shown new directions
for building technology with better sound insulation are very important.
Much more can be done along this line in the future, and it is important to
spread the information to the public that good sound insulation in houses
is possible. A major problem may be to have a general acceptance among
architects and in the building industry that sound insulation is a design
parameter that must be taken seriously.
REFERENCES
H. Barkhausen (1926). Ein neuer Schallmesser für die Praxis (A new sound meas-
uring device for practical application, in German). Zeitschrift für technische
Physik 7, 599–601.
R. Berger (1911). Über die Schalldurchlässigkeit. (On the Sound Transmission, in
German). Dissertation, Munich, Germany.
P.V. Brüel (1946). Lydisolation og rumkustik (Sound insulation and room acous-
tics, in Danish). Jul. Gjellerups Forlag, Copenhagen.
L. Cremer (1942): Theorie der Schalldämmung dünner Wände bei schrägem
Einfall. (Theory of sound insulation of thin walls at oblique incidence, in
German). Akustische Zeitschrift 7, 81–104.
F. Ingerslev (1949). Akustik. Lærebog i bygningsakustik for ingeniører (Acoustics.
Textbook on building acoustics for engineers, in Danish). Teknisk Forlag,
A-S Dansk Ingeniørforenings Forlag, Copenhagen.
ISO/DIS 19488 (2017). Acoustics—Acoustic classification of dwellings. (Under
development.) International Organization for Standardization, Geneva,
Switzerland.
V.O. Knudsen (1930). Measurement and calculation of sound-insulation. Journal
of the Acoustical Society of America 2, 129–140.
V.O. Knudsen and C.M. Harris (1950). Acoustical Designing in Architecture. John
Wiley & Sons, Inc., New York. Reprint 1978 by the Acoustical Society of
America, New York.
J.W. Kopec (1997). The Sabines at Riverbank. Their Role in the Science of
Architectural Acoustics. Acoustical Society of America, New York.
A. London (1950). Transmission of reverberant sound through double walls.
Journal of the Acoustical Society of America 22, 270–279.
E. Meyer and L. Keidel (1935). Röhrenvoltmeter mit logarithmischer Anzeige
und seine Anwendungen in der Akustik, (Vacuum-tube voltmeter with log-
arithmic display and its applications in acoustics, in German). Elektrische
Nachrichtentechnik (ENT) 12(2), 37–46.
Introduction xxv
This chapter summarizes the most important basic concepts used in building
acoustics. The definitions and symbols are in accordance with International
Organization for Standardization (ISO) technical report (TR) 25417.
1.1 INTRODUCTION
The sound insulation of walls and floors in a building is often of major con-
cern for the people using the building. It is of special interest in residential
buildings, but also in, e.g. offices, hospitals and school buildings. Sound
insulation in buildings has been under research and standardization for a
long time – many contributions were made in the 1950s. However, since that
time, the behaviour and habits of people, as well as their expectations, have
changed. Noise sources have become louder and cover a broader frequency
spectrum, and it is possible that the demands of the tenants have increased.
At the same time, there are a gradual development and change in build-
ing elements and methods. Thus, knowledge of physical and other aspects
of sound insulation in buildings is essential for an acoustician. Moreover,
basic knowledge of sound insulation is also important in other branches of
acoustics, e.g. in the development of vehicles, ships and aeroplanes.
Sound insulation is divided into airborne sound insulation, where the
noise source is in the air inside or outside the building, and structure-borne
sound insulation, where the noise source is located at a building structure.
The most important case of structure-borne sound insulation is the impact
noise sound insulation, which includes noise generated by people walk-
ing. Chapter 5 gives a basic overview of the principles of sound insulation.
Chapters 8 and 9 deal more thoroughly with airborne sound insulation,
and Chapter 10 addresses the impact noise.
Airborne noise sources in a building are typically people talking, sound
from stereo equipment, televisions and musical instruments. The develop-
ment of hi-fi stereo equipment means that these sources are much louder
and cover more of the low frequencies than was the case when the building
regulations for sound insulations were developed. It is thus not given that
1
2 Sound Insulation in Buildings
p2
Lp = 10 ⋅ lg 2
(dB) (1.1)
pref
where pref = 20 μPa is the reference sound pressure. The sound pressure is
the root-mean-square (RMS) value of the time-varying sound pressure. For
harmonic vibrations, p 2 = 1/2 |p|2 , where |p| is the amplitude of the sound
pressure. In rough terms, the range of audible sound is 0 dB to 120 dB.
A sound source is characterized by its emitted sound power, which can
also be expressed in decibels. If P is the sound power in watts, the sound
power level is
P
LW = 10 ⋅ lg (dB) (1.2)
Pref
where Pref = 10 −12 W is the reference sound power. The relation between
sound power level and sound pressure level is addressed in Chapter 4.
The decibel scale is also used for many other acoustic parameters,
e.g. vibratory acceleration, vibratory velocity and sound intensity (see
Table 1.1). The reference values are defined in International Organization
for Standardization (ISO) 1683.
1.3.1 Harmonic vibrations
If the velocity of particle motions is small compared to the speed of sound
propagation in the medium, the vibrations can be considered harmonic
vibrations. Concepts such as sound pressure, velocity, acceleration, and
force can be described mathematically by either real or complex numbers as
y = y cos(ω t + ϕ ) (1.3)
In the latter case, y is a complex number that can be considered in two dif-
ferent ways: either having a size (the modulus |y|) and an angle (ωt + φ) in
the complex plane or having a real part (the cosine term) and an imaginary
part (the sine term). Whatever notation is used, the physical vibration that
can be observed is just the cosine part. However, the complex notation has
some mathematical advantages.
The imaginary unit j = −1 may also be written as j = 1ej(π /2) , i.e. as the
unit vector turned perpendicular to the real axis.
In the aforementioned expressions, t is the time, φ is a phase angle and ω
is the angular frequency:
2π
ω = 2π f = (rad/s) (1.5)
T1
The factor ejωt is called the time factor, and it is often omitted in theo-
retical derivations if it does not directly affect the result or calculations.
Sometimes, the time factor e−iωt is used instead, which is equally fine; but it
is very important to be consistent and use the same time factor throughout
any derivations or calculations.
The place factor e− jkx describes the propagation in the positive direction
of the x-axis, and it contains the angular wave number k, which is
2π ω
k= = (m −1) (1.7)
λ c
1.3.2 Speed of sound
The speed of sound is the distance travelled by sound per unit time. It is
found as the distance travelled during one period (one wavelength λ) multi-
plied by the number of periods per second (the frequency f):
c = λf (m/s) (1.8)
In air and other gases or fluids, the speed of sound is a function of the bulk
modulus K s (Pa) and the density ρ (kg/m3) of the media it travels in
Ks
c= (m/s) (1.9)
ρ
1.3.3 Frequency bands
Acoustic measurements are usually made in frequency bands. The
bandwidth Δf is the difference between the upper and lower limiting
frequencies:
∆f = f2 − f1 (Hz) (1.10)
The relative bandwidth is the ratio between the bandwidth and the centre
frequency:
∆f f2 f
= − 1 (1.12)
fcenter f1 f2
In building acoustics, the most often used frequency bands are one-third
octave bands:
f2
= 101/10 ≅ 3 2 ∆f ≅ 0.23fcenter (1.13)
f1
f2
= 103/10 ≅ 2 ∆f ≅ 0.71fcenter (1.14)
f1
In music, the octave is defined as a frequency interval where the upper fre-
quency is exactly two times the lower frequency. (The name comes from the
Latin Octavus, meaning ‘eight’.) Playing an ascending scale on the white keys
of a keyboard, the octave is step number 8. Using both black and white keys,
the octave is divided into 12 semitones. Thus, the one-third octave is the interval
of four semitones, which is called a major third in music. (In German literature
“Terz” is often used for one-third octave.) Using just intonation, the major third
has the frequency interval = 5/4 = 1.25. In the modern equal temperament, the
semitones all have equal interval = 21/12, and the major third is 21/3 ≈ 1.259921.
For comparison, the exact interval of the one-third octave is 101/10 ≈ 1.258925.
When applied to filters for acoustic measurements, ISO 266 defines the
exact and nominal center frequencies to be periodic within a decade, i.e.
the one-third octave bands are, strictly speaking, one-tenth decade bands
(see Table 1.2). If n denotes the band number, the exact center frequency is
calculated from the following equation:
Table 1.2 Exact and nominal centre frequencies in hertz of the one-third octave
bands often used in building acoustics
Band no. Exact Nominal Band no. Exact Nominal
14 25.12 25 26 398.11 400
15 31.62 31.5 27 501.19 500
16 39.81 40 28 630.96 630
17 50.12 50 29 794.33 800
18 63.10 63 30 1000.00 1000
19 79.43 80 31 1258.93 1250
20 100.00 100 32 1584.89 1600
21 125.89 125 33 1995.26 2000
22 158.49 160 34 2511.89 2500
23 199.53 200 35 3162.28 3150
24 251.19 250 36 3981.07 4000
25 316.23 315 37 5011.87 5000
∑ 10
0,1(Lp ,i + Lcorr,i )
LA = 10 lg (dB ) (1.16)
i
where L corr,i is the correction at frequency band i (Table 1.3). Actually, the
A-weighting filter is a rough approximation of the 40 phon curve.
Basic concepts in acoustics 7
130
110
100 phon
90
80 phon
Sound pressure level, dB
70
60 phon
50
40 phon
30
20 phon
10
Hearing threshold
–10
10 100 1,000 10,000
Frequency, Hz
Figure 1.1 Equal loudness contours and hearing threshold. (Adapted from ISO 226,
Acoustics – Normal Equal-Loudness-Level Contours, International Organization
for Standardization, Geneva, Switzerland, 2003.)
∂2 u 1 ∂2 u (1.17)
=
∂ x2 c 2 ∂t 2
where u is the particle velocity. In the longitudinal case, this velocity will be
in the x-direction (particle movement in the same direction as the propaga-
tion), whereas in the transversal case, it will be in the y and z directions
(more or less perpendicular to the direction of propagation). Note, how-
ever, that in the case of general non-plane wave motion, the differential
equations will be a little bit more complicated, especially in the transverse
case. As the waves are described by the ordinary wave equation, they will
(a) λI CI (b) λs CS
(c) λk CK (d) λB CB
Figure 1.2 Important types of waves. (a) Longitudinal; (b) quasilongitudinal; (c) shear; (d)
bending. (Reproduced from Kristensen, J., Rindel, J.H., Bygningsakustik – Teori og
praksis (Building acoustics – Theory and practice, in Danish), SBI-Anvisning 166,
Danish Building Research Institute, Hørsholm, Denmark, 1989. With permission.)
Basic concepts in acoustics 9
behave in the same way as the acoustic wave in air; only the magnitude of
the wave speed c will differ. Also, note that c will be constant with respect
to the frequency. However, a combined wave such as the bending wave can-
not be described by the ordinary wave equation, and the wave speed will be
frequency dependent. This is called dispersion.
1.5.1 Longitudinal waves
The bulk modulus of a solid is the ratio between a small pressure increase
and the resulting decrease in volume or increase in density. It is slightly
different under adiabatic or isothermal conditions, which depends on the
thermal conductivity of the material. In solid structures, the bulk modulus
and thus the speed of longitudinal waves depend on the actual shape of the
structure. The simplest case is that of a bar whose sides are free to move
as the compression travels through the bar. Hence, the bulk modulus of
the bar equals the Young’s modulus E, and the speed of the compression
wave is
E
cL,bar = (m/s) (1.18)
ρm
E
cL,plate = (m/s) (1.19)
ρ m (1 − µ 2 )
E(1 − µ)
cL,extended = (m/s) (1.20)
ρ m (1 + µ)(1 − 2µ)
Most solid materials have Poisson ratios between 0.1 and 0.5, the latter
being the theoretical upper limit. Some examples are given in Table 1.4.
h λB CB
surrounding air. The reason for this is partly due to the transverse out-
of-plane motion and the low impedance associated with this motion, and
partly due to the frequency dependency of the wave speed, as will be dis-
cussed later.
A plate with thickness h can be characterized by the mass per unit area,
m = ρ m h (kg/m 2 ) (1.21)
Eh3
B= (Nm) (1.22)
12(1 − µ 2 )
E
G= (N/m 2 ) (1.23)
2(1 + µ)
The pure bending wave, typical for thin plates and long wavelengths, is
controlled by the bending wave equation. In the case of a plane wave propa-
gating in the x-direction, the bending wave equation is
∂4 v ∂2 v
B 4
+m 2 =0 (1.24)
∂x ∂t
order in the time derivative. This implies that the wave speed is frequency
dependent and given by
B Eh2
cB = ω 4 = 2π f 4 (m/s) (1.25)
m 12ρ m (1 − µ 2 )
The bending wave speed is thus proportional to the square root of the fre-
quency. This is in good agreement with reality at low frequencies (and thin
plates). However, it is unrealistic to have a wave speed that approaches
infinity in the limit of high frequencies. The bending wave equation and
wave speed are low-frequency approximations.
Another transversal wave is the shear wave, in which the sections of a
plate are shifted in parallel, perpendicular to the direction of propagation
(Figure 1.2). The speed of a shear wave is
G E
cS = = (m/s) (1.26)
ρm 2ρ m (1 + µ)
−1/3
1 1
cB,eff ≅ 3 + 3 (1.27)
cB cS
1000
Phase speed, m/s
c = 344 m/s
fc fs
100
10 100 1,000 10,000
Frequency, Hz
cB cS cB,eff
Figure 1.4 The effective speed of transverse waves in a plate showing the transition from
bending to shear. In this example, the speed of shear waves is three times
c, the critical frequency is 80 Hz, and the crossover frequency is 720 Hz.
1.5.3 Critical frequency
The acoustic wave speed in air is independent of the frequency (c ≈ 344 m/s
at 20 °C), whereas the bending wave speed and the effective transverse wave
speed are frequency dependent. This has the consequence that there exists a
frequency where the two wave types have the same speed and wavelength,
and therefore couple easily. A good coupling means that the plate will easily
radiate sound at and above this frequency. Actually, it is tempting to adopt
the Mach number M known from fluid dynamics and aircraft technology.
This is the ratio of speed of propagation to the speed of sound in air. Thus,
the Mach number of a structural wave is M = cB,eff /c.
The critical frequency fc is defined as the frequency at which cB = c = 344 m/s
(Mach 1). Thus, from Equation 1.25, it follows that
c2 m c 2 3ρ m (1 − µ 2 ) Kc
fc = = = (1.28)
2π B πh E h
c s2 m 1 3E(1 − µ) (1.29)
fs = =
2π B 2π h ρ m (1 + µ)
Basic concepts in acoustics 13
f s cS2 E (1.30)
= 2 = 2
fc c 2c ρ m (1 + µ)
i.e. only dependent on the material properties. Examples for some building
materials are given in Table 1.4.
14 Sound Insulation in Buildings
1.6 AVERAGING
1.6.1 Arithmetic average
1
a=
2
( x + y) (1.31a)
n
a=
1
n ∑x i (1.31b)
i =1
1.6.2 Geometric average
a= xy (1.32)
1.6.3 Harmonic average
2xy
a= (1.33)
x+y
The harmonic average is, in fact, the arithmetic average of the reciprocal
values, 1/a = (1/x + 1/y)/2.
Example: The characteristic distance from a reflecting surface is the
harmonic average of the distance to the source and the distance to the
receiver.
1.6.4 Energetic average
1 n
La = 10 lg
n
∑10
i =1
0.1⋅Li
(dB)
(1.34)
1.7.1 Trigonometrical formulae
1.7.2 Euler’s formulae
ejx + e− jx
cos ( x) = (1.39)
2
ejx − e− jx
sin ( x) = (1.40)
2j
sin ( x)
sinc ( x) = (1.42)
x
Note that sinc (0) = 1. This function is also called the 0th order spherical
Bessel function of the first kind, j0(x). The function is shown graphically in
Figure 1.5.
1.7.4 Logarithmic functions
The common logarithm, base 10:
lg ( xy ) = lg ( x ) + lg ( y ) (1.44)
16 Sound Insulation in Buildings
0.8
0.6
0.4
sinc (x)
0.2
–0.2
–0.4
0 5 10 15 20
x
( )
lg xn = n lg ( x )
(1.45)
y = ex ⇔ x = ln ( y ) (1.46)
lg ( x)
ln ( x) = (1.47)
lg (e)
ln ( x )
lg ( x ) = (1.48)
ln (10)
Modulus : z = Re { z } + Im { z } 2
2
(1.52)
Im { z }
Argument : θ = arctan (1.53)
Re { z }
z1 z1 j(θ1−θ2 )
Division : = e (1.57)
z2 z2
REFERENCES
Mechanical vibrations
mg
xs = (2.1)
ks
where g = 9.81 ms−2 is the acceleration due to gravity and ks is the static
stiffness (N/m).
In the case of a dynamic excitation of the system by the force F, we have
the equation of motion from Newton’s second law:
d2 x dx
m 2
+r + kd x = F (2.2)
dt dt
19
20 Sound Insulation in Buildings
Figure 2.1 (a) Simple mechanical system. (b) Static deflection due to mass load and gravity.
(c) Dynamic excitation by an external force.
where x is the displacement from the rest position, t is the time and kd is the
dynamic stiffness (N/m).
As long as the external force is active, the system is said to display forced
vibrations. However, when the force stops, the system can still exhibit
vibrations for a shorter or longer time after the force is stopped. These are
resonant or free vibrations.
Let us assume that the external force is harmonic, i.e. it varies with time
as a sinusoidal function:
F = F0 cos(ω t + ϕ ) (2.3)
Imaginary
|F |
‹
F
ωt
φ
Real
Figure 2.2 A harmonically varying force displayed as a phasor in the complex plane.
Again, we consider the simple mechanical system in Figure 2.1. The system
is excited by the harmonic force F:
F = F ej(ωt+ϕ ) (2.5)
The vibrations exhibited by the system can be characterized by the deflec-
tion x, the velocity v, or the acceleration a. As long as the force is acting, the
system exhibits forced vibrations, implying that the vibrations have the same
angular frequency ω as the acting force. Thus, the velocity can be written as
v = vejωt (2.6)
where we have chosen to set the phase angle of the velocity to zero.
Mechanical impedance Zm of a system is defined as the complex ratio
between a harmonic force and the resulting velocity:
F F jϕ
Zm = = e (2.7)
v v
The unit is N s/m. The phase angle between force and velocity is φ.
The force and the velocity can be determined either in the same point
or in two different points. Therefore, we distinguish between the driving
point impedance and the transfer impedance. The latter is defined by
F1
Zm,12 = (2.8)
v2
22 Sound Insulation in Buildings
where F 1 is a harmonic force that excites the system in one point, and v 2 is
the resulting velocity in another point of the system, both taken as complex
quantities.
The mobility Ym of a mechanical system is the inverse of mechanical
impedance, i.e.
v v − jϕ
Ym = = e (2.9)
F F
The mobility is sometimes called mechanical admittance.
For a mass m, we assume a harmonic vibration (Equation 2.6) and get
the following relation from Newton’s second law:
dv
F=m = jω mv (2.10)
dt
F = kd x = kd ∫ v dt = kjωv
d
(2.11)
F = rv (2.12)
Table 2.1 shows the point impedance and mobility for the three basic
mechanical elements and for an extended system, namely, a large homo-
geneous plate.
An important system for applications in building acoustics is a large and
thin, homogeneous plate. This may be excited in a point by a harmonic
force in the normal direction perpendicular to the surface of the plate.
Bending waves are generated propagating radially in the plate outgoing
from the excitation point. The driving point impedance of the plate is found
to be surprisingly simple (Cremer and Heckl, 1967):
F0 4c 2m ′′
Z0 = = 8 m ′′B = (2.13)
v0 π fc
where m″ is the mass per unit area and B is the bending stiffness per unit
width. The critical frequency fc (Equation 1.28) is inserted in the last term.
It is noted that the point impedance is independent of frequency and real,
i.e. force and velocity are in phase. It can be shown that the impedance
(Equation 2.13) is also a good approximation for the average velocity due
to point excitation of a finite plate.
Mechanical vibrations 23
Dashpot: r (Ns/m) r 1
r
Plate 8 m′′B 1
Mass per unit area: 8 m′′B
m″ (kg/m2)
Bending stiffness:
B (Nm)
2.3 FREE VIBRATIONS
Free vibrations of the mechanical system are described through the solution
of the equation of movement (Equation 2.2) when the force is F = 0:
d2 x dx
m +r + kd x = 0 (2.14)
dt 2 dt
r < mkd : The system is weakly damped and can exhibit free harmonic
vibrations.
r = mkd : The system is critically damped.
r > mkd : The system is overdamped and a deflection slowly returns to the
steady-state position.
Only the first case is relevant for acoustical applications and is discussed
in the following.
The equation of movement (Equation 2.14) can be rewritten in the case
of a weakly damped system as
d2 x dx (2.15)
2
+ 2δ + ω 02 x = 0
dt dt
24 Sound Insulation in Buildings
r
δ = (2.16)
2m
kd
ω0 = (2.17)
m
where x0 and the phase angle θ depend on the initial conditions and the
angular frequency of the damped vibration is
ω r = ω 02 − δ 2 (2.19)
The time function of the vibration is shown in Figure 2.3. The phase angle
is set to zero. The amplitude of the vibration is x 0 e−δt, and thus, it decreases
exponentially with time. It can be noted from Equation 2.19 that the angu-
lar frequency of the system, ωr, is lower than the angular frequency of the
undamped system ω 0.
1
Deflection x / x0
–1
Time t
Figure 2.3 Time function of the free vibration of a weakly damped system with one
degree of freedom.
Mechanical vibrations 25
The system is said to exhibit a natural mode of vibration, and the associ-
ated frequency is called the natural frequency fr = ωr/2π.
The undamped natural frequency f 0 = ω 0/2π is often referred to as the
resonance frequency. However, this is only correct if a velocity resonance
is understood (see more about this in Section 2.5). The undamped natural
frequency is
ω0 1 kd
f0 = = (Hz) (2.20)
2π 2π m
1 2π 2π
T1 = = = (s) (2.21)
fr ω r ω 02 − δ 2
From Equation 2.18, it can be seen that the ratio between deflections x
at time t and t + T1 is e δT 1. The natural logarithm of this ratio is called the
logarithmic decrement Λ:
x(t)
Λ = ln = δ T1 (2.22)
x(t + T1)
1
τ= (s) (2.23)
δ
The energy of the vibration is related to the velocity v, which is easily found
when Equation 2.18 is written in the complex notation:
dx (2.25)
v= = (jω r − δ )x0e−δ t ejω rt e− jθ
dt
1 2 1 (2.27)
E= m v = mω 02 x02e−2δ t = E0e−2δ t
2 2
26 Sound Insulation in Buildings
−dE / dt 2δ r r
η= = = = (2.28)
ω 0E ω0 mkd ω 0m
The reverberation time T is primarily used in room acoustics, but it may
also be used in relation to the decay of free vibrations. By definition, T is
the time for the energy to decrease to 10 −6 of the initial energy, which cor-
responds to a decay of 60 dB. For t = T, we get from Equation 2.27
E
= e−2δ T = 10−6 = e−6 ln(10) (2.29)
E0
and the reverberation time is
2.4 FORCED VIBRATIONS
dx d2 x
= jω x and = −ω 2 x (2.31)
dt dt 2
we get
Fˆ Fˆ
xˆ = = (2.32)
−ω m + jω r + kd jω r + j(ω m − kd / ω)
2
The velocity of the vibration is
dx F
v= = jω xˆejωt = (2.33)
dt r + j(ω m − kd / ω)
Mechanical vibrations 27
F
Zm = = r + j(ω m − kd / ω) (2.34)
v
It may be noted that this formula is analogous to Ohm’s law for a simple
electrical circuit.
The modulus and argument of the mechanical impedance are
Zm = r 2 + (ω m − kd / ω )2 (2.35)
ω m − kd / ω
θ = arctan (2.36)
r
where Zm = |Zm|ejθ.
Applying the mechanical impedance of the system, we get the following
formulas for the force, velocity and deflection in complex notation:
F = F ejωt (2.37)
F F j(ωt −θ )
v= = e (2.38)
Zm Zm
F F
x= = ej(ωt −θ −π /2) (2.39)
jω Zm ω Zm
Here we have chosen the force as a reference for the phase angle. Figure 2.4
shows these three parameters as phasors in the complex plane. In the physi-
cal world, the parameters appear as projections on the real axis with har-
monic variation (the exponential functions are replaced by cosine functions).
The displacement x can be found from Equation 2.39 by inserting
Equation 2.34 and using Equations 2.17 and 2.28:
F F / kd
x= = (2.40)
m 2 ω2
( 2
jω r + j ω − ω 0 jη ⋅
ω
)ω
+ 1 − 2
ω0 ω0
Imaginary
|F|
θ |F|
v=
ωt |Zm|
Real
π
2
|F |
x=
ω|Zm|
Figure 2.4 Display of force, velocity and displacement as vectors in the complex plane.
The vectors rotate with the angular frequency ω, but the relative positions remain
constant.
π ω 0 ω −1
θ + = arctan η ⋅ − (2.42)
2 ω ω0
Figure 2.5 shows the amplitude and phase of the displacement as functions
of frequency for different values of the loss factor. (Note that the relative
angular frequency and the relative frequency are identical, i.e. ω/ω 0 = f/f 0.)
At low frequencies (ω ≪ ω 0), the deflection amplitude is almost indepen-
dent of the frequency and the phase angle is close to zero, which means that
the deflection follows the force without delay. The mechanical impedance is
dominated by the stiffness term |Zm| ≈ kd /ω.
At the resonance frequency (ω = ω 0), the deflection amplitude increases
dramatically for small loss factors, and more moderately for large loss fac-
tors. The phase angle is π/2 and the deflection is one-fourth period behind
the force, while the velocity is in phase with the force. The mechanical
impedance is dominated by the damping term |Zm| ≈ r.
At high frequencies (ω ≫ ω 0), the deflection amplitude decreases rapidly
with frequency and the phase angle is close to π, which means that deflec-
tion and the force are almost in opposite phase. The velocity is behind
the force and the acceleration is in phase with the force. The mechanical
impedance is dominated by the mass term |Zm| ≈ ωm.
Mechanical vibrations 29
10
0,1
(x/F) kd
1
0,2
0,5
1
0.1
0.1 1 10
f /f0
3
Phase angle, θ + π/2 (rad)
2
0,1
0,2
0,5
1
1
0
0.1 1 10
f /f 0
Figure 2.5 Amplitude of deflection (upper graph) and phase angle between deflection
and force (lower graph) for different values of the loss factor.
2.5 RESONANCE
2.5.1 Resonance frequency
If we consider the deflection of the system, it follows from Equation 2.40
that the maximum deflection occurs when the denominator
2 2
ω ω2
η ⋅ +
ω 1 −
0 ω 02
1
ω res, x = ω 0 1 − η 2 = ω 02 − 2δ 2 (2.43)
2
F F
x max = = (2.44)
1 r ω −δ2
2
k d η 1 − η2 0
F F / mkd
v= = (2.45)
ω0mη + jm / ω (ω 2 − ω02 ) η + j(ω / ω0 − ω0 / ω)
From this, it is obvious that the angular resonance frequency for velocity
resonance is
ω res, v = ω 0 (2.46)
F F
v max = = (2.47)
η mkd r
Mechanical vibrations 31
Table 2.2 N
atural frequencies and resonance frequencies of a simple
mechanical system with one degree of freedom
Natural frequency
Undamped Damped
Angular kd ω r = ω 02 − δ 2
frequency ω0 =
m
(rad/s)
1 kd η2
Frequency (Hz) f0 = fr = f0 1 −
2π m 4
Resonance frequency
Deflection Velocity Acceleration
Angular ω 02 − 2δ 2 ω0 ω 02
frequency
(rad/s) ω − 2δ 2
2
0
η2 f0
Frequency (Hz) f0 1 − f0
2 η2
1−
2
ω0 ω0
ω res, a = = (2.48)
1 ω 0 − 2δ 2
2
1 − η2
2
2
v 1 / (mkd ) 1 / (mkd )
2 = = (2.49)
F η 2 + (ω / ω 0 − ω 0 / ω )2 η 2 + (f / f0 − f0 / f )2
32 Sound Insulation in Buildings
η 2 = (f / f0 − f0 / f )2
1 1 2 2 2
⇒ f = ± η f0 ± η f 0 +f 0
2 4
1
f1 = fm − η f0
2
(2.50)
1
f2 = f m + η f0
2
where the mean frequency fm is
1 1
fm =
2
( f1 + f2 ) = f0 1 + η 2
4
(2.51)
1
Relative |v|2
Br = η f 0
0.5
0
f1 f0 f2
Frequency
Br = f2 − f1 = η f0 (2.52)
1 2 2
fc = f1f2 = fm2 − η f0 = f0 (2.53)
4
2.6 VIBRATION INSULATION
T1
1
P = v⋅F =
T1 ∫ v ⋅ Fdt (2.54)
0
where T1 = 1/f = 2π/ω here is the period duration of the forced vibration. The
velocity and the force are assumed to be harmonic functions and, thus, the
effective RMS (root-mean-square) values and the amplitudes are related as
2 1 2 1 2
veff = v and Feff2 = F
2 2 .
Inserting the real part of v and F from Equations 2.37 and 2.38 yields
34 Sound Insulation in Buildings
T1
P=
1
T1 ∫ v ⋅ cos(ωt − θ) ⋅ F ⋅ cos(ωt)dt
0
T1
=v⋅F
1
T1 ∫ 12 [cos(2ωt − θ) + cosθ ] dt
0
(2.55)
1
= v ⋅ F cos θ
2
Note the importance of the phase angle θ.
The same derivation can be made by using complex notation:
1
P=
2
v ⋅ F ⋅ Re ejθ { }
1
=
2
{
⋅ Re v e− jω t ejθ ⋅ F ⋅ ejω t } (2.56)
1
= ⋅ Re {v* ⋅F }
2
where v* is the complex conjugate of v (see Equation 1.54).The power
input can be related to the velocity at the excitation point by means of the
mechanical point impedance Zm (Equation 2.7):
1 v 1 2 F
P= 2
⋅ Re v * ⋅F = v ⋅ Re = veff ⋅ Re {Zm } (2.57)
2 v 2 v
In a similar way, the power input can be related to the excitation force and
the mobility Ym (Equation 2.9):
1 2 v
P= F ⋅ Re = Feff2 ⋅ Re {Ym } (2.58)
2 F
It may be noted that the following relation exists:
1 Re {Zm }
Re {Ym } = Re = 2 (2.59)
Zm Zm
m v
k r
F΄
dx
F′ = r + kd x = ( r − j kd ω ) v (2.61)
dt
F′ r − jkd / ω
H= = (2.62)
F r + j(ω m − kd ω)
Introducing the natural frequency ω 0 (Equation 2.17) and the loss factor η
(Equation 2.28), yields
η − jω0 / ω
H= (2.63)
η + j(ω / ω0 − ω0 ω)
Figure 2.8 displays the numerical value of the transmissibility, i.e. the ratio
between the amplitude of the forces expressed in dB, as a function of the
relative frequency for various values of the loss factor. Note that ω/ω 0 = f/f 0.
The transmissibility in dB is calculated from
η 2 + (ω 0 / ω )
2
F′ 2
20 lg H = 10 lg = 10 lg (2.64)
F2 η 2 + (ω / ω 0 − ω 0 ω )2
At low frequencies, i.e. for ω ≪ ω 0, the vibration isolator is stiff and the force
is transmitted to the floor without attenuation. If the excitation frequency
36 Sound Insulation in Buildings
ω is near ω 0, the transmitted force is stronger than the excitation force. The
amplification is due to resonance, and if the loss factor is small, the ampli-
fication can grow very high. Thus, the frequency range near the resonance
frequency should always be avoided.
Only when the frequency ω > 2ω 0, does the attenuation start. It is vital
that vibration isolators are designed to give a resonance frequency well
below the lowest excitation frequency, and a factor of three may be a practi-
cal minimum (ω ≈ 3ω 0) to obtain an efficient isolation.
If the loss factor is small and η ≪ ω 0/ω ≪ 1, the approximate transmis-
sibility is
( )
−2
≅ (ω 0 / ω )
4
H 2 ≅ 1 − (ω / ω 0 )2 (2.65)
This means a slope of –12 dB per octave. However, at very high frequen-
cies, ω 0/ω ≪ η, the slope changes to –6 dB per octave, which follows from
the approximation:
40
20
Transmissibility, 20 log |H| (dB)
–20
–40
–60
–80
0.1 1 10 100
Relative frequency, f /f0
Figure 2.8 The numerical value of the transmissibility with viscous damping, shown as a
function of the relative frequency for four different values of the loss factor.
Mechanical vibrations 37
2
η2 ηω 0
H 2≅ ≅ (2.66)
η + (ω / ω 0 )
2 2
ω
η2 + 1
20 lg H = 10 lg (2.68)
( )
2
η 2 + (ω / ω 0 ) − 1
2
This is displayed in Figure 2.9. The main difference between this and vis-
cous damping is that the damping above the resonance frequency is practi-
cally independent of the loss factor. The slope is –12 dB per octave.
2.6.4 The Q-factor
When a resonant system is excited with a frequency near the resonance
frequency f 0, the system will cause an amplification instead of an attenua-
tion. The Q-factor is defined for a lightly damped system as a measure of
the sharpness of the resonance peak as
f0 1
Q= = (2.69)
Br η
40
20
Transmissibility, 20 log |H| (dB)
–20
–40
–60
–80
0.1 1 10 100
Relative frequency, f /f0
Figure 2.9 The numerical value of the transmissibility with hysteretic damping, shown as
a function of the relative frequency for four different values of the loss factor.
1
20 lg H = 10 lg 1 + 2 = 10 lg 1 + Q2 ≈ 20 lg Q
η
( ) (2.70)
2.6.5 Insertion loss
The efficiency of vibration isolators is not solely determined by the trans-
missibility of the vibrating system. The supporting floor may also influence
the result. Hence, a better measure for the efficiency of vibration isolators
is the insertion loss, defined by
P
∆L = 10lg (dB) (2.71)
P′
where P and P′ refer to the power transferred to the floor without or with
the vibration isolator, respectively (Figure 2.10).
The point impedance of the floor is called Zm and the velocity generated
in the floor is v or vʹ without or with the vibration isolator, respectively.
From Equation 2.57, we get
Mechanical vibrations 39
(a) (b)
F
m v
F kd r
F΄
m
vu v΄
2
2
veff ⋅ Re {Zm } v
∆L = 10lg = 10lg 2 (dB) (2.72)
v′eff ⋅ Re {Zm }
2
v′
The insertion loss is simply the difference between the velocity levels mea-
sured on the floor without and with the vibration isolator.
The force F acts against the point impedance, which is Zm + jωm without
the vibration isolator, where m is the total mass of the machine. With refer-
ence to Figure 2.10, we have
F F′
= Zm + jω m , = Zm (2.73)
v v′
2 2
F ⋅ Zm jω m
∆L = 10lg 2 2
= −20lg H − 20lg 1 + (dB) (2.74)
F′ ⋅ Zm + jω m Zm
(ω m) (dB)
2
where m″ is the mass per unit area (kg/m 2) and B is the bending stiff-
ness (Nm) of the floor. Obviously, the efficiency of the vibration isolators
depends on whether the mass and stiffness of the floor are sufficient to pro-
vide a reasonable withhold against the isolators. In order to insulate against
40 Sound Insulation in Buildings
Figure 2.11 Vibration isolation of a machinery: (a) Insufficient resistance from the floor,
(b) increased mass of floor, (c) increased stiffness of floor. Drawing by Bernt
Forsblad. (Reproduced with permission from Arbetarskyddsnämnden, Buller
bekämpning. Principper och tillämpning. (Noise abatement. Principles and
applications, in Swedish), Arbetarskyddsfonden, Sweden, 1977.)
m v
kd r
ve
ve − v
r ( ve − v ) + kd = jω mv (2.76)
jω
leading to
v r − jkd / ω
= =H (2.77)
ve r + j (ω m − kd / ω )
Comparing this with Equation 2.62, we realize that the transmissibility H
of the system describes the vibration insulation in this case too. The only dif-
ference being that H is now a ratio of velocities, instead of a ratio of forces.
An example is the large anechoic chamber at the Technical University of
Denmark, DTU. For such a facility, it is crucial that the insulation against
noise and vibrations from the environment is as good as possible. The
building is a box-in-box system, where the inner box is made from 40 cm
thick concrete. The total mass is 1200 tons supported by 24 vibration iso-
lators made of rubber. Six vibration isolators are placed near each corner
(Figure 2.13).
To check the efficiency, vertical vibrations were measured simultane-
ously below and above the vibration isolators. The accelerometers were
mounted on small steel plates, which can be seen in Figure 2.13. The source
of vibrations was pile driving for the foundation of a neighbour building.
The measured difference between the velocity levels is shown in Figure 2.14
as a function of the frequency. The resonance of the system appears as an
amplification around 8 Hz; but at frequencies above 20 Hz, the attenuation
is better than 10 dB.
Figure 2.13 Vibration isolators of rubber under the large anechoic chamber at DTU.
(Photo by J. H. Rindel.)
42 Sound Insulation in Buildings
10
0
Transmissibility (dB)
–10
–20
–30
1 10 100
Frequency (Hz)
n
f = (Hz) (2.78)
60
The resonance frequency of the system can then be decided. If there are no
detailed requirements, we may apply the simple criterion:
f
f0 ≅ (Hz) (2.79)
3
1 kd g
f0 = (Hz) (2.80)
2π ks xs
where
g is the gravity acceleration (9.81 m/s2)
kd is the dynamic stiffness (N/m)
Mechanical vibrations 43
1
f0 ≅ (Hz) (2.81)
3xs
Applying the design criterion (Equation 2.79) leads to the very simple
design guide:
1 3
xs ≅ ≅ (m) (2.82)
3f02 f 2
100.0
Static deflection (mm)
10.0
1.0
0.1
1 10 100
Frequency of excitation (Hz)
Stiff 0 dB 10 dB 20 dB 30 dB
Figure 2.15 Diagram for the design of vibration isolators. The necessary static deflection
is determined from the excitation frequency and the attenuation in dB. The
shaded area is the resonance region and must be avoided. Example: excita-
tion at 24 Hz, 20 dB attenuation requires 6 mm static deflection.
44 Sound Insulation in Buildings
Figure 2.16 Water pump and electric motor mounted on a concrete block, which is
supported by four vibration isolators (not visible on the photo). (Photo by
J. H. Rindel.)
The human body can sense vibrations in the frequency range from 1 Hz
to 80 Hz. The sensitivity depends on the direction of vibration relative
to the body. Hence, in a standing position, vertical vibrations are sensed
Mechanical vibrations 45
1 120
Acceleration (RMS), m/s2
0.01 80
0.001 60
1 2 4 8 16 32 63
Frequency (Hz)
Vertical Horizontal Combined
10 140
Velocity (RMS), mm/s
Lv, dB re 10–9m/s
1 120
0.1 100
0.01 80
1 2 4 8 16 32 63
Frequency (Hz)
Vertical Horizontal Combined
Figure 2.17 Base curves for building vibrations: (top): acceleration; (bottom): velocity.
(Adapted from ISO 2631-2, Evaluation of human exposure to whole-body vibra-
tion – Part 2: Continuous and shock-induced vibration in buildings (1 to 80 Hz),
International Organization for Standardization, Geneva, Switzerland, 1989.)
46 Sound Insulation in Buildings
–10
Weighting (dB)
–20
–30
–40
0.1 1 10 100 1000
Frequency (Hz)
Acceleration, Wm, dB Velocity, Wmv, dB
v w ≅ k ⋅ aw (m/s) (2.83)
Lv , w ≅ La, w + 29 dB (2.84)
REFERENCES
pi j(ωt −kx)
ui = e (3.1b)
ρc
For a harmonic sound wave, the particle velocity component along a given
axis direction x can be obtained through partial differentiation of the sound
pressure:
∂Φ 1 ∂p
ux = − =− ⋅ (3.2)
∂x jωρ ∂x
with Φ = Φ ejωt being the velocity potential, and the sound pressure being
∂Φ
p=ρ = jωρ ⋅ Φ (3.3)
∂t
49
50 Sound Insulation in Buildings
ur
pr
x
pi
v
ui
x=0
Normally, the reflected sound wave will have lower amplitude and a dif-
ferent phase compared to the incoming sound wave. Both can be consid-
ered by introducing a complex reflection factor, R. Consequently, the sound
pressure and particle velocity of the reflected wave can be written as
pr = R ⋅ pi ej(ωt+kx) (3.4a)
pi j(ωt+kx)
ur = −R ⋅ e (3.4b)
ρc
( )
p = pi ⋅ e− jkx + R ⋅ ejkx ejωt (3.5a)
pi
u=
ρc
( )
⋅ e− jkx − R ⋅ ejkx ejωt (3.5b)
The particle velocity at the wall surface is perpendicular to the wall and
equal to the normal component of the wall velocity: v. Setting x = 0 in
Equation 3.5b yields
pi
v = u(x=0) = ⋅ (1 − R) ejωt (3.6)
ρc
The sound field in front of a wall 51
If the wall is sufficiently hard and heavy, v = 0 resulting in R = 1, and the
wall is said to be totally reflecting. In front of such a wall, we can find p
and u from Equation 3.5 as
pi
u = −2j ⋅ sin (kx) ejωt (3.7b)
ρc
Notice that none of these expressions contain a term of the form ejkx, i.e.
the combined wave is not propagating; we have a ‘standing wave’ in front
of the wall.
The energy density in the sound field is distributed between a potential
energy density and a kinetic energy density given by
where
2
1 p 2 1 p (3.9a)
wpot = ⋅ 2 = ⋅ 2
2 ρc 4 ρc
1 1
⋅ ρ u 2 = ⋅ ρ u (3.9b)
2
wkin =
2 4
2 2 2
1 pi 1 pi 1 pi (3.10)
wi = ⋅ 2 + ⋅ ρ = ⋅ 2
4 ρc 4 (ρ c) 2
2 ρc
In the standing wave in front of the wall, we have from Equation 3.7:
2 2 2
p = 4 pi ⋅ cos2 (kx) = 2 pi ⋅ (1 + cos(2kx)) (3.11a)
2 2
2 pi 2
pi
u =4 ⋅ sin (kx) = 2 ⋅ (1 − cos(2kx)) (3.11b)
( ρ c )2 ( ρ c )2
From this it can be seen that the potential energy has a maximum and the
kinetic energy has a minimum (zero) when 2kx = n2π, i.e. for x = nλ/2 where
52 Sound Insulation in Buildings
n = 0, 1, 2, 3, etc. The kinetic energy is maximum and the potential energy
is minimum (zero) when 2kx = π + n2π, i.e. for x = λ/4 + nλ/2. These are the
characteristic nodal points and maxima in a standing wave in front of a
wall.
The total energy density is found by inserting Equation 3.11 in Equations
3.9 and 3.8:
2 2 2
1 pi 1 pi pi
w= ⋅
4 ρc2
(
2 1 + cos ( 2kx ) + ⋅ ρ )
4 ( ρ c )2
2 1 − cos ( 2kx ) =
ρc2
(
(3.12) )
It appears that the total energy density is independent of the distance from
the wall. By comparison with Equation 3.10, it can also be seen that the
energy density is simply doubled compared to that of the incoming plane
wave alone.
pi
uix = cos θ ⋅ e− jk(x cos θ +y sin θ )ejωt (3.13b)
ρc
urx
ury
pr ur
θ
x
θ
pi
ui
uiy v
uix
pi
uiy = sin θ ⋅ e− jk(x cos θ +y sin θ )ejωt (3.13c)
ρc
in which the components of the particle velocity in the two axis directions
have been found using Equation 3.2.
Likewise for the reflected wave, and remembering the change in sign for
the x term, we get
pi
urx = −R ⋅ cos θ ⋅ e− jk(− x cos θ +y sin θ )ejωt (3.14b)
ρc
pi
ury = R ⋅ sin θ ⋅ e− jk(− x cos θ +y sin θ )ejωt (3.14c)
ρc
Through addition of Equations 3.13 and 3.14, we obtain for the total sound
field in front of a hard wall with reflection factor R = 1:
pi
ux = −2j ⋅ cos θ ⋅ sin(kx cos θ )e− jky sin θ ejωt (3.15b)
ρc
pi
uy = 2 ⋅ sin θ ⋅ cos(kx cos θ )e− jky sin θ ejωt (3.15c)
ρc
By comparing with Equation 3.7, we see that the sound field corresponds to
a standing wave with the distance λ/(2 cos θ) between the nodal points and
between maxima; but at the same time, this field propagates parallel with
the wall with the speed given by
c (3.16)
cy =
sin θ
Notice that the speed of the phase is c y > c, i.e. supersonic. The ‘trace’ of the
sound field seems to move with the phase speed along the y-axis.
By squaring the amplitudes in Equation 3.15, we obtain
2
p = 2 ⋅ pi
2
(1 + cos (2kx cos θ )) (3.17a)
54 Sound Insulation in Buildings
2
pi
2
ux = 2 ⋅ (
cos2 θ ⋅ 1 − cos ( 2kx cos θ ) ) (3.17b)
( ρ c )2
2
pi
2
uy = 2 ⋅ (
sin2 θ ⋅ 1 + cos ( 2kx cos θ ) ) (3.17c)
( ρc ) 2
2
pi
2 2
u = ux + uy = 2 ⋅
2
(
⋅ 1 − cos (2θ ) cos (2kx cos θ ) )
( ρc )
2
(3.18)
Here we have made use of the mathematical relation (Equation 1.36).
While the potential energy equals zero in certain nodal points (see
Equation 3.17a), this is not the case for the kinetic energy when θ ≠ 0. The
reason is that only the normal component of the particle velocity, ux is zero
at the wall surface. The tangential component uy reaches a maximum value
at the wall and is proportional to p:
2
2 p
uy = sin2 θ
( ρ c )2 (3.19)
The total energy density can be found by entering Equations 3.17a and 3.18
into Equations 3.8 and 3.9 and making use of the mathematical relation
(Equation 1.38):
2
pi
w=
ρc2
(
⋅ 1 + sin2 θ ⋅ cos ( 2kx cos θ )
(3.20) )
It is worth mentioning that the total energy density is not constant in all
parts of the sound field in the case of oblique incidence. The maxima occur,
e.g. at the wall surface.
3.3 RANDOM INCIDENCE IN A
DIFFUSE SOUND FIELD
In this section, we shall assume a hard wall only with reflection factor
R = 1. If the intensity of the incoming wave is the same from all angles of
incidence (or the probability of all angles of incidence is the same) and if
all incoming sound waves are uncorrelated, then it is possible to calculate
the potential and kinetic energy densities in the diffuse field by integrating
The sound field in front of a wall 55
the expressions valid for a single angle of incidence over the solid angle 2π.
From Equation 3.17a, we get for the pressure squared:
2π π 2
1
∫ ∫ 2 ⋅ p (1 + cos (2kx cos θ )) ⋅ sin θ d θ
2 2
p = dφ i
2π
0 0
sin(2kx)
= 2 ⋅ pi (1 +
2
2kx
2
(
) = 2 ⋅ pi 1 + sinc ( 2kx ) ) (3.21)
Here and in the following, the sinc function (Equation 1.42) is applied for
simplicity. Notice that in large distances from the wall and very close to the
wall, we have, respectively:
2 2
p ∞ = 2 ⋅ pi for x → ∞ (3.22a)
2 2 2
p x = 0 = 4 ⋅ pi = 2 ⋅ p ∞ for x → 0 (3.22b)
Likewise, from Equation 3.18, we get for the squared particle velocity:
sin ( 2kx )
pi
2
− cos ( 2kx )
2 sin(2kx) 2 kx
u = 2⋅ 2 1 − +
ρc 2kx (kx)2
Close to the wall, Equation 3.23 cannot be used, but we find the limit by
setting x = 0 in Equation 3.18 and repeating the integration:
2 π 2
pi
∫ (1 − cos (2θ)) ⋅ sinθ dθ
2
u x=0 = 2 ⋅
ρc2
0
π 2
∫ 2 ⋅ sin θ d θ
2 3
= u∞
0
= 4
u∞
2
for x → 0 (3.24b)
3
56 Sound Insulation in Buildings
The total energy density is found by inserting Equations 3.21 and 3.23 in
Equation 3.8:
Far from the wall, Equation 3.25 yields, as expected, recalling Equation 3.12:
2
pi
w∞ → = 2w i for x → ∞ (3.26a)
ρc2
Very close to the wall, we find by applying Equations 3.22b and 3.24b:
2
pi 2 5
wx =0 → 1 + = w∞ for x → 0 (3.26b)
ρc2 3 3
Hence, we see that in a diffuse sound field, the total energy density at the
walls (with R = 1) is 10 log (5/3) = 2.2 dB higher than the asymptotic value
far from the walls. Likewise, from Equations 3.22 and 3.24, we see that the
sound pressure is 3 dB higher and the particle velocity is 1.2 dB higher than
the respective asymptotic values far from the walls. In Table 3.1, the results
for the sound field at the surface of a hard wall relative to the asymptotic
values far from the wall are summarized.
The main results so far are illustrated in Figure 3.3. For normal, oblique
and random sound incidence on a hard wall, the patterns of potential,
kinetic and total energy densities are shown – all relative to the asymp-
totic values far from the wall (two times the corresponding values in the
incoming wave). In the case of random incidence, notice how the ampli-
tudes of the fluctuations decrease quickly with the distance from the wall.
Already at the distance of λ/4, the sound pressure has reached its asymp-
totic value, and only minor deviations from this value occur further away
from the wall.
Table 3.1 L
evels of sound pressure, particle velocity and total energy density at the
surface of a hard wall, relative to the asymptotic values far from the wall
Sound pressure (dB) Particle velocity Total energy density
Normal incidence 3 → −∞ dB 0 dB
Oblique incidence 3 10 lg (2 sin2θ) 10 lg (1 + sin2θ)
Random incidence 3 1.2 dB 2.2 dB
The sound field in front of a wall 57
0 0 0
2 kx 2 kx 2 kx
(d) 2 (e) 2 (f ) 2
–
ρ2 υ2 E
1 1 1
θ = 30˚
0 0 0
2 kx 2 kx 2 kx
0 0 0
2 kx 2 kx 2 kx
0 0 0
2 kx 2 kx 2 kx
Figure 3.3 Sound field energy components in front of a totally reflecting wall. Four rows
from above: plane incident sound waves with angles of incidence 0°, 30°, 60°
and diffuse field random incidence. In each case, the three columns show
potential, kinetic and total energy densities relative to the values far from the
wall. (Reproduced from Waterhouse, R.V., J. Acoust. Soc. Am., 27, 247–258,
1955, Figure 1. With the permission of the Acoustical Society of America.)
For physical reasons, one should expect the total energy density to be
equally distributed between potential and kinetic energy in the total sound
field, i.e.
∞ ∞
∫
0
∫
wpot d x = wkin d x (3.27)
0
58 Sound Insulation in Buildings
As we see immediately from Equations 3.9, 3.22a and 3.24a, the asymp-
totic values are equal:
2
pi
wpot, ∞ = wkin, ∞ = 2 ⋅ for x → ∞ (3.28)
ρc2
∞ ∞ ∞
wpot − wpot, ∞ p2 1 π λ
∫ wpot, ∞ ∫ p
∫
d x = 2 − 1 d x = sinc ( 2kx ) d x =
2
⋅ = (3.29)
k 2 8
0 0 ∞ 0
As shown by Waterhouse (1955), the same result is obtained for the kinetic
energy density and, thus, for the total energy density, too. It is remarkable
that acoustic energy is slightly increased near the walls in a diffuse sound
field. The additional energy corresponds to a sound field with constant
energy density also occupying the volume extending λ/8 beyond the physi-
cal walls. In a large room (dimensions wavelength) , the total energy is
approximately
λ
Etot = w ⋅ V + S (3.30)
8
where w is the energy density averaged over the room volume avoiding
regions very close to the walls, V is the room volume and S the total area of
the inner wall surfaces.
This implies that to find the total energy in a room based on a measure-
ment of the average energy density in the central part of the room, the
measurement result should be corrected with the amount:
λS
CW = 10 lg 1 + (dB) (3.31)
8V
This correction is sometimes called the ‘Waterhouse correction’ because
it was first derived by Waterhouse (1955) in a pioneering paper that also
treats the sound field at the edge between two walls and in wall corners.
While negligible at high frequencies, the correction amounts to a few dB
below 500 Hz, depending on the volume. Let us summarize the assump-
tions on which the Waterhouse correction is based:
where axial modes dominate the sound field. Axial modes correspond
to the case of normal incidence, where there is no correction.
• Only the energy condition of a large surface is considered. The energy
conditions near edges and corners are not included.
2 at a wall
p∞
2
2 = 4 at an edge (3.32)
pi
8 in a corner
Figure 3.4 Illustration of (a) a point of observation (black spot) and its images in a wall;
(b) a right-angled edge; and (c) a right-angled corner.
60 Sound Insulation in Buildings
12
9
Wall Edge Corner
Sound pressure level re far field (dB)
–3
–6
–9
–12
0 5 10 15 20 25
2kd
Table 3.2 Some characteristics of the sound pressure level near walls, edges and
corners assuming hard reflective surfaces and a diffuse sound field
Relative sound
Relative sound pressure level ΔL
pressure level Distance d to Distance d to at first minimum
ΔL at d = 0 (dB) first ΔL = 0 dB first minimum (dB)
Wall 3 0.25 λ 0.35 λ −1
Edge (2D) 6 0.32 λ 0.45 λ −3
Corner (3D) 9 0.53 λ 0.78 λ −11
Note: ΔL is sound pressure level relative to the asymptotic values in the far field.
a single wall to an edge and further to a corner. In the latter case, we have
variations from +9 dB to −11 dB and the distance from the corner must be
d > λ in order to reach small fluctuations within ca. ±1 dB. More details from
the curves are listed in Table 3.2.
In the previous section, we have treated the sound field in front of a wall
for a single frequency. If the sound field contains white or pink noise in a
certain frequency band Δf = f 2 − f1, we can carry out a simple energy summa-
tion of the mutually uncorrelated frequency components.
62 Sound Insulation in Buildings
For noise incident from one direction with angle of incidence θ relative
to the normal of the totally reflecting wall, Equation 3.17a gives us the
squared sound pressure amplitude. Using k = 2πf/c and the distance from
the wall x = d, we get
2 f2
p 1 4π
2 ⋅ pi
2 =
f2 − f1 ∫ 1 + cos
f1
c
fd cos θ d f
4π 4π
sin f2d cos θ − sin f1d cos θ
1 c c
= f2 − f1 +
f2 − f1 4π
f2d cos θ
c
where
2π
X=
c
fm d cos θ (f m = 12 (f2 + f1)) (3.39a)
2π
Y= ∆fd cos θ (∆f = f2 − f1) (3.39b)
c
This result was first derived and validated by Rindel (1978).
It appears that the interference pattern in front of the wall depends on
two parameters – X, which is controlled by the arithmetic mean frequency
of the frequency band, and Y, which is controlled by the bandwidth. In
addition, the distance from the wall and the angle of incidence are variables
in both parameters. The difference between the centre frequency, which is
the geometric mean, and the arithmetic mean frequency of the frequency
band is small for octave bands (fm /fc = 1.06) and negligible for one-third
octave bands (fm /fc = 1.0066).
Unless the bandwidth is zero, Equation 3.38 has the asymptotic value
of unity. This value corresponds to a simple energy summation of inci-
dent and reflected noise without any interference. Consequently, this is a
natural normalization factor for the interference function in Equation 3.38.
Figure 3.6 shows four examples of what the interference pattern looks like
for different bandwidths of noise. The abscissa is X = (2π/c)fmd cos θ. As
can be seen, a standing wave occurs near the wall – like in the case of
a pure tone of frequency fm; but the larger the bandwidth, the faster the
The sound field in front of a wall 63
(a) 2
|p|2
1
2•|pi|2
0
0.1 1 10 100
x
(b) 2
0
0.1 1 10 100
x
(c) 2
0
0.1 1 10 100
x
(d) 2
0
0.1 1 10 100
x
Figure 3.6 Examples of interference pattern corresponding to Equation 3.38 for four dif-
ferent values of the noise bandwidth. Valid for totally reflecting walls and a
single angle of incidence. (a) pure tone; (b) one-third octave bandwidth; (c)
octave bandwidth; and (d) wide band. The dashed lines are the sinc (Y) function.
p2
∆Lint = 10lg = 10lg (1 + cos(2X) ⋅ sinc(Y )) (dB) (3.40)
2⋅ p 2
i
64 Sound Insulation in Buildings
(a) 3
–3
ΔLint (dB)
–6
–9
–12
–15
0.1 1 10 100
X
(b) 3
–3
ΔLint (dB)
–6
–9
–12
–15
0.1 1 10 100
X
Table 3.3 M
inimum distance (in m) from a reflecting wall in order to have |ΔLint|
< 1 dB at 100 Hz one-third octave band or 125 Hz octave band for
oblique incidence at various angles or for a line source covering the
range from −60° to +60°
0 5.8 2.7
15 6.0 2.8
30 6.7 3.1
45 8.2 3.8
60 11.6 5.4
75 22.3 10.5
Line (±60°) 2.2 1.2
0
ΔLint (dB)
–3
–6
–9
0.1 1 10 100
X
Figure 3.8 Interference correction for one-third octave bands, averaging angles of inci-
dence with in the range from −60° to +60°. The parameter is X = (2π/c)fmd.
66 Sound Insulation in Buildings
Table 3.4 T
heoretical measurement errors due to interference for one-third
octave bands with the microphone 2 m in front of the façade
Frequency (Hz) 45° ΔLint (dB) Traffic ΔLint (dB)
100 1.7 −1.3
125 2.7 1.2
160 −2.3 0.8
200 −2.2 −0.9
250 2.0 0.8
315 −2.0 −1.0
400 −0.6 0.1
500 0.1 0.1
Note: Either loudspeaker measurements using angle of incidence 45° or measurements with
traffic noise covering the range from −60° to +60°.
3
ΔLint (dB)
–3
0.1 1 10 100
X
Figure 3.9 Interference correction for one-third octave bands and random incidence
(3D diffuse field). The parameter is X = (2π/c)fmd.
The sound field in front of a wall 67
case the noise contains substantial pure tones, the problems occur again.
Sometimes, the problem can be solved by carrying out a spatial average in
an area in front of the façade, as is the normal procedure in closed rooms.
For random incidence or 3D diffuse sound field, the effect of the band-
width can be found through numerical integration of Equation 3.21 over
the frequency range of relevance. For the one-third octave band, the result
is shown in Figure 3.9. The first minimum occurs at a distance d = 0.35 λ
from the wall, i.e. at a little longer distance than in the case of normal inci-
dence (0.25 λ).
REFERENCES
∂2 p ∂2 p ∂2 p
+ + + k2 p = 0 (4.1)
∂ x2 ∂ y 2 ∂ z 2
where p is the sound pressure and k = ω/c is the angular wave number, ω is
the angular frequency and c is the speed of sound in air. The equation can
be solved by separation of the variables, and it is assumed that the solution
can be written in the form:
1 ∂2 X 1 ∂2 Y 1 ∂2 Z
+ + + k2 = 0
X ∂ x2 Y ∂ y 2 Z ∂ z 2
This can be separated in various directions, and for the x-direction, it yields
1 ∂2 X
+ kx2 = 0
X ∂ x2
69
70 Sound Insulation in Buildings
Similar equations hold for the y and z directions. The angular wave num-
ber k is divided into three along the directions as
X(x) = Cx cos ( kx x + ϕ x )
1 ∂p
ux = − =0 for x = 0 and x = lx
jωρ ∂x
π
kx = ⋅ nx , where nx = 0,1, 2,3, … (4.3)
lx
Two similar boundary conditions hold for the y and z directions. With
these conditions, the solution to Equation 4.1 is
x y z
p = p0 ⋅ cos π nx ⋅ cos π ny ⋅ cos π nz (4.4)
lx ly lz
The time factor ejωt is omitted. The amplitude of the sound pressure does not
move with time, so the waves that are solutions to Equation 4.4 are called
standing waves. They are also called the modes of the room, and each of
them is related to a certain natural frequency (or eigenfrequency) given by
ω n ck c
fn = = = kx2 + ky2 + kz2
2π 2π 2π
2 2
(4.5)
2
c nx ny nz
fn = + +
2 lx ly lz
Some examples are shown in Figure 4.1. It is observed that the set of numbers
(nx, ny, nz) indicate the number of nodes (places with p = 0) along each coordi-
nate axis. Although it is usual to display the room modes as the distribution
of p2 , it may actually be more relevant to look at the distribution of the sound
pressure level in dB; see the last example, mode (2,2,0), in Figure 4.1.
(a) p2 p2
(2,0,0) (2,1,0)
p2 p2
(1,2,0) (3,1,0)
(b) SPL, dB
p2
(2,2,0) (2,2,0)
Figure 4.1 Examples of room modes. (a) The vertical axis is the sound pressure squared,
and the horizontal axes are along the length and width of the room. Mode (2,0,0)
is an axial mode and the others are tangential modes and (b) Mode (2,2,0) is
shown both as the sound pressure squared and as the sound pressure level in dB.
72 Sound Insulation in Buildings
n
p=
p0
8 ∑expjπ ± nl x
x
x±
ny
ly
y ± z z
lz
(4.6)
As before, the time factor ejωt is omitted, and the summation is taken over
all eight possible combinations of the + and − signs. Equation 4.6 shows
that the total sound pressure can be interpreted as the interference between
eight plane waves travelling in different directions. The wave number of the
propagation along the x-axis is kx = πnx /lx. The direction of propagation
expressed as an angle relative to the x-axis is
nx
lx nx ⋅ c
cos ϕ x = =± (4.7)
2 2 2 2 ⋅ lx ⋅ fn
nx ny nz
l + l + l
x y z
−1
nx
( ny
) (
l 2 ⋅ ln (1 − α x1 )(1 − α x2 ) + l 2 ⋅ ln (1 − α y1 )(1 − α y 2 ) )
55.3 ⋅ fn x y
(4.8)
Tn =
−c 2 n
lz
(
+ 2z ⋅ ln (1 − α z1 )(1 − α z 2 ) )
Introduction to room acoustics 73
(a)
(b)
(c)
A B
Figure 4.2 Two examples of tangential modes. A: (1,2,0) mode, left side and B: (3,1,0)
mode, right side. (a) nodal lines; (b) wave fronts in two positions with a short
time delay; and (c) directions of propagation, perpendicular to the wave
fronts.
kz
nz/lz
φz
φy
ky
φx
nx/lx
ny/ly
kx
−1
55.3 ⋅ fn nx ny nz
Tn = 2 ⋅ 2 + 2 + 2 ⋅ ln (1 − α m ) (4.9)
−c 2 lx ly lz
This shows that the axial modes have longer reverberation times than tan-
gential modes, and oblique modes have shorter reverberation times.
p(ω) = ∑ω 2
An
− ω − j2δnω n
2
n
(4.10)
n
4.1.4 Modal density
A closer inspection of Equation 4.5 shows that the natural frequen-
cies of a rectangular room may be interpreted in a geometrical way. A
three-d imensional frequency space is shown in Figure 4.5. The natu-
ral frequencies of the one-dimensional modes are marked on each of
the axes, representing the axial modes of the length, the width and the
height, respectively. Interestingly, the points in the grid represent the
oblique modes, the distance to each point from the origin is the natural
Introduction to room acoustics 75
60
50
40
SPL, dB
30
20
10
0
20 40 60 80 100 200
Frequency, Hz
Figure 4.4 Transfer function in a rectangular room with dimensions 7.7 m, 5.5 m and
3.5 m. At low frequencies, it is possible to identify the modes by their modal
numbers (compare with Table 4.1).
fz
c/2lz fy
c/2lx
c/2ly
fx
Figure 4.5 Frequency grid in which each grid point represents a room mode.
3
π f 3 8V 4 π V f
N= ⋅ 3 = ⋅ (4.11)
6 c 3 c
The axial modes are marked along each of the three axes and the tan-
gential modes are found in each of the three planes between two axes. If
these modes are also taken into account, the number of modes with natural
frequencies below the frequency f, the mode count, is (Pierce 1989, p. 293)
3 2
4π V f πS f Lf 1
N= c + 4 c + 8 c + 8 (4.12)
3
where V is the volume of the room, S is the total area of the surfaces and L
= 4 (lx + ly + lz) is the total length of all edges. At high frequencies, the oblique
modes dominate and the first term in Equation 4.12 is a good approxima-
tion for any room, not only for rectangular rooms.
The modal density is the average number of modes per hertz. From
Equation 4.12, we get
Introduction to room acoustics 77
d N 4π V 2 π S L
n(f ) = = 3 f + 2f+ (4.13)
df c 2c 8c
4π V 2
n(f ) ≅ f (4.14)
c3
A more detailed analysis of the modal grid points in the frequency domain
(Rodríguez Molares 2010, Table 7.1) leads to the modal densities of axial,
tangential and oblique modes in a three-dimensional room:
L
nax (f ) ≅ (4.15)
2c
πS L
ntan (f ) ≅ f− (4.16)
c2 2c
Number of modes per 1/3 octave band
100
n(ax)
10 n(tan)
n(ob)
n(total)
1
n(count)
0.1
20 25 31.5 40 50 63 80 100 125 160 200
1/3 octave center frequency, Hz
4π V 2 π S L
nob (f ) ≅ 3
f − 2f+ (4.17)
c 2c 8c
Adding the modal densities of the three groups of modes yields again
Equation 4.13.
Each normal mode is in itself a resonant system that can be associated
with a bandwidth around the natural frequency. From Equation 4.6, it fol-
lows that the squared sound pressure of mode n is
2
An
pn2 = (4.18)
(ω )
2
2
− ω n2 + 4ω 2δ n2
∆ω δ 2.2
Br = ≅ = (4.19)
2π π T
Measurement band
Room response, Lp, dB
3 dB
Br
Frequency
In a room, the natural frequencies of the normal modes are quite spread
at low frequencies, but at higher frequencies, they are very close together.
The number of modes within the bandwidth of the modes is called the
modal overlap index, M:
8.8 ⋅ π ⋅ V 2
M = Br ⋅ n(f ) ≈ f (4.20)
T ⋅ c3
The modal overlap index in a room increases rapidly with increasing fre-
quency, which is clearly seen in the approximation where Equations 4.14
and 4.19 are inserted. Figure 4.7 shows an example of low modal overlap.
There are three modes within the measurement band, which is 4.2 times
the modal bandwidth. Thus, M = 4.2/3 ≈ 0.7.
T
f g = 2000 (4.21)
V
It is obvious that the direct sound field near a sound source is not included
in the diffuse sound field. Neither are the special interference phenomena
that are known to give increased energy density near the room boundaries
and corners. The diffuse sound field is an ideal sound field that does not
exist in any room. However, in many cases, the diffuse sound field can be a
good and very practical approximation to the real sound field.
p12 = I1 ⋅ ρ c
In a diffuse sound field, the RMS sound pressure pdiff is the result of sound
waves propagating in all directions, and all having the sound intensity I1.
By integration over a sphere with the solid angle ψ = 4π, the RMS sound
pressure in the diffuse sound field is
2
pdiff =
∫
ψ = 4π
I1 ⋅ ρ c dψ = 4π ⋅ I 1 ⋅ ρ c
(4.22)
In the case of a plane wave with the angle of incidence θ relative to the
normal of the surface, the incident sound power per unit area on the
surface is
2
pdiff
Iθ = I1 cos θ = cos θ (4.23)
4πρ c
where pdiff is the RMS sound pressure in the diffuse sound field. This is just
the sound intensity in the plane propagating wave multiplied by the cosine,
which is the projection of a unit area as seen from the angle of incidence
(Figure 4.8).
The total incident sound power per unit area is found by integration over all
angles of incidence covering a half sphere in front of the surface (Figure 4.8b).
Introduction to room acoustics 81
(a) p1 (b)
pdiff
θ Iθ Iinc
Figure 4.8 (a) Plane wave at oblique incidence on a surface and (b) diffuse incidence on
a surface.
The integration covers the solid angle ψ = 2π and Figure 4.9 shows the defini-
tion of the angles.
2π π / 2
2
1 pdiff
Iinc =
∫
ψ = 2π
Iθ dψ =
4π ∫∫ 0 0
ρc
cos θ sin θ d θ d φ
1
1 p2 2
1 pdiff 1
=
4π
⋅ 2π ⋅ diff
ρc ∫
0
sin θ d(sin θ ) = ⋅
2 ρc 2
⋅ (4.24)
2
p diff
Iinc =
4ρc
It is noted that this is four times less than in the case of a plane wave of
normal incidence.
sinθ dφ
dθ
θ
φ y
A= ∑Sα
i
i i = S1α 1 + S2α 2 + … = Sα m
(4.25)
where S is the total surface area of the room and αm is the mean absorption
coefficient. The unit of A is m 2 . In general, the equivalent absorption area
may also include sound absorption due to the air and due to persons or
other objects in the room.
p2
E = ( wpot + wkin ) V = 2wpotV = V (4.26)
ρc2
Here and in the following, p denotes the RMS sound pressure in the diffuse
sound field (called pdiff in Section 4.2.2). The energy absorbed in the room
is the incident sound power per unit area (Equation 4.24) multiplied by the
total surface area and the mean absorption coefficient, i.e. the equivalent
absorption area (Equation 4.25),
p2
Pa,abs = Iinc Sα m = Iinc A = A (4.27)
4ρ c
If Pa is the sound power of a source in the room, the energy balance equa-
tion of the room is
dE
Pa − Pa,abs =
dt
Introduction to room acoustics 83
p2 V d 2
Pa − A= (p ) (4.28)
4ρ c ρc2 d t
With a constant sound source, a steady state situation is reached after some
time, and the right side of the equation is zero. Hence, the absorbed power
equals the power emitted from the source, and the steady state sound pres-
sure in the room is
4Pa
ps2 = ρc (4.29)
A
This equation shows that the sound power of a source can be determined by
measuring the sound pressure generated by the source in a room, provided
that the equivalent absorption area of the room is known. It also shows
how the absorption area in a room has a direct influence on the sound pres-
sure in the room. In some cases, it is more convenient to express Equation
4.29 in terms of the sound pressure level Lp and the sound power level LW,
4A
Lp ≅ LW + 10 lg 0 ( dB ) (4.30)
A
where A 0 = 1 m 2 is a reference area. The approximation comes from neglect-
ing the term with the constants and reference values:
A 2
4ρ c
V d 2
p (t) + 2
ρc d t
p (t) = 0 ( ) (4.31)
The solution to this equation can be written as
cA
− (t – t 0 )
p2 (t) = ps2e 4V (4.32)
where ps2 is the mean square sound pressure in the steady state and t 0 = 1 s is
the time when the source is turned off. It can be seen that the mean square
sound pressure, and hence, the sound energy, follow an exponential decay
84 Sound Insulation in Buildings
function. On a logarithmic scale, the decay is linear, and this is called the
decay curve (Figure 4.10).
If instead the source is turned on at time t = 0, the sound build up in the
room follows a similar exponential curve, also shown in Figure 4.10.
−
cA
t
p2 (t) = ps2 1 − e 4V (4.33)
The reverberation time T is defined as the time it takes for the sound energy
in the room to decay to one millionth of the initial value, i.e. a 60 dB decay
of the sound pressure level. Hence, for t = T
(a)
1
Relative sound pressure squared
0.8
0.6
0.4
0.2
0
0 0.5 1 1.5 2 2.5 3
Time, s
(b)
0
Relative sound pressure level, dB
–10
–20
–30
–40
–50
–60
0 0.5 1 1.5 2 2.5 3
Time, s
Figure 4.10 Build up and decay of sound in a room. Here, the source is turned on at t = 0
and turned off at t0 = 1 s. The reverberation time is 1.8 s. (a) Linear scale
(sound pressure squared), and (b) logarithmic scale (dB).
Introduction to room acoustics 85
cA
− T
p2 (t) = ps210−6 = ps2 e 4V
4V 55.3V
T = 6 ⋅ ln(10) ⋅ = (4.34)
cA cA
None of these conditions are fulfilled in ordinary rooms like classrooms and
offices that typically have a sound absorbing ceiling and highly reflecting
walls.
In Figure 4.10 we see the rapid build-up of sound energy after the source
is turned on. The point of −3 dB (half of the asymptotic final energy) is
reached after 5 % of the reverberation time, −1 dB after 11 % and −0.1 dB
after 27 %. This means that using a stationary sound source, the sound
field can be considered stationary after one-third of the reverberation time,
and measurements can start.
ps2 55.3V
Pa = ⋅ (4.35)
4 ρc cT
The reverberation time and the average sound pressure level in the rever-
beration room are measured, and the sound power level is calculated from
86 Sound Insulation in Buildings
2
pref ⋅ 55.3 ⋅ V
LW = Lp + 10lg
Pref ⋅ 4 ρc 2 ⋅ T
(4.36)
V T
= Lp + 10lg − 10lg − 14dB
V0 t0
Pa = I r ⋅ 4π r 2 (4.37)
Thus, the sound pressure squared of direct sound in the distance r from the
source is
Pa
2
pdir = ρc (4.38)
4π r 2
4Pa
ps2 = ρc
A
2
The reverberation distance rrev is defined as the distance where pdir = ps2
when an omnidirectional point source is placed in a room. It is a descrip-
tor of the amount of absorption in a room since the reverberation distance
depends only on the equivalent absorption area:
A (4.39)
rrev = = 0.14 A
16π
Introduction to room acoustics 87
At a distance closer to the source than the reverberation distance, the direct
sound field dominates, which is called the direct field. At longer distances,
the reverberant sound field dominates, and in this so-called far field, the
stationary, diffuse sound field may be a usable approximation.
An expression for the combined direct and diffuse sound field can
derived by the simple addition of the squared sound pressures of the two
sound fields. However, since direct sound is treated separately, it should be
extracted from the energy balance equation, which was used to describe
the diffuse sound field. To do this, the sound power of the source should
be reduced by a factor of (1 −αm), which is the fraction of the sound power
emitted to the room after the first reflection. Hence, the squared sound
pressure in the total sound field is
r2
2
ptotal 2
= pdir + ps2 (1 − α m ) = ps2 rev2 + 1 − α m (4.40)
r
1 4
2
ptotal = Pa ⋅ ρ c + (1 − α m ) (4.41)
4π r 2
A
The absorption area A divided by (1 − αm) is sometimes called the room
constant.
Normal sound sources like a speaking person, a loudspeaker, or a musical
instrument radiate sound with different intensities in different directions.
The directivity factor Q is the ratio of the intensity in a certain direction to
the average intensity,
4π r 2
Q= I⋅ (4.42)
Pa
Q ⋅ Pa
2
pdir = ρc (4.43)
4π r 2
This leads to a general formula for the sound pressure level as a function of
the distance from a sound source in room.
4 A0 r2
Lp ≅ LW + 10 lg + 10 lg Q rev2 + 1 − α m ( dB ) (4.44)
A r
where A 0 = 1 m 2 . This formula is displayed in Figure 4.11 for the case of
an omnidirectional sound source (Q = 1). In a reverberant room with little
sound absorption (e.g., αm < 0.1), the sound pressure level in the far field
will be approximately as predicted by the diffuse field theory, i.e. the last
88 Sound Insulation in Buildings
15
10
1
5
Relative sound pressure level, dB
0 5
10
–5
20
–10 50
100
–15
∞
–20
0.1 0.2 0.5 1 2 5 10 20
Distance from sound source r, m
term will be close to zero. In the case of a highly directive sound source
like a trumpet (Q 1) , the direct field can be extended to distances much
longer than the reverberation distance. In the latter situation, the last
term in Equation 4.44 raises the sound pressure level above the diffuse
field value.
In large rooms with medium or high sound absorption (e.g., αm > 0.2), the
sound pressure level will continue to decrease as a function of the distance
because the diffuse field theory is not valid in such a room. Instead, the slope
of the spatial decay curve may be taken as a measure of the degree of acoustic
attenuation in a room. Hence, in large industrial halls, the attenuation in dB
per doubling of the distance may be a better descriptor than the reverberation
time.
A reverberation room can be used to determine the sound power of a
sound source by measuring the average sound pressure level in the room.
If measurements are made in positions that avoid the direct sound, the last
term in Equation 4.44 becomes more correctly (Vorländer 1995):
Introduction to room acoustics 89
−A A
10 lg (1 − α m ) → 10 lg e S ≅ −4.34 ⋅ ( dB )
S
where A is the equivalent absorption area, S is the total surface area and tak-
ing the air attenuation into account (see Section 4.3.2). For correct results
at low frequencies, it is also necessary to apply the Waterhouse correction
(Equation 3.31)), and, thus, the sound power level is determined from
A A cS
LW = Lp + 10 lg + 4.34. + 10 lg 1 + (dB) (4.45)
4 A0 S 8Vf
This is how the sound power of sound sources can be measured according
to ISO 3741. In addition, the possible variation of temperature and static
pressure may also be taken into account for very high precision.
Figure 4.12 A plane wave travelling as a ray from wall to wall in a room.
90 Sound Insulation in Buildings
∑l = c ⋅t = n ⋅l
i m
(4.47)
i
where lm is the mean free path. Hence, the squared sound pressure is
c
⋅ln(1− α m )⋅t
p2 (t) = p02 ⋅ e lm (4.48)
When the squared sound pressure has dropped to 10 −6 of the initial value,
the time t is by definition the reverberation time T:
c
⋅ln(1−α m )⋅T c
10−6 = e lm ⇒ −6 ⋅ ln(10) = ⋅ ln(1 − α m ) ⋅ T
lm
13.8 ⋅ lm 13.8 ⋅ lm
T= ≈ (4.49)
− c ⋅ ln(1 − α m ) c ⋅αm
The last approximation is valid if αm < 0.3, i.e. only in rather reverberant
rooms. The approximation comes from
1 α 2 α 3
− ln(1 − α m ) = ln = αm + m + m +
1− αm 2 3
With the assumption that all directions of sound propagation appear with
the same probability, it can be shown from Kosten (1960) that the mean
free path in a three-dimensional room is
Introduction to room acoustics 91
4V
lm =
S
( three-dimensional ) (4.50)
π Sx
lm =
U
( two-dimensional ) (4.51)
55.3 ⋅ V
T= (4.52)
− c ⋅ S ⋅ ln (1 − α m )
In a reverberant room (αm < 0.3), it gives the same result as Sabine’s for-
mula, but in highly absorbing rooms, Eyring’s formula is theoretically
more correct. In practice, the absorption coefficients are not the same
for all surfaces and the mean absorption coefficient is calculated as in
Equation 4.20:
αm =
1
S
⋅ ∑Sα i i (4.53)
i
c ct
⋅ln(1−α m )⋅t ⋅( ln(1−α m )− m⋅lm )
2 2
p (t) = p ⋅ e lm
⋅e − mct 2
= p ⋅e lm (4.54)
0 0
13.8 ⋅ lm 13.8 ⋅ lm
T= ≈ (4.55)
(
c − ln (1 − α m ) + m ⋅ lm )
c (α m + m ⋅ lm )
A= ∑ S α + 4mV
i i (4.57)
i
ct
Figure 4.13 Rectangular room with a sound source and image sources, shown in two
dimensions. Image sources located inside the circle with radius ct will con-
tribute reflections up to time t.
Introduction to room acoustics 93
4
π (ct)3
N (t) = 3 (4.58)
V
The reflection density is then the number of reflections within a small time
interval dt, and by differentiation:
dN c3
= 4π t 2 (4.59)
dt V
The reflection density increases with the time squared, so the higher order
reflections are normally so dense in arrival time that it is impossible to
distinguish separate reflections. If Equation 4.59 is compared to Equation
4.14, it is striking to observe the analogy between reflection density in the
time domain and modal density in the frequency domain.
A= ∑ S α + ∑ n A + 4mV
i
i i
j
j j (4.61)
Table 4.2 Typical values of the absorption coefficient α for some common materials
Frequency (Hz)
Material 125 250 500 1000 2000 4000
Brick, bare concrete 0.01 0.02 0.02 0.02 0.03 0.04
Parquet floor on studs 0.16 0.14 0.11 0.08 0.08 0.07
Needle-punch carpet 0.03 0.04 0.06 0.10 0.20 0.35
Window glass 0.35 0.25 0.18 0.12 0.07 0.04
Curtain draped to half its 0.10 0.25 0.55 0.65 0.70 0.70
area, 100 mm air space
13.8 ⋅ lm l
T≈ ≈ 0.04 ⋅ m ( lm in m ) (4.62)
c ⋅αm αm
αwalls = 0.1
5m
20 m
10 m
Volume V = 5 × 10 × 20 = 1000 m3
Surface area S = 700 m 2
Equivalent absorption area A = 200 × 0.8 + 500 × 0.1 = 210 m 2
Mean absorption coefficient αm = A/S = 210/700 = 0.30
Mean absorption coefficient (height) αm = (0.8 + 0.1)/2 = 0.45
Mean free path (three-dim.) lm = 4V/S = 4 × 1000/700 = 5.7 m
Mean free path (two-dim.) lm = πSx /U = π × 200/60 = 10.5 m
(1 − α m ) = (1 − α1 )(1 − α 2 ) (4.63)
SPL, dB
Noise on Noise off
Decay
Background
noise level
t=0 Time, s
Figure 4.15 Typical decay curve measured with noise interrupted at time t = 0. SPL,
sound pressure level.
Introduction to room acoustics 97
close to a straight line between the excitation level and the background level.
The dynamic range is seldom more than around 50 dB and the whole range
of the measured decay curve is not used. The lower part of the decay curve is
influenced by the background noise and the upper part may be influenced by
the direct sound, which gives a steeper start of the curve. Hence, the part of the
decay curve used for evaluation begins 5 dB below the average stationary level
and ends normally 25 dB or 35 dB below the same level. The evaluation range
is, thus, either 20 dB or 30 dB and the slope is determined by fitting a straight
line or automatically by calculating the slope of a linear regression line. From
the slope of the decay curve in dB per second, the reverberation time is calcu-
lated, being the time for a 60 dB drop following the straight line. According to
ISO 3382-2, the preferred evaluation range is 20 dB and the result is denoted as
T20. Similarly, T30 is used if the 30 dB evaluation range has been applied. If the
sound field in the room is sufficiently diffuse, the decay curves are reasonably
straight and T20 and T30 are approximately the same.
Sometimes, the decay curves are not nice and straight and it is difficult
to measure a certain reverberation time; T20 and T30 yield different results.
One reason can be that it is a measurement at low frequencies in a small
room and maybe only a few modes are excited within the frequency band
of the measurement. In this case, interference between the modes can cause
very irregular decay curves.
The reverberation time is measured in a number of source and receiver
positions, often two source positions and three microphone positions, i.e.
six combinations are used. If the interrupted noise method is used, several
decays should be measured and averaged in each position because the noise
signal is a stochastic signal and, thus, the measured decay curves are always a
little different. However, if the integrated squared impulse response method
is used, a single measurement in each position is sufficient; the decay curve
obtained with this method corresponds to averaging an infinite number of
decays with the interrupted noise method (Schroeder, 1965). Figure 4.16
shows an example of a measured squared impulse response (the curve with
fluctuations) and the corresponding integrated decay curve.
Annex B of ISO 3382-2 contains two parameters that are useful for
describing the quality of reverberation time measurements. One is the
degree of non linearity of the part of the decay curve within the evaluation
range. Since the evaluation of the decay curve is made by a least-squares fit
of a straight line, the correlation coefficient r is known. The non linearity
parameter ξ is defined as the permillage deviation from perfect linearity:
(
ξ = 1000 1 − r 2 ) (‰) (4.64)
Values higher than 10 ‰ indicate a decay curve being far from a straight
line, and the value of the reverberation time derived from the curve may be
suspicious.
98 Sound Insulation in Buildings
–10
–20
SPL, dB
–30
–40
–50
0 0.5 1 1.5
Time, s
Figure 4.16 Typical decay curve measured using the integrated squared impulse response
method. The lower curve is the squared impulse response and the upper
curve is the decay curve.
T
C = 100 30 − 1 (%) (4.65)
T20
Values higher than 10 % indicate a bent decay curve; the decay may be a
combination of room modes with very different reverberation times.
REFERENCES
D.A. Bies and C.H. Hansen (1988). Engineering Noise Control, Theory and Practice,
Unwin Hyman Ltd., London.
C.M. Harris (1966). Absorption of sound in air versus humidity and temperature.
Journal of the Acoustical Society of America 40, 148–159.
ISO 3382-2 (2008). Acoustics – Measurement of room acoustic parameters – Part
2: Reverberation time in ordinary rooms, International Organization for
Standardization, Geneva, Switzerland.
ISO 3741 (2010). Acoustics – Determination of sound power levels and sound
energy levels of noise sources using sound pressure – Precision methods for
reverberation test rooms, International Organization for Standardization,
Geneva, Switzerland.
C.W. Kosten (1960). The mean free path in room acoustics. Acustica 10, 245–250.
H. Kuttruff (1973). Room Acoustics, Applied Science Publishers, London.
A.R. Molares (2010). A Monte Carlo approach to the analysis of uncertainty in
acoustics. PhD Dissertation, Escola Técnica Superior de Enxeñeiros de
Telecomunicación Universidade de Vigo, Spain.
A.D. Pierce (1989). Acoustics. An Introduction to Its Physical Principles and
Applications. 2nd Edition, The Acoustical Society of America, New York.
Introduction to room acoustics 99
This chapter describes the basic principles of sound insulation with simpli-
fied assumptions. The methods for measuring airborne and impact sound
insulation are briefly described. Also presented are the methods of objective
evaluation of the measurement results. As a supplement to the fundamental
theoretical models, the principles are illustrated through examples of mea-
surement results on typical constructions.
P2
τ= (5.1)
P1
However, the sound transmission coefficients are typically very small num-
bers, and it is more convenient to use the sound reduction index R with the
unit decibel (dB). It is defined as:
101
102 Sound Insulation in Buildings
P1 1
R = 10 lg = 10 lg = −10 lg τ (dB) (5.2)
P2 τ
Another name for the same term is sound transmission loss, which will be
used throughout this book.
p12S
P1 = Iinc S = (5.3)
4ρ c
The area of the wall is S. In the receiving room, the sound pressure
p 2 is generated from the sound power P 2 radiated into the room (see
Equation 4.29),
4P2
p22 = ρc (5.4)
A2
Room 1 Room 2
P1 P2
A2
L1 L2
Figure 5.1 Airborne sound transmission from source room (1) to receiving room (2).
Introduction to sound insulation 103
p12S p2 S S
R = 10 lg = 10 lg 21 = L1 − L2 + 10 lg (dB) (5.5)
4 ρ cP2 p2 A2 A2
L1 and L 2 are the spatially averaged sound pressure levels in the source and
receiving rooms, respectively. This important result is the basis for trans-
mission loss measurements.
55.3V2
A2 = (5.6)
cT2
where V2 and T2 are the volume and reverberation time of the receiving room,
respectively, and c is the speed of sound. Only under special laboratory con-
ditions can the transmission loss of a wall be measured without influence
from other transmission paths. In a normal building, sound will not only be
transmitted through the separating construction but the flanking construc-
tions will also influence the result, as discussed later in Section 5.5.
For measurements of sound insulation in buildings, the apparent sound
reduction index is
S
R′ = L1 − L2 + 10 lg (dB) (5.7)
A2
The apostrophe after the symbol indicates that flanking transmission can
be assumed to influence the result. Thus, R′ is typically a few dB less than
104 Sound Insulation in Buildings
R for the same partition wall measured in a laboratory. For the character-
ization of sound insulation in buildings, the R′ has a problem – the area
S of the partition must be known, and this is not always the case. For
example, the source and receiving rooms may be located with an offset
(staggered rooms) so that the area S does not exist or is very small? In the
latter case, Equation 5.7 would yield a very low result, not corresponding
to the perceived sound insulation.
To overcome this problem in field measurements, the area S can be set
to a constant value of 10 m 2 , and the result using Equation 5.7 is called the
normalized level difference Dn. An alternative measure that overcomes the
problem with the partition area is the standardized level difference DnT :
T2
DnT = L1 − L2 + 10 lg (dB), T0 = 0.5 s (5.8)
T0
This is the level difference between the source and receiving rooms if the
reverberation time of the receiving room is 0.5 s at all frequencies. This rever-
beration time is typical for a furnished living room in a dwelling. According
to ISO 16283-1, the standardized level difference DnT provides a straightfor-
ward link to the subjective impression of airborne sound insulation.
the relation between DnT and R′ is approximately: From Equations 5.6
through 5.8,
V2
DnT = R′ + 10 lg (dB), d0 = 3 m (5.9)
S ⋅ d0
the surfaces of the source room is denoted by Iinc, the total incident sound
power on the partition is
n
The total area of the partition is called S. The total sound power transmit-
ted through the partition is
n
P2 = ∑τ S I i i inc
i =1
τ res =
P2 1
=
P1 S ∑ τ S (5.10)
i =1
i i
The same result can also be written in terms of the transmission losses Ri
of each element
1 n
Rres = −10 lg τ res = −10 lg
S
∑ S 10
i =1
i
−0,1Ri
(5.11)
In the simple case of only two elements, the graph in Figure 5.2 may be used.
0 S2/(S1+S2), %
20 1
2
15
5
Rres–R2 (dB)
10 10
20
5
40
80
0
0 5 10 15 20 25 30
R1–R2 (dB)
Sap/S
60 10–6
50 10–5
40 10–4
Rres (dB)
30 10–3
20 10–2
10 10–1
0
0 10 20 30 40 50 60 70 80
R1 (dB)
1
(
)
Rres = −10 lg S110−0,1R1 + Sap ≅ −10 lg 10−0,1R1 +
S
Sap
S
(5.12)
Figure 5.3 can illustrate the result. The relative area of the aperture defines
an upper limit of the sound insulation that can be achieved.
speed of sound c (also longitudinal waves). The symbols and notation are
explained in Figure 5.4.
The sound pressure is equal on either side of the two transition planes:
pi + pr = p1 + p4
pt = p2 + p3 (5.13)
Also, the particle velocity is equal on either side of the two transition planes:
ui − ur = u1 − u4
ut = u2 − u3 (5.14)
pi pr pt
= = = Z0 = ρ c
ui ur ut
p1 p2 p3 p4
= = = = Zm = ρ mcL (5.15)
u1 u2 u3 u4
Z0
pi − pr =
Zm
( p1 − p4 )
Z0
pt =
Zm
( p2 − p3 ) (5.16)
pi p1 p2 pt
pr p4 p3
Figure 5.4 Thick wall with incident, reflected and transmitted sound waves. p1, p2 , p3 and
p4 denote the sound pressures of the longitudinal waves in the solid material,
incident and reflected at the boundaries. pi, pr and pt denote the incident,
reflected and transmitted sound pressures in the air.
108 Sound Insulation in Buildings
Assuming propagation from one side of the material to the other without
losses means that there is only a phase difference between the pressures at
the two intersections:
p2 = p1e− jkm h
p4 = p3e− jkm h (5.17)
The resulting transmission loss is shown in Figure 5.5, where some dips
can be observed at high frequencies. They occur at frequencies where the
thickness is equal to half a wavelength in the solid material, or a multiple of
half wavelengths. However, the dips are very narrow and they are mainly
of theoretical interest.
100
80
60
R0 (dB)
40
20
0
10 100 1,000 10,000
Frequency (Hz)
Figure 5.5 Transmission loss at normal incidence of sound on a 600 mm thick concrete wall.
Introduction to sound insulation 109
Two special cases can be studied. First, the case of a thin wall: Zm ≫ Z 0
and kmh ≪ 1
Z 2 ωρ h 2
R0 ≅ 10 lg 1 + m sin2 (km h) ≅ 10 lg 1 + m (5.19)
2Z0 2ρ c
The other special case is a very thick wall: Zm ≫ Z 0 and kmh ≫ 1
2
Z ρ c
R0 ≅ 10 lg m ≅ 20 lg m L (5.20)
2Z0 2ρ c
The cross-over frequency from Equation 5.19 to Equation 5.20 is the fre-
quency f h at which kmh = 1:
cL
fh = (5.21)
2π h
This is the frequency at which the thickness is approximately one-sixth of
the longitudinal wavelength λL in the material:
cL λ
h= = L
2π f 2π
The result for the thin wall (Equation 5.19) is the so-called mass law, which
will be derived in a different way in the next section. The result for a very
thick wall (Equation 5.20) means that there is an upper limit on the sound
insulation that can be achieved by a single-leaf construction, and this limit
depends on the density of the material. For wood, it is 68 dB, for concrete
80 dB and for steel 94 dB. These numbers should be reduced by 5 dB in the
case of random incidence, instead of normal incidence – see Section 5.2.3.
dvn
∆ p = pi + pr − pt = m = jω mvn (5.22)
dt
where vn is the velocity of the wall vibrations (in the direction normal to the
wall). Wall impedance Zw is introduced:
∆p
Zw = = jω m (5.23)
vn
110 Sound Insulation in Buildings
vn
pi
vt = vn /cos θ
θ
pt
pr
The wall impedance will be more complicated if the bending stiffness of the
wall is also taken into account (see Section 5.2.4).
The particle velocities in the sound waves are called u with the same
indices as the corresponding sound pressures. Due to the continuity
requirement, the normal component of the velocity on both sides of the
wall is:
vn
pt = pi − pr = Z0 (5.25)
cos θ
2 2
pi Zw cos θ
Rθ = 10 lg = 10 lg 1 + (dB) (5.26)
pt 2Z0
2
ωm π fm
R0 = 10lg 1 + j ≅ 20lg (dB) (5.27)
2ρc ρc
Since m = ρmh, this result is the same as that derived earlier in a different
way (Equation 5.19). The mass law for sound insulation was first suggested
in a dissertation by Berger (1911).
Introduction to sound insulation 111
1
τ (θ ) = 2 (5.28)
ωm
1+ cos2 θ
2ρ c
Random incidence means that the sound field on the source side of the
partition is approximately a diffuse sound field. In a diffuse sound field, the
incident sound power P1 on a surface is found by integration over the solid
angle ψ = 2π, assuming the same sound intensity I1 in all directions. Since,
in each direction, the transmitted sound power is equal to the incident
sound power multiplied by the transmission coefficient, the ratio between
transmitted and incident power is:
π /2
P ∫ψ = 2π
τ (θ )I1S dψ ∫ τ (θ) cos θ sin θ dθ
τ= 2 = = 0
∫
π /2
P1 I1S dψ
ψ = 2π
∫ cos θ sin θ dθ
0
π /2 1
=2
∫0
∫
τ (θ ) cos θ sin θ dθ = τ (θ )d(cos2 θ )
0
( )
2
d(cos2 θ ) 2ρ c
∫ ln 1 + (ω m 2ρ c )
2
= =
1 + (ω m 2ρ c ) cos θ ω m
2 2
0
where R0 is the mass law for normal incidence, Equation 5.27. This is the
theoretical result for random incidence, and for typical values (R0 between
30 dB and 60 dB), it means that R is 8 dB to 11 dB lower than R0. However,
this is not true in real life, and it can be shown that the aforementioned
result is related to partitions of infinite size. Taking the finite size into
account, the result is approximately:
R ≅ R0 − 5 dB (5.30)
Eh3
B= (5.31)
12(1 − µ 2 )
where E is Young’s modulus of the material and μ is Poisson’s ratio (μ ≅ 0.3
for most rigid materials).
The speed of propagation of bending waves in a plate with bending stiff-
ness per unit width B and mass per unit area m is (see Equation 1.25):
B f
cB = ω 4 =c (5.32)
m fc
c2 m
fc = (5.33)
2π B
A sound wave with the angle of incidence θ propagates across the wall
with the phase speed c/sin θ, i.e. the phase speed is in general higher than c
(Figure 5.7). If the bending wave speed happens to be equal to the phase
speed of the incident sound wave, this is called coincidence:
c pr pt c
θ θ
c /sin θ
c
pi θ
Figure 5.7 Thin wall with bending wave and indication of speed of propagation along the wall.
Introduction to sound insulation 113
cB = c / sin θ
The coincidence leads to a significant dip in the sound transmission loss.
The coincidence dip will be at the coincidence frequency, which is higher
than or equal to the critical frequency:
fco = fc sin2 θ
(5.34)
The coincidence effect was first described by Lothar Cremer (1905–1990)
in a pioneering paper (Cremer 1942). Taking the bending stiffness into
account, the wall impedance (Equation 5.23) is replaced by:
2
f
Zw = jω m 1 − sin4 θ (5.35)
fc
Insertion in the general form, Equation 5.26, leads to the sound transmis-
sion loss at a certain angle of incidence:
70
60
50
Transmission loss, Rθ (dB)
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
frequency. Only bending waves are assumed. This model does not include
the influence of a finite area, the normal modes at low frequencies, nor
the shear waves at high frequencies. These topics will be dealt with in
Chapter 8.
The following results are valid for sound insulation between rooms
with approximately diffuse sound fields. In the frequency range below the
critical frequency, f < fc , the sound insulation is mass-controlled and the
approximate transmission loss is:
R ≅ R0 + 20 lg 1 − ( f fc ) − 5 dB (5.37)
2
Introduction to sound insulation 115
100
90
80
70
60
Rθ (dB)
50
40
30
20
10
0
0
An 3
gle 0
of 0
in 1000
cid 60
en 1000
ce 90 Hz)
,θ 100 ncy (
Freque
Figure 5.9 Contour plot of the transmission loss of a single panel as a function of angle
of incidence and frequency. The step between contours is 10 dB. The panel
has m = 25 kg/m2 and fc = 1000 Hz.
In the frequency range above the critical frequency, f ≥ fc , the sound insula-
tion is stiffness-controlled and the resonant transmission due to the struc-
tural modes dominates the transmission. The approximate transmission
loss in this region is:
2η f
R ≅ R0 + 10 lg (dB) (5.38)
π fc
where η is the loss factor (see Equation 2.28). This result will be derived
and explained later in Chapter 7.
A sketch of the transmission loss as a function of frequency is shown
in Figure 5.10. Well below the critical frequency, the course increases by
6 dB per octave, but it changes towards the dip at the critical frequency as
described by Equation 5.37. The maximum transmission loss below fc will
be at f = fc / 3 ≅ 0.58 ⋅ fc . Insertion of this frequency and Equation 1.28
in Equation 5.37 leads to:
c2 ρ 3
ρ E
( )
Rmax, f < fc = 10 lg 2 m 1 − µ 2 − 8.5 dB (5.39)
This is a theoretical maximum, and for some building materials (Table 1.4),
this can be calculated: around 33 dB for gypsum board; 35 dB for glass,
116 Sound Insulation in Buildings
R (dB)
fc
Frequency (Hz)
brick and porous concrete; 37 dB for concrete; 44 dB for steel and 59 dB for
lead. In reality, the maximum value may be a few dB lower because reso-
nant transmission contributes to the total sound transmission, especially
when the physical dimensions are relatively small, like windows, or when
the plate is supported by laths, like gypsum boards. Hence, a maximum
around 30 dB to 32 dB is typical for single layers of gypsum board and glass.
40
30
Transmission loss, R (dB)
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
6.5 mm glass 4 mm glass
40
30
Transmission loss, R (dB)
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Laminated glass, 2 × 3.2 mm glass with 1.1 mm interlayer 6.5 mm glass
Figure 5.12 Transmission loss of a laminated and a solid glass pane with same mass
per unit area. (Adapted from Marsh, J.A., Building Acoustics, Oriel Press,
London, 1971, Figure 5.3.)
118 Sound Insulation in Buildings
39
30 22
50
40
Transmission loss, R (dB)
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
With sand, 33 kg/m2 Without sand, 17 kg/m2
40
1 2
30
15 15
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Without carving With carving
A simple and widely used method to increase the mass without too much
trouble with the critical frequency is simply to layer two or more sheets on
top of each other. With two layers instead of one, the bending stiffness is
doubled and so is the mass, i.e. the ratio is constant and the critical fre-
quency is the same as for the single sheet. Figure 5.15 shows the transmis-
sion loss of one or two layers of 13 mm gypsum board, with the critical
40
1 2 30
Transmission loss, R (dB)
20
13 26
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
13 mm gypsum board Two layers of 13 mm gypsum board
70
60
50
Transmission loss, R (dB)
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
250 mm porous concrete with plaster 250 mm porous concrete without plaster
60
50
40
Transmission loss, R (dB)
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Empty cavity With mineral wool
m1 d m2
pi p1 p2 pt
pr p4 p3
v1 v2
Figure 5.18 A double construction showing sound pressures and wall velocities. p1, p2 ,
p3 and p4 denote the sound pressures in the cavity, incident and reflected at
the boundaries. pi, pr and pt denote the incident, reflected and transmitted
sound pressures. v1 and v 2 are the velocities of the two walls when excited
by sound.
pi + pr − (p1 − p4 ) = Z1v1
p2 + p3 − pt = Z2v2 (5.40)
The velocity of each wall equals the particle velocity on either side:
1
v1 = ui − ur = (pi − pr )
Z0
1
v1 = u1 − u4 = (p1 − p4 )
Z0
1
v2 = u2 − u3 = (p2 − p3)
Z0
1
v2 = ut = pt (5.41)
Z0
Assuming propagation from one side of the cavity to the other without
losses means that there is only a phase difference between the pressures at
the two intersections:
p2 = p1e− jkd
From Equations 5.40 through 5.42, the ratio between the sound pressures pi
and pt can be derived and, thus, the transmission loss can be expressed by:
2
p
R0 = 10 lg i
pt
2
Z + Z2 Z1 + Z2 Z1Z2
= 10 lg 1 + 1 cos(kd) + j 1 + + sin(kd) (5.43)
2Z0 2Z0 2Z02
If only the mass of each wall is taken into account, the wall impedances are:
Z1 = jω m1
Z2 = jω m2 (5.44)
Neglecting the smaller parts and inserting Z 0 = ρc together with Equation
5.44 yields:
ω (m + m ) 2
ω (m1 + m2 ) ω 2m1m2
2
R0 ≅ 10 lg 1 2
sin(kd) + cos(kd) − sin(kd) (5.45)
2ρ c 2ρ c 2(ρ c)2
m1 + m2 ρ c
tan(kd) = (5.46)
m1m2 ω
For a cavity that is narrow compared to the wavelength (kd << 1), we get:
ωd
tan(kd) ≅ kd =
c
c ρ 1 1
f0 ≅ + (5.47)
2π d m1 m2
If the depth d of the cavity is comparable to the wavelength, there are many
solutions to Equation 5.46, and they are approximately kd = nπ, where n is
124 Sound Insulation in Buildings
c
fd = (5.48)
2π d
This is quite similar to the result found for the sound transmission through
a solid material (Equation 5.21). Only, in this case, the transmission is
through air. The spring-like behaviour of the air cavity changes from that
of a simple spring below the cross-over frequency to that of a transmission
channel at higher frequencies.
ω (m1 + m2 )
R0 ≈ 20 lg = R(1+ 2) (5.49)
2ρ c
This means that the construction behaves as a single construction with the
mass per unit area (m1 + m2). In the frequency range above the resonance
frequency and below the cross-over frequency, f 0 < f < fd ,
ω 3m1m2d
R0 ≈ 20 lg ≈ R1 + R2 + 20 lg(2kd) (5.50)
2ρ 2c 3
In this range, a much better sound insulation can be obtained, and the
transmission loss depends on the product of the three parameters, m1, m2
and d. At frequencies above fd where the cavity is wide compared to the
wavelength, sin (kd) is replaced by its maximum value 1, and for f ≥ fd:
ω 2m1m2
R0 ≈ 20 lg ≈ R1 + R2 + 6 dB (5.51)
2(ρ c)2
In this high-frequency range, d is no longer an important parameter.
A sketch of the transmission loss as a function of frequency is shown in
Figure 5.19. We recall that the transmission loss of a single leaf will exhibit
a dip at the critical frequency, and if the two leafs have different critical fre-
quencies, there will be two corresponding dips in the transmission loss of
Introduction to sound insulation 125
R (dB)
f0 fd fc1 fc2
Frequency
2
2
ω
R0 = R(1+ 2) + 10 lg sin (kd) + cos(kd) − ⋅ sinc(kd)
2
ω0
2
f 2
≅ R(1+ 2) + 10 lg 1 − (5.52)
f0
60
50
40
Transmission loss, R (dB)
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
2 × 4 mm glass with 85 mm distance 2 × 4 mm glass with 12 mm distance
Figure 5.20 Sound insulation of a double window with two 4 mm glass panes, and the
influence of the cavity depth. (Adapted from Ingemansson, S., Ljudisolerande
fönsterkonstruktioner (In Swedish), Stockholm, Sweden, 1968, Figure 8.)
60
50
40
Transmission loss, R (dB)
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
6.5 mm and 4 mm glass with 100 mm distance
2 × 4 mm glass with 100 mm distance
2 × 6.5 mm glass with 100 mm distance
filled with 50 mm mineral wool. The effect of the mineral wool is signifi-
cant in the entire frequency range.
Asymmetric double constructions have a special application in relation
to heavy- or medium-weight single walls. Such walls have a relatively low
critical frequency and, thus, the sound radiation from resonant modes in the
wall is very important for sound transmission. If the structural vibrations
are transferred to a thin plate with high critical frequency, the sound radia-
tion is reduced and, thus, the sound insulation is significantly improved, as
shown in the example in Figure 5.23. In fact, it is possible to improve the
sound insulation even more by applying a similar light construction to the
other side of the wall.
128 Sound Insulation in Buildings
60
50
40
Transmission loss, R (dB)
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
50 mm mineral wool in cavity Empty cavity
Figure 5.22 The influence of absorption material in the cavity on the sound insulation of
a double construction.
In Figure 5.3 and Equation 5.12, it was shown that an aperture in a wall
leads to an upper limit of the sound insulation. As a first approximation, this
limit is independent of the frequency, which in practice means that the high
frequency sound insulation is limited while the lower frequencies can remain
more or less unaffected by a small slit or a leakage. In the following, we shall
see a few examples of how slits and leaks influence the sound insulation.
The effect of the width of a slit is seen in Figure 5.24. While transmis-
sion loss of the wall without slits increase with frequency, a 1 mm wide
slit means an upper limit around 40 dB. In addition, we see very efficient
transmission around 2000 Hz and 4000 Hz. This can be explained by the
depth of the slit that equals one or two half wavelength at these frequencies.
Introduction to sound insulation 129
70
60
50
Transmission loss, R (dB)
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 5.23 Effect of a thin panel construction to reduce sound radiation from a single
wall, shown here as a 68 mm lightweight concrete wall. The thin panel is
3 mm of hardboard on 25 mm wooden laths with 25 mm mineral wool
in the cavity. (Adapted from Brandt, O., Akustisk planering. (In Swedish),
V. Pettersons Bokindustri AB, Stockholm, Sweden, 1958, Figure 3.24b.)
In Figure 5.25, the 3 mm slit from the previous figure is filled with felt,
and we see that such a filling can be efficient; the transmission loss with felt
in the slit is nearly the same as that without any slit in Figure 5.24.
Doors can be particularly difficult in terms of sound insulation, because
it sometimes becomes necessary to avoid any sealing lists at the floor, e.g.
in hospitals. In such cases, it is possible to apply special moving sealing lists
that close when the door is closed. However, it is also possible to leave the
130 Sound Insulation in Buildings
60
50
40
Transmission loss, R (dB)
30
20
10
0
100 200 500 1000 2000 5000 10000
Frequency (Hz)
No slit b = 1 mm b = 3 mm b = 10 mm
Figure 5.24 Transmission loss of a 1 m2 wall with a 1 m long slit. The depth is 75 mm and
the width of the slit varies from 0 mm to 10 mm. (Adapted from Gösele, K.,
Schalldämmung von Türen, Berichte aus der Bauforschung, Heft 63, 1969,
Figure 17.)
slit under the door open, if there is an efficient sound absorbing lining in
the door, like a resonator. In the measured result shown in Figure 5.26, the
underside of the door has an opening into a 100 mm deep cavity filled with
mineral wool.
Sound insulation of doors is also particularly difficult in practice due to
the risk of leaks and because the sealing between the frame and the sur-
rounding wall is important. Good workmanship is important here.
Openable windows are also an issue, where leaks can easily lead to poor
sound insulation. It is important that a soft sealing list is applied on all
sides of the sash, and if it is possible to have two sets of continuous sealing
that is even better that a single sealing (Figure 5.27).
Introduction to sound insulation 131
60
50
40
Transmission loss, R (dB)
30
20
10
0
100 200 500 1000 2000 5000 10000
Frequency (Hz)
3 mm slit with felt 3 mm slit
Figure 5.25 Effect of filling a slit with felt. Transmission loss of a 1 m2 wall with a 1 m
long and 75 mm deep slit. The width of the slit is 3 mm. (Adapted from
Gösele, K., Schalldämmung von Türen, Berichte aus der Bauforschung, Heft
63, 1969, Figure 24.)
5.5 FLANKING TRANSMISSION
P1 S
R′ = 10 lg = L1 − L2 + 10 lg (dB) (5.53)
P2 + P3 A2
132 Sound Insulation in Buildings
60
50
40
Transmission loss, R (dB)
30
20
10
0
100 200 500 1000 2000 5000 10000
Frequency (Hz)
2 mm slit with absorber 2 mm slit without absorber
P3 = ∑P
i
F ,i (5.54)
P1
RF ,i = 10 lg (dB) (5.55)
PF ,i
Introduction to sound insulation 133
60
50
40
Transmission loss, R (dB)
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 5.27 Effect of sealing in an openable double window with two panes of 4 mm glass
and 85 mm cavity depth. (Adapted from Ingemansson, S., Ljudisolerande
fönsterkonstruktioner (In Swedish), Stockholm, Sweden, 1968, Figure 6.)
R′ = −10 lg 10−0,1R +
∑ 10
i
−0,1RF ,i
(dB) (5.56)
Figure 5.28 Direct transmission and three flanking transmission paths via the floor.
(Reproduced from Kristensen, J. and Rindel, J.H., Bygningsakustik – Teorti og
praksis. SBI-Anvisning 166 (In Danish), Danish Building Research Institute,
Hørsholm, Denmark, 1989. With permission.)
Figure 5.30 Examples of junctions between roof and internal wall (upper part) and
between floor and internal wall (lower part) in vertical view. Solutions are
shown in pairs, those with a flanking transmission issue to the left.
• Covering the flanking surface with a board having a high critical fre-
quency, i.e. preferably above 2000 Hz, the higher the better. This can
be a secondary surface with one or two layers of plasterboard. The
reduction of flanking transmission is efficient on both surfaces of the
flanking structure, in the source room as well as in the receiving room.
5.6 ENCLOSURES
A noise source is supposed to radiate sound power Pa. Consider that the
noise source is totally covered by an enclosure with surface area S, absorp-
tion coefficient α on the inside and the enclosure is made from a plate
with transmission loss R or transmission coefficient τ. The average sound
pressure in the enclosure pencl can be estimated, if a diffuse sound field is
assumed:
4Pa
2
pencl = ρ c (5.57)
αS
The sound power incident on the inner surface of the enclosure is (still with
the assumption of a diffuse sound field):
2
pencl S
Pinc = (5.58)
4ρ c
136 Sound Insulation in Buildings
τ
Pout = τ Pinc = Pa (5.59)
α
The insertion loss of the enclosure is the difference in radiated sound power
level without and with the enclosure:
Pa α
∆L = 10 lg = 10 lg = R + 10 lg α (dB) (5.60)
Pout τ
This result cannot be considered to be very accurate. The assumption of
a diffuse sound field inside the enclosure is especially doubtful. However,
the result is not bad as a rough estimate for the design of an enclosure. It is
clearly seen from Equation 5.60 that both transmission loss and absorption
coefficient are important for an efficient reduction of noise by an enclosure.
The tapping machine is placed on the floor in the source room in a number
of positions. The calibrated sound pressure level L 2 is measured in the room
below – or any other room in the building (see Figure 5.31). The reverbera-
tion time in the receiving room must also be measured in order to calculate
the absorption area A 2 . The impact sound pressure level is the sound pres-
sure level in dB re 20 μPa that would be measured if the absorption area is
A 0 = 10 m 2:
A2
Ln = L2 + 10 lg (dB) A0 = 10 m 2 (5.61)
A0
The frequency range is the same as for airborne sound insulation, i.e. the
16 one-third octave bands from 100 Hz to 3150 Hz. However, it is recom-
mended that the frequency range be extended down to 50 Hz, especially in
the case of lightweight floors and heavyweight constructions with a floating
floor.
Introduction to sound insulation 137
Room 1
P2
Room 2
A2
L2
Figure 5.31 Principle of measuring the impact sound pressure level from a floor to a
receiving room (Room 2).
T2
Ln,′ T = L2 − 10 lg (dB) T0 = 0.5 s (5.62)
T0
Just like DnT for airborne sound insulation, Ln′ T is corrected for the
reverberation time in the receiving room if different from 0.5 s. According
to ISO 16283-2, Ln′ T provides a straightforward link to the subjective
impression of impact sound insulation.
The relationship between the standardized and the normalized impact
sound pressure level is approximately:
V2
Ln′ T = Ln′ − 10 lg (dB), V0 = 30 m3 (5.63)
V0
where V2 is the volume of the receiving room. For typical volumes in a
dwelling around 50 m3 to 80 m3, the difference will be −2 dB to −4 dB.
138 Sound Insulation in Buildings
5.7.2 Historical note
The tapping machine originated from the German standard DIN 4110
(1938), but the impact sound pressure level was measured in Phon as a
broad-band level applying a frequency weighting filter B. The practice to
measure impact sound in one-third octave bands was developed in Denmark
(Ingerslev et al. 1947) – they used the frequency range from 100 Hz
to 2000 Hz. An alternative method was developed in America, using a tap-
ping machine with different specifications and measuring in seven octave
bands from 63 Hz to 4 kHz (Lindahl and Sabine 1940).
90
80
Impact sound pressure level, Ln (dB)
70
ΔL
60
50
40
30
20
10
50 100 200 500 1000 2000 5000
Frequency (Hz)
Bare slab With wooden floor on laths on soft pads
Figure 5.32 Impact sound pressure level of a 200 mm clinker concrete slab with 30 mm
concrete finish, measured without and with a 22 mm wooden floor on
laths (38 mm × 57 mm) on 10 mm fiberboard pads. (Data from Lydteknisk
Institut, Building Acoustic Laboratory Measurements. Bygningsakustiske labora-
toriemålinger, in Danish), Lyngby, Denmark, LL 906/80.
Introduction to sound insulation 139
R + Ln = 38 + 30 lg f (5.64)
With a floating floor or a soft flooring, the relationship changes above the
resonance frequency f 0 of the flooring:
R + Ln = 38 + 30 lg f0 − ∆L + ∆R (f > f0 ) (5.65)
80
70
ΔR
60
Transmission loss, R (dB)
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Bare slab With wooden floor on laths on soft pads
160
150
140
130
Sum R + Ln (dB)
ΔL – ΔR
120
110
100
90
80
50 100 200 500 1000 2000 5000
Frequency (Hz)
Bare slab With wooden floor on laths on soft pads Theory
Figure 5.34 Example of sum curves obtained by adding R and L n, using the measured
results from Figures 5.32 and 5.33.
5.8 SINGLE-NUMBER RATING OF
SOUND INSULATION
80
70
60
Transmission loss, R (dB)
Rw = 30 dB
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 5.35 Determination of the weighted sound reduction index. Full line is the mea-
sured transmission loss, in this example, a double window measured in the
laboratory. The dotted curve is the reference curve, which has been shifted
to the position 30 dB at 500 Hz. The sum of the unfavourable deviations is
7.0 + 8.0 + 8.0 + 4.0 + 2.0 + 3.0 = 32.0 dB. The result is Rw = 30 dB.
142 Sound Insulation in Buildings
• The frequency range is 125 Hz to 4000 Hz, i.e. there are still 16 one-
third octave bands but they are shifted one band up; the 100 Hz value
of the reference curve is not used and the reference curve is extended
horizontally up to 4000 Hz.
• The measured transmission loss results are rounded to whole dB val-
ues before comparison with the reference curve.
• An additional criterion is that the maximum unfavourable deviation
is 8 dB; in the case of a dip in the frequency curve of sound insulation,
this may be decisive for the single-number value.
• The result is written STC xx (without dB), where xx is the single-
number value in dB.
The Rw and STC values are usually very close, within 1 or 2 points, but
larger differences may occur in the case of a resonance dip.
80
L'n,w = 48 dB
70
60
Impact sound pressure level, L'n (dB)
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 5.36 Determination of the weighted impact sound pressure level. The full line is
the measured impact sound pressure level, in this example, a field measure-
ment of a concrete slab with a floating floor. The dotted curve is the refer-
ence curve, which has been shifted to the position 48 dB at 500 Hz. The sum
of the unfavourable deviations is 6.3 + 3.7 + 1.7 + 4.4 + 3.0 + 5.0 + 2.3 = 26.4 dB.
The result is L′n,w = 48 dB.
80
L'B = 60 dB
70
60
Impact sound pressure level, L'n (dB)
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 5.37 The Bodlund curve and its application. The full line is the measured impact
sound pressure level, in this example, a field measurement of a concrete
slab with a floating floor. The dotted curve is the reference curve, which has
been shifted to the position 60 dB at 500 Hz. The sum of the unfavourable
deviations is 3.3 + 10.9 + 11.7 + 3.3 = 29.2 dB. The result is L′B = 60 dB.
Introduction to sound insulation 145
It has been found that the result, using the Bodlund curve, is highly cor-
related with the ISO method when extended with the spectrum adaptation
term going down to 50 Hz (the spectrum adaptation term is explained in
Section 5.8.6). As an example, we find that more than 80 % should be
satisfied with the impact sound if L′B < 55 dB, or approximately L′n,w + C I,50–
2500 < 49 dB, because there is an average difference of 6.4 dB between the
two ratings (Hveem et al. 1996, 35).
5.8.5 Historical note
It is interesting to note that ISO/R 717, which is specifically meant for mea-
surements in dwellings, only mentions the normalized measures for sound
insulation R′ and L n, not the standardized measures that adjust the result to
0.5 s reverberation time. Since some countries were using the standardized
measures, it is not surprising that there was some controversy about the
proposal among the ISO member counties; 20 countries approved the rec-
ommendation, but six countries opposed the approval (Belgium, Denmark,
France, Italy, Norway and USA).
146 Sound Insulation in Buildings
C j = −10 lg ∑ 10
i =1
(Lij − Xi )/10
− Xw (dB) (5.67)
where
Introduction to sound insulation 147
Table 5.1 S pectra for calculation of the spectrum adaptation terms according to
ISO 717-1
Spectrum 1 Spectrum 1 Spectrum 2
From 50 Hz or From 50 Hz or
100 Hz up to 100 Hz up to Any frequency
3150 Hz 5000 Hz range
Frequency
(Hz) Lj,i (dB) Lj,i (dB) Lj,i (dB)
50 −40 −41 −25
63 −36 −37 −23
80 −33 −34 −21
100 −29 −30 −20
125 −26 −27 −20
160 −23 −24 −18
200 −21 −22 −16
250 −19 −20 −15
315 −17 −18 −14
400 −15 −16 −13
500 −13 −14 −12
630 −12 −13 −11
800 −11 −12 −9
1000 −10 −11 −8
1250 −9 −10 −9
1600 −9 −10 −10
2000 −9 −10 −11
2500 −9 −10 −13
3150 −9 −10 −15
4000 −10 −16
5000 −10 −18
Note: Spectrum 1 is A-weighted pink noise and spectrum 2 is A-weighted traffic noise.
148 Sound Insulation in Buildings
Note that there is no need to calculate the weighted value with the shifted
reference curve if only the extended frequency range is relevant. For
example, Rw,50 is calculated directly from:
n
CI = 10 lg ∑10
i =1
Li /10
− 15 − Xw (dB) (5.69)
Introduction to sound insulation 149
–10
–20
–30
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 5.38 Traffic noise spectra in third octave bands relative to the A-weighted level.
The ISO 717-1 spectrum used for the spectrum adaptation term Ctr, the
ASTM E1332 spectrum used for OITC, and the city street traffic spectrum
from Buratti et al. (2014).
where
C I is the spectrum adaptation term; C I, 50–2500 in the case of extended
frequency range.
i is index for the one-third octave band in the relevant range, up to n.
Li is the level of the measured impact sound pressure level in the fre-
quency band i, given to 0.1 dB.
Xw is the single-number quantity calculated with the shifted reference
curve, e.g. L n,w or Ln′ T ,w .
The number 15 in Equation 5.69 has been so determined that the value of
C I is approximately 0 dB for solid floors with effective floor coverings.
For simplicity, L n,w,50 is sometimes used instead of the more correct
L n,w + C I, 50–2500 and, similarly, for Ln′ T ,w,50 . Note that there is no need to
calculate the weighted value with the shifted reference curve if only the
extended frequency range is relevant. For example, L n,w,50 is calculated
directly from:
Ln,w,50 = 10 lg ∑10
i =1
Li /10
− 15 (dB)
(5.70)
150 Sound Insulation in Buildings
Requirements for sound insulation are usually given in the national build-
ing codes and may differ from one country to another. Some countries
have detailed and complicated requirements for different types of buildings
and rooms, while other countries have few and simple requirements. Some
examples of requirements between dwellings are shown in Table 5.2. The
highest level of sound insulation requirements is found in the Nordic coun-
tries and in Austria, while Portugal has significantly lower requirements.
The latter is characteristic for countries in southern Europe. It appears that
sound insulation requirements depend on the climate of the country, and
a warm climate means that people spend most time with open windows or
outdoors, while the opposite is true in a cold climate.
In 2003, the UK changed the compliance limits for England and Wales
to use the traffic noise spectrum adaptation term C tr for the evaluation of
party walls and floors; Australia did the same a few years later. The reason-
ing behind this rather surprising adoption was that low-frequency sound
from music, Hi-Fi equipment, etc. was recognized as an increasing problem
between neighbours in multi-unit buildings, but there was at the same time
a strong resistance against extending the measurements to include frequen-
cies below 100 Hz. The traffic noise spectrum would certainly put more
weight on the lower frequencies (e.g. 100 Hz to 500 Hz), but instead new
problems have appeared with complaints about insufficient sound insula-
tion at the mid and high frequencies.
In the USA, most of the states have adapted the International Building
Code (IBC), which requires STC 50 and IIC 50. However, this is the require-
ment for laboratory tests of the building components, and in the case of
a field test, 5 dB lower values are allowed (field sound transmission class
(FSTC) and field impact insulation class (FIIC)). It should be noted that the
American measurement standard ASTM 336 requires that ‘the flanking
paths be shielded’ during measurement of transmission loss in the field.
Therefore, FSTC 45 is a rather weak requirement for airborne sound insu-
lation. Canada has no requirement for impact sound and the requirement
for airborne sound insulation is similar to that in the USA, except that field
measurements should be made in terms of the apparent sound transmission
class (ASTC), which includes the flanking paths in the building.
The progress of lightweight building systems has led to increased aware-
ness of the importance of low frequencies for evaluation of impact sound,
and as a consequence, a revised building code SS 25267 was introduced in
Sweden in 2015. This requires, as a minimum, the sound insulation to be
evaluated from 50 Hz, and impact sound to be measured down to 20 Hz in
case of sound classes above the minimum requirement.
REFERENCES
ASTM E90-09 (2009). Standard test method for laboratory measurement of air-
borne sound transmission loss of building partitions and elements, American
Society for Testing and Materials, West Conshohocken, PA.
ASTM E336-16 (2016). Standard test method for measurement of airborne sound
attenuation between rooms in buildings, American Society for Testing and
Materials, West Conshohocken, PA.
ASTM E413-16 (2016). Classification for rating sound insulation, American Society
for Testing and Materials, West Conshohocken, PA.
ASTM E492-09 (2016). Standard test method for laboratory measurement of impact
sound transmission through floor-ceiling assemblies using the tapping machine.
American Society for Testing and Materials, West Conshohocken, PA.
ASTM E989-06 (2012). Standard classification for determination of impact
insulation Class (IIC), American Society for Testing and Materials, West
Conshohocken, PA.
ASTM E1007-16 (2016). Standard test method for field measurement of tapping
machine impact sound transmission through floor-ceiling assemblies and asso-
ciated support structures, American Society for Testing and Materials, West
Conshohocken, PA.
ASTM E1332-16 (2016). Standard classification for rating outdoor-indoor sound atten-
uation, American Society for Testing and Materials, West Conshohocken, PA.
R. Berger (1911). Über die schalldurchlässigkeit. (On the Sound Transmission, in
German). Dissertation, Munich, Germany.
K. Bodlund (1985). Alternative reference curves for evaluation of the impact sound
insulation between dwellings. Journal of Sound and Vibration 102, 381–402.
152 Sound Insulation in Buildings
J.S. Bradley (1982). Subjective Rating of the Sound Insulation of Party Walls.
Building Research Note 196. National Research Council Canada, Division of
Building Research.
O. Brandt (1958). Akustisk planering. (Acoustical planning, in Swedish). V. Pettersons
Bokindustri AB, Stockholm, Sweden.
C. Buratti, E. Belloni, E. Moretti (2014). Façade noise abatement prediction: New
spectrum adaptation terms measured in field in different road and railway traf-
fic conditions. Applied Acoustics 76, 238–248.
L. Cremer (1942): Theorie der Schalldämmung dünner Wände bei schrägem Einfall.
(Theory of sound insulation of thin walls at oblique incidence, in German).
Akustische Zeitschrift 7, 81–104.
J. Davy (2004). Insulating buildings against transportation noise. Proceedings of
ACOUSTICS 2004, 3–5 November 2004, Gold Coast, Australia, 447–453.
DIN 4109 (1959). Schallschutz im Hochbau, Entwurf. (In German). Berlin 1959.
(Final edition 1962).
DIN 4110 (1938). Technische Bestimmung für die Zulassung neuer Bauweisen. (In
German), 2nd edition. Berlin.
W. Fasold, E. Sonntag (1978). Bauakustik. Bauphysikalisches Entwurfslehre, Band 4,
3rd edition. VEB Verlag für Bauwesen. Berlin.
K. Gösele (1969). Schalldämmung von Türen (Sound insulation of doors, in German).
Berichte aus der Bauforschung, Heft 63.
S. Hveem, A. Homb, K. Hagberg, and J.H. Rindel (1996). Low-frequency footfall
noise in multi-storey timber frame buildings. NKB Work and Committee
Report 1996:12 E. Nordic Committee on Building Regulations, Acoustics
Group, Helsinki.
S. Ingemansson (1968). Ljudisolerande fönsterkonstruktioner. (Sound insulating
window constructions, in Swedish). Report from the Building Research, 3/68,
Stockholm, Sweden.
F. Ingerslev, A. Kjerbye Nielsen, and A. Falck Larsen (1947). The measuring of
impact sound transmission through floors. Journal of the Acoustical Society of
America 19, 981–987.
ISO 10140-2 (2010). Acoustics – Laboratory measurement of sound insulation
of building elements – Part 2: Measurement of airborne sound insulation,
International Organization for Standardization, Geneva, Switzerland.
ISO 10140-3 (2010). Acoustics – Laboratory measurement of sound insulation
of building elements – Part 3: Measurement of impact sound insulation,
International Organization for Standardization, Geneva, Switzerland.
ISO 16283-1 (2014). Acoustics – Field measurements of sound insulation in build-
ings and of building elements – Part 1: Airborne sound insulation, International
Organization for Standardization, Geneva, Switzerland.
ISO 16283-2 (2015). Acoustics – Field measurements of sound insulation in build-
ings and of building elements – Part 2: Impact sound insulation, International
Organization for Standardization, Geneva, Switzerland.
ISO 16283-3 (2016). Acoustics – Field measurements of sound insulation in build-
ings and of building elements – Part 3: Façade sound insulation, International
Organization for Standardization, Geneva, Switzerland.
ISO 717-1 (1996). Acoustics – Rating of sound insulation in buildings and of build-
ing elements – Part 1: Airborne sound insulation, International Organization
for Standardization, Geneva, Switzerland.
Introduction to sound insulation 153
ISO 717-2 (1996). Acoustics – Rating of sound insulation in buildings and of build-
ing elements – Part 2: Impact sound insulation, International Organization for
Standardization, Geneva, Switzerland.
ISO/R717 (1968). Rating of sound insulation for dwellings, International
Organization for Standardization, Geneva, Switzerland.
J. Kristensen (1973). Lydisolation: Teori, måling, vurdering og bestemmelser, (Sound
insulation: Theory, measurement, evaluation and requirements, in Danish). SBI-
notat 24 (In Danish). Danish Building Research Institute, Aalborg University
Copenhagen.
J. Kristensen, J.H. Rindel (1989). Bygningsakustik – Teori og praksis (Building
acoustics – Theory and practice, in Danish) SBI-Anvisning 166. Danish
Building Research Institute, Hørsholm, Denmark.
P.T. Lewis (1971). Real windows, Chapter 7, in Building Acoustics (Eds. T. Smith,
P.E. O’Sullival, B. Oakes, R.B. Conn). British Acoustical Society, Special Volume
No. 2, Oriel Press, London.
R. Lindahl and H.J. Sabine (1940). Measurement of impact sound transmission
through floors. Journal of the Acoustical Society of America 11, 401–405.
J.A. Marsh (1971). The airborne sound insulation of glass, Chapter 5, in Building
Acoustics (Eds. T. Smith, P.E. O’Sullival, B. Oakes, R.B. Conn). British
Acoustical Society, Special Volume No. 2, Oriel Press, London.
NT ACOU 061 (1987). Windows: Traffic Noise Reduction Indices, Nordtest
Method. Helsinki, Finland.
SS 25267 (2015). Acoustics – Sound Classification of Spaces in Buildings, Swedish
Standards Institute, Stockholm.
Lydteknisk Institut. Building Acoustic Laboratory Measurements (Bygningsakustiske
laboratoriemålinger, in Danish). Lyngby, Denmark.
M. Mesihovic, J.H. Rindel, and I. Milford (2016). The need for updated traffic noise
spectra, used for calculation of sound insulation of windows and facades.
Proceedings of InterNoise 2016, Hamburg, Germany, 3890–3897.
Chapter 6
The sound radiation from plates is described in this chapter. Both forced and
resonant vibrations are considered, and the application of Rayleigh’s method
of radiation calculation is demonstrated in the case of radiation from forced
bending waves. The influence of the finite area of a plate is also shown. Forced
vibrations are the part of the vibrations that are directly due to the surround-
ing sound field exciting the plate. In contrast, resonant vibrations are the free
vibrations caused by reflections of the forced vibrations from the boundaries.
155
156 Sound Insulation in Buildings
are looking for are, therefore, to have zero value at the boundary, and the
second-order derivative should also be zero at this location. A function
fulfilling these conditions is sin (kxx) (convince yourself). It is also easy
to show that this function fulfils the bending wave equation (once again,
convince yourself). These two facts prove that the normal modes we are
looking for are sine functions. The wave number k is, however, still not
fixed. If there is a simple support both at x = 0 and x = lx, then kx must equal
mπ
kx, m = (6.1)
lx
nπ
ky, n = (6.2)
ly
mπ nπ
φm, n (x, y) = sin x sin y (6.3)
lx ly
w(x, y) = ∑ Aφ
m , n =1
n m, n (x, y) (6.4)
However, that is not what we shall do here. Instead, the starting point will
be the Pythagoras relationship between the free bending wave numbers and
the mode numbers m and n,
2 2
mπ nπ
kb2 = km2 , x + kn2, y = + (6.5)
lx ly
If the bending wavelength is used instead of the wave numbers, the relation-
ship that shall be fulfilled is
2 2 2
2 m n
λ = l + l (6.6)
B x y
Sound radiation from plates 157
λ/2
λy/2 ly
lx
λx/2
where, as before, m and n are natural numbers (Figure 6.1). The numbers
represent the number of half wavelengths along the x- and y-axis, respec-
tively. Some examples are shown in Figure 6.2. With simply supported
edges, the lowest natural frequency is the one with (m, n) = (1, 1).
In general, it is a little complicated to calculate the natural frequen-
cies of the normal modes, because of the transition from bending to shear
(a) (b)
(c) (d)
c2 m 2 n 2
fmn = + (6.7)
4 fc lx ly
1 1 1
N = π ⋅ lx ⋅ ly − ( lx + ly ) + (6.8)
λB
2
λB 4
Assuming bending waves and introducing the area S = lx∙ly and the perim-
eter of the panel U = 2(lx + ly), the modal density yields:
∆N π S U fc π S (6.9)
= 2 fc − ≅ 2 fc
∆f c 4c f c
∆N 2π S U 2π S (6.10)
= 2 f− ≅ 2 f
∆f cs 2c s cs
We notice that the modal density in a plate is constant in the bending wave
region, but increases with frequency in the shear wave region.
The modal density of a homogeneous, isotropic plate depends not only on
the length and width but also on the thickness. If the plate is thin (thickness
less than approximately one-sixth of the wavelength of bending waves),
bending waves can be assumed; but if the plate is thick compared to the
wavelength, shear waves will occur instead. If the bandwidth is one-third
octave, Δf = 0.23 f and, thus, the approximate number of modes within the
band is:
0.23π
2
Sfc f (f < f s / 2) (bending)
c
∆N1/3 ≅ (6.11)
0.46π Sf 2 (f > f s / 2) (shear)
c s2
Sound radiation from plates 159
Here fc is the critical frequency, fs = fc (cs /c)2 is the cross-over frequency for
transition to shear waves and cs is the speed of shear waves. The two for-
mulas in Equation 6.11 are shown graphically in Figure 6.3. The curves
cross at the frequency fs /2, so this is the cross-over frequency from bending
to shear when the modal density is considered. In other words, the cross-
over for modal density is one octave lower than the cross-over for speed of
propagation.
As an example, the first modes have been calculated for a 10 m 2 brick
wall, 250 mm thick (Table 6.1). The actual number of modes in each one-
third octave band is displayed in Figure 6.4 and compared with the sta-
tistical estimate (Equation 6.11). The shear wave cross-over frequency is
fs = 4 kHz, so for modal density, the bending wave assumption holds up
to 2 kHz. It is noted that up to 250 Hz, many one-third octave bands are
without any modes and the first modes from f 11 to f 22 are widely separated.
Actually, this interval is two octaves (f 22 = 4 f11) and there may be only
two or three modes within that range. Thus, the statistical estimates in
Equation 6.11 should only be used for frequencies above f 22 = 4 f11, that is
366 Hz in this example.
100.0
Bending
Number of modes per one-third octave band
Shear
10.0
1.0
1/2 fs
0.1
50 100 200 500 1,000 2,000 5,000 10,000
Frequency (Hz)
Figure 6.3 Number of normal modes per one-third octave bands in a 250 mm thick
brick wall, calculated for bending waves and shear waves with Equation 6.11.
fc = 72 Hz, fs ≈ 4000 Hz, f11 = 91.4 Hz. Below 315 Hz, there is statistically less
than one mode per band. Note that the ordinate is logarithmic.
160 Sound Insulation in Buildings
Table 6.1 F
irst 14 modes calculated for a 250 mm
thick brick wall, 4 m × 2.5 m
m n fmn (Hz)
1 1 91.4
2 1 168
1 2 289
3 1 297
2 2 366
4 1 477
3 2 494
1 3 617
4 2 674
2 3 694
5 1 708
3 3 823
5 2 905
6 1 990
Number of modes per one-third octave band
7 Actual number
Statistical
6
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 6.4 Number of normal modes per one-third octave bands in a 250 mm thick brick
wall, 4 m × 2.5 m. The first normal mode is f11 = 91.4 Hz. The statistical modal density
calculated from Equation 6.11 is also shown. Bending waves only.
Consider a large plate located in the xy-plane with sound radiation into
the side where z > 0. A bending wave is assumed to travel in the plate as a
Sound radiation from plates 161
plane wave in the x-direction. Hence, the velocity of the vibration can be
written as:
v = v e− jkBx (6.12)
The time factor ejωt is omitted here and in the following. The angular wave
number of the bending wave is k B = ω/cB . In the air above the plate, there
is no variation in the y-direction, so the sound pressure can be written in
the form:
where kx and kz are the wave numbers that represent the sound propaga-
tion in the x- and z-directions, respectively. When this general solution is
inserted in the Helmholtz equation:
∇2 p + k2 p = 0 (6.14)
1 ∂p p kz − j(kx x+kz z)
uz = − = e (6.16)
jωρ ∂z ωρ
At the surface of the plate (z = 0), this must equal the normal velocity of the
plate (Equation 6.12) and thus:
p kz
v= and kx = kB (6.17)
ωρ
kz = k2 − kB2 (6.18)
Insertion in Equation 6.13 yields the following result for the sound pressure:
ρc v − j k2 −kB2 z
p= e− jkBxe (6.19)
1 − (kB / k)2
The first case is k B < k (meaning that cB > c and f > fc). The solution rep-
resents a sound field propagation in both the x- and z-directions. The angle
of propagation (Figure 6.5c) is determined by the following:
c k fc
sin θ = = B = (6.20)
cB k f
In this case, the sound pressure can be written in terms of the angle of
radiation, θ:
The second case is k B > k (meaning that cB < c and f < fc). The sound pres-
sure (Equation 6.19) yields:
jρ c v − j kB2 − k2
p= e− jkB xe z (6.22)
2
(kB / k) − 1
This represents a sound field that only propagates in the x-direction, but
not in the z-direction; the amplitude of the sound pressure decreases expo-
nentially with the distance from the plate. There is an acoustic near field,
but no sound radiation (Figure 6.5a).
–
θ
+ CB CB
CB C
CB
C
–
–
f < fc f = fc f > fc
Figure 6.5 A propagating bending wave in a plate cannot radiate sound if the speed
is below c (case a). If the speed equals c, the sound is radiated along the
surface of the plate (case b). A plane wave is radiated in the direction θ if
the speed exceeds c (case c). (Reproduced from Kristensen, J., Rindel J.H.,
Bygningsakustik – Teori og praksis (Building acoustics – Theory and practice,
in Danish). SBI Anvisning 166. Danish Building Research Institute, Hørsholm,
Denmark, 1989. With permission.)
Sound radiation from plates 163
(a) (b)
Figure 6.6 Simulated particle movements in the air due to a bending wave travelling from
left to right. (a) cB < c. (b) cB > c. (Reproduced from Heckl, M. and Heckl,
M., Darstellung von abstrahlvorgängen im film. Fortschritte der Akustik – DAGA
1976, VDI-Verlag, Düsseldorf, 165–168, 1976. With permission.)
The result in the previous section is for structures with infinite extent.
However, real structures are finite and this has an important effect on the
sound radiation. The rest of this chapter is about finite structures. In order
to study the radiation of these structures, radiation efficiency is introduced.
The concept of radiation efficiency was first introduced by Gösele (1953).
Radiation efficiency is a measure of the radiated sound power from a source
relative to the sound power radiated from an equivalent piston source that
generates a plane wave. The sound power from a very large piston with a
harmonic vibration v = |v| ejωt and the area S is:
2
Ppiston = 1
2 ρ cS v = ρ cS v 2 (6.23)
Here v denotes the RMS (root mean squared) value of the velocity, and the
brackets mean the average taken over the surface of the vibrating source.
If the actual radiated sound power is P from a surface with area S and
velocity v, the radiation efficiency σ is defined as:
P
σ = (6.24)
ρ cS v 2
164 Sound Insulation in Buildings
The radiation efficiency has no dimension and it may take values between
0 and around 4. Values higher than unity are possible because bending waves
in a plate may radiate more efficiently than the piston, as we shall see later.
The radiation efficiency will be used to describe the radiated sound power
from various kinds of vibrations. When sound is transmitted through a
plate, the radiated sound power can be assigned to two different kinds of
vibrations in the same plate: forced and resonant vibrations. As these two
vibration fields have different physical origins and different wave numbers
associated with them, the radiation efficiency and velocity magnitude will
also be very different. The resonant vibration is associated with free bend-
ing wave numbers and the forced vibration is associated with the wave
number in the sound field in the air. The total transmission through the
wall is found by adding together the transmitted sound powers:
In this section, a general method for calculating the sound radiation from a
vibrating surface is introduced – the so-called Rayleigh’s method to be used
in the following sections. (Rayleigh, 1896, §278, Equation 3). The method
is described in Lord Rayleigh’s “The Theory of Sound” first p ublished 1877,
and the steps are the follows:
dS΄
β0 r
dS ro α0
Iy X
2
Iy θr
2
Ix Ix
2 2
Z
vdS − jkr
p = jωρ e (6.26)
2π r
Here and in the following, the time factor e− jω t is omitted. The total sound
pressure at the point of observation can be found by integration over the
surface of the plate:
∫ ver
− jkr
jωρ
p= dS (6.27)
2π S
The distance from the centre of the plate to the point of observation is r 0
and it is assumed that this is large compared to the dimensions of the plate.
In Equation 6.27, r appears in two places – in the divisor, the approxima-
tion r ≈ r 0 is sufficient, whereas in the exponent, a more accurate approxi-
mation is needed:
Here x and y are the coordinates of the source point dS, and the angles α 0
and β 0 are as defined in Figure 6.7. With these approximations, the equa-
tion becomes:
p=
jωρ − jkr0
2π r0
e ∫ ve
S
jk( x cos α0 + y cos β0 )
dS (6.29)
With this equation, the sound pressure can be calculated in any direction
(α 0, β 0) if the velocity v of the plate is known as a function of the position
(x, y).
2
The radiated sound intensity at the point of observation is I = 12 p ρ c,
and the total sound power radiated into the half sphere S′ = 2π r 0 is:
2
P= ∫ S'
IdS′ =
r02
2ρc ∫ 2π
2
p dψ (6.30)
166 Sound Insulation in Buildings
We are now ready to study sound radiation and the radiation efficiency of a
finite plate. The theoretical derivation and experimental verification of the
results were done by Rindel (1975). To start with, consider the case of forced
vibrations (the resonant vibration case will be discussed in Section 6.6).
It is assumed that the vibrations of the plate are generated by a plane
sound wave incident from the side z < 0. The angles of incidence relative to
the axis of the coordinate system are α, β and θ, respectively (Figure 6.8).
The sound pressure of the incident sound can be described as:
1l 1l
2 y 2 x
p=
jωρ
2π r0
v e− jkr0 ∫ ∫e jkx(cos α0 −cos α ) jky (cos β0 −cos β )
e dx dy
− 12 ly − 12 lx
Even if this result looks complicated, we note that the variables are sepa-
rated. With the use of Euler’s formula (Equation 1.40), the result can be
rewritten in the simple form:
jωρ
= v Se− jkr0 ⋅ sinc(X) ⋅ sinc(Y ) (6.33)
2π r0
Sound radiation from plates 167
The function sinc (X) = sin X/X is symmetric (same value for positive and
negative X), it has a maximum of 1 for X = 0 and → 0 for X → ∞ (Figure 1.5).
This means that the sound radiation is maximum in the direction α 0 = α and
β 0 = β, i.e. in the same direction as the incident sound, as could be expected.
The result also shows that smaller amounts of sound are radiated in other
directions; especially for small values of klx and kly, i.e. if the plate dimen-
sions are small compared to the wavelength. If the ratio of dimension to
wavelength is very small, the radiation approaches an omnidirectional pat-
tern in a half sphere.
2S lxly
a= = (6.35)
U lx + ly
where S is the area and U is the perimeter. 2a is in fact the harmonic aver-
age of lx and ly (see Equation 1.33), and a must be between 0.5 and 1
times the shorter dimension of the plate. In the special case of a square
plate, a = 1 / 2 S . If the sides are very different in length, the characteristic
dimension does not exceed the shorter dimension.
α
θ
β
pi
Figure 6.8 Definition of the angles of incidence of the plane sound wave generating the
forced vibrations in the plate.
168 Sound Insulation in Buildings
(a)
1
0.9
0.8 ka = 1/2
Radiation function, sinc2 (X)
0.7 ka = 1
0.6 ka = 2
θ = 0° 0.5
ka = 4
0.4
ka = 8
0.3
ka = 16
0.2
ka = 32
0.1
0
–90 –60 –30 0 30 60 90
Angle of radiation, (θ0)
(b)
1
0.9
0.8 ka = 1/2
Radiation function, sinc2 (X)
0.7 ka = 1
0.6 ka = 2
θ = 45° 0.5
ka = 4
0.4
ka = 8
0.3
ka = 16
0.2
ka = 32
0.1
0
–90 –60 –30 0 30 60 90
(c) Angle of radiation, (θ0)
1
0.9
ka = 1/2
Radiation function, sinc2 (X)
0.8
0.7 ka = 1
0.6 ka = 2
θ = 90° 0.5 ka = 4
0.4
ka = 8
0.3
ka = 16
0.2
ka = 32
0.1
0
–90 –60 –30 0 30 60 90
Angle of radiation, (θ0)
Figure 6.9 Radiation function sinc2(X) showing the sound radiation pattern from forced
bending waves. Three examples of the angle of incidence, (a) 0°, (b) 45°,
(c) 90°, are shown.
Sound radiation from plates 169
The radiation pattern is shown in Figure 6.9 for the two-dimensional case,
i.e. β0 = β = π/2. It is seen that the radiation has little directivity for low ka, but
gets more directive for high values of ka. In the case of grazing incidence,
θ = 90°, the radiation covers a relatively wide angle even for high values of ka.
The total radiated sound power into the half sphere z > 0 is found by
inserting Equation 6.33 in Equation 6.30:
P=
1
2ρc ∫ ω2ρ2 2 2
2 π 4π
2
v S (sincX ⋅ sincY ) dψ
2
=
ρc 2 2 2
8π 2
⋅k S v ∫ 2π
(sincX ⋅ sincY )
2
dψ
(6.36)
σθ =
1
2 ρcS v
P
2
=
k2 S
4π 2 ∫ 2π
(sincX ⋅ sincY )
2
dψ (6.37)
Table 6.2 Calculated radiation efficiency for forced vibrations in a square plate
ka 0° 15° 30° 45° 60° 75° 90° Random
0.5 −8.3 −8.3 −8.3 −8.4 −8.5 −8.6 −8.6 −8.50
0.75 −5.0 −5.1 −5.2 −5.4 −5.6 −5.7 −5.7 −5.53
1 −3.0 −3.0 −3.3 −3.6 −3.9 −4.1 −4.2 −3.79
1.5 −0.6 −0.8 −1.1 −1.7 −2.2 −2.6 −2.7 −2.01
2 0.3 0.2 −0.2 −0.8 −1.3 −1.8 −2.0 −1.15
3 −0.1 0.2 0.5 0.4 −0.2 −0.7 −0.9 −0.14
4 −0.3 0.0 0.7 1.0 0.6 0.1 −0.2 0.41
6 0.0 0.2 0.6 1.6 1.7 1.1 0.7 1.16
8 0.1 0.2 0.5 1.6 2.4 1.8 1.4 1.61
12 0.0 0.2 0.6 1.4 3.1 2.9 2.3 2.18
16 0.0 0.2 0.6 1.5 3.2 3.7 3.0 2.57
24 0.0 0.2 0.6 1.5 2.9 4.7 3.9 3.04
32 0.0 0.2 0.6 1.5 3.0 5.4 4.6 3.36
48 0.0 0.2 0.6 1.5 3.0 6.0 5.5 3.77
64 0.0 0.2 0.6 1.5 3.0 6.1 6.2 4.04
Source: Sato, H., J. Acoust. Soc. Jpn., 29, 509–516, 1973.
170 Sound Insulation in Buildings
10
5
ka ∞
64
10 lg σθ (dB)
32
16
8
4
0
2
1
–5
0° 15° 30° 45° 60° 75° 90°
Angle of incidence (θ)
1
σθ → for ka → ∞ (6.38)
cos θ
Sound radiation from plates 171
10
Random
10 lg σθ (dB)
0 90°
60°
45°
0°
–5
–10
0.5 1 2 5 10 20 50 100
ka
Figure 6.11 Radiation efficiency for a square plate with forced bending waves for differ-
ent angles of incidence and for random incidence.
6.5.1 Approximate results
The following approximation to the result in Equation 6.37 was derived by
combining three asymptotic results for high frequencies, for low frequen-
cies and for grazing incidence θ = 90° (Rindel 1993).
−1/ 4
4
π π
2 2
π
σ for,θ ≅ cos θ − 0.6 + π 0.6 +
2
(6.39)
ka ka 2 ( ka )2
in Figure 6.11. A numerical approximation for a square plate and valid for
ka > 0.5 is:
1
σ for, d ≅
2
(0.20 + ln(2ka)) (6.40)
(excitation by a diffuse sound field, so the index d). Usually, Equation 6.40
is sufficiently accurate for an estimate. However, a more exact formula (still
for ka > 0.5) was derived by Sewell (1970):
σ for, d =
1
2( (
0.160 + ln(2ka) + 16π (ka)2 )
−1
)
+ C(Λ) (6.41)
Here C is a function of the edge ratio Λ = lx /ly, where lx ≥ ly. In the original
paper by Sewell (1970), C was given in a table for values of Λ between
1 and 10. The function can be approximated by:
0
10 lg σ (dB)
Λ=1
–5
Λ=2
Λ = 10
–10
0.1 1 10 100
ka
Figure 6.12 R
adiation efficiency for a plate with forced bending waves excited by a dif-
fuse sound field as a function of ka for three different edge ratios.
Sound radiation from plates 173
2
σ for ≅
π
(ka )2 (6.43)
This is valid for ka < 0.5 and for any angle of incidence including random
incidence.
10
5
ka ∞
64
10 lg σ (dB)
32
16
8
4
0
2
1
–5
1 2 5 10
(cB/c)2 ≈ f/fc
Figure 6.13 Radiation efficiency for a square plate as a function of the phase speed of
bending waves relative to the speed of sound in air.
174 Sound Insulation in Buildings
+ + –
+
– – +
(1,1) (1,2) (2,2)
Figure 6.14 Radiation from low-numbered modes, (a) (1,1), (b) (1,2) and (c) (2,2) in
a plate. The shaded areas are those radiating sound. Areas with + and − are
in opposite phase.
or above the critical frequency of the plate. If f < fc , we have cB < c, λB < λ,
and vice versa for f > fc.
At the first normal mode (1,1) of a simply supported plate (Figure 6.14a),
all parts of the plate move in phase and the sound radiation is similar to
that from the diaphragm of a loudspeaker. At the second and third normal
mode (1,2) or (2,1), the movements of the plate are divided by one nodal
line (Figure 6.14b). Below the critical frequency, the pressure variations
are cancelled across the modal line, but sound is radiated from a zone near
the edges. If we consider the next normal mode (2,2), the sound radiation
is further limited to four zones near the corners (Figure 6.14c). The radia-
tion efficiency of these first modes is displayed in Figure 6.15 as functions
of the relative speed cB /c. All modes have radiation efficiency σ ≅ 1 above
10
0
10 lg σm,n (dB)
–10
(1,1)
(1,2)
–20
(2,2)
–30
0.1 1
cB/c
Figure 6.15 Radiation efficiency for low-numbered modes, (1,1) (1,2) and (2,2). (Principle
adapted from Wallace, C.E., J. Acoust. Soc. Am., 51, 946–952, 1972, Figure 2.)
176 Sound Insulation in Buildings
the critical frequency (cB /c > 1); but at lower frequencies, the radiation is
less efficient. The sound radiation from normal modes in plates has been
analysed by Wallace (1972), but here it is sufficient to look at approximate
results for a square panel.
27 Sf 2 4Sf 2
σ res (1,1) = ⋅ ≅ 2 mode(1,1) (6.44a)
π 3 c2 c
25 S 2 f 4 3.4S 2 f 4
σ res (1, 2) = ⋅ ≅ mode(1,2) (6.44b)
3π c 4 c4
8π S3 f 6 1.7 S3 f 6
σ res (2, 2) ≅ ⋅ ≅ mode(2,2) (6.44c)
15 c 6 c6
Modes with higher mode numbers radiate from the edges or from the cor-
ners when the frequency is below the critical frequency, but from the entire
area above the critical frequency (Figures 6.16 and 6.17). There is cancel-
lation of sound pressures across the nodal lines of the plate as long as the
wavelength in the plate is shorter than the wavelength in the air. Therefore,
some mode patterns give radiation along two edges and other mode pat-
terns only allow the four corners to radiate sound. Above the critical fre-
quency, the entire surface contributes to the radiation and σ → 1 at high
frequencies.
The following result for the radiation efficiency was derived by Maidanik
(1962) for the case of resonant vibrations in a plate generated by random
incidence sound at frequencies below the critical frequency.
c2 Uc
σ res = 2
⋅ g1(M) + ⋅ g 2 (M) (f11 < f < fc ) (6.45)
Sfc Sfc
8(1 − 2M 2 )
(M < 2 / 2)
g1(M) = π 4 M 1 − M 2 (6.46)
0 (M ≥ 2 / 2)
10
C: Surface
0
10 lg σm,n (dB)
–10
B: Edges
–20
A: Corners
–30
0.1 1
cB/c
cB,eff f
M= ≈ (6.48)
c fc
The parameter M is the Mach number, as it gives the ratio between the
speed of the travelling bending wave and the speed of sound in air. The
approximation using the critical frequency is only valid for thin plates,
whereas a more general result is obtained by using the ratio between the
propagation speed in the plate and in the air, cB,eff /c.
178 Sound Insulation in Buildings
The first assumptions means in reality that the result is not valid below
f 22 = 4 f11 (see Section 6.1).
Above the critical frequency, we have radiation from supersonic vibra-
tions and for ka ≫ 1, the radiation efficiency is:
( )
−1/2
≅ (1 − fc f )
−1/2
σ res = 1 − M −2 (f > fc ) (6.49)
π fU
σ res ≅ (f ≈ fc ) (6.50)
16c
Combining the aforementioned equations into one:
10
4000
10 lg σ (dB)
2000
–10
1000
500
–20 250
125
–30
50 100 200 500 1000 2000 5000
Frequency (Hz)
where σ1, σ 2 and σ 3 are from Equations 6.45, 6.49 and 6.50, respectively. In
this way, the transition between each equation is smoothed.
Some calculation examples are shown in Figure 6.18. The previous obser-
vations concerning a peak near the critical frequency for thin plates and a
very flat radiation course for thick plates are clearly seen. For frequencies
above the critical frequency, the results shown in Figure 6.18 are in fact the
same as those in Figure 6.13.
6.7.1 Approximate results
Assuming bending waves, i.e. plates that are relatively thin and suffi-
ciently large so that the first mode of vibration f 11 ≪ fc , we get the following
approximations for the frequency range below the critical frequency (also
shown in Figure 6.19).
Uc f 3c 1
σ res ≅ < f < fc (6.52)
π 2 Sfc fc U 2
c2 3c
σ res ≅ f11 < f < (6.53)
Sfc2 U
4Sf 2
σ res ≅ (f < f11) (6.54)
c2
10
fc
fb
10 lg σ (dB)
–10
–20
Figure 6.19 Radiation efficiency from resonant vibrations in a large thin plate (f11 < fc).
The maximum radiation from Equation 6.50 is shown as a dashed line.
180 Sound Insulation in Buildings
10
fc
0
10 lg σ (dB)
–20
10 20 50 100 200 500 1,000 2,000 5,000 10,000
Frequency (Hz)
Figure 6.20 Radiation efficiency from resonant vibrations in a small thick plate (f11 > fc).
The maximum radiation from Equation 6.50 is shown as a dotted line.
Table 6.3 E
xamples of key frequencies for a 6.5 mm
glass pane and a 250 mm brick wall
Glass pane Brick wall
h, mm 6.5 250
lx , m 1.4 4
ly, m 1.4 2.5
S, m2 2 10
f11, Hz 15 91.5
f21, Hz 37 168
f12, Hz 37 289
f22, Hz 60 366
fa, Hz 89 43
3c/U, Hz 183 79
fc, Hz 2000 72
fb, Hz 2305 204
Sound radiation from plates 181
5c
fb = fc + (6.55)
U
At very low frequencies, the crossing of Equations 6.50 and 6.54 is at the
frequency fa:
πU
fa = 3
64c
( f11fc ) (6.56)
2
k2 S
3 (
sincX1 + sincX2 ) (6.57)
2
σ res,θ =
2π
where
1 1
X1 = klx ( sin θ r + c cB ) and X2 = klx ( sin θ r − c cB ) (6.58)
2 2
182 Sound Insulation in Buildings
50 50
Radiation efficiency,(σ)
Radiation efficiency,(σ)
40 40
30 30
20 20
10 10
0 0
–90 –75 –60 –45 –30 –15 0 15 30 45 60 75 90 –90 –75 –60 –45 –30 –15 0 15 30 45 60 75 90
Angle of radiation,(θr) Angle of radiation,(θr)
ka = 2π ka = 4π ka = 8π ka = 16π ka = 2π ka = 4π ka = 8π ka = 16π
(c) f = fc (d) f = 4/3 fc
60 60
50 50
Radiation efficiency,(σ)
Radiation efficiency,(σ)
40 40
30 30
20 20
10 10
0 0
–90 –75 –60 –45 –30 –15 0 15 30 45 60 75 90 –90 –75 –60 –45 –30 –15 0 15 30 45 60 75 90
Angle of radiation,(θr) Angle of radiation,(θr)
ka = 2π ka = 4π ka = 8π ka = 16π ka = 2π ka = 4π ka = 8π ka = 16π
50 50
Radiation efficiency,(σ)
Radiation efficiency,(σ)
40 40
30 30
20 20
10 10
0 0
–90 –75 –60 –45 –30 –15 0 15 30 45 60 75 90 –90 –75 –60 –45 –30 –15 0 15 30 45 60 75 90
Angle of radiation,(θr) Angle of radiation,(θr)
ka = 2π ka = 4π ka = 8π ka = 16π ka = 2π ka = 4π ka = 8π ka = 16π
It is seen from the result (Equation 6.57) that the radiation has two domi-
nating, symmetric lopes as long as cB > c (f > fc). The main direction of
radiation is:
θ r = ± arcsin(c cB ) = ± arcsin ( )
fc f (6.59)
(a) (b)
f < fc f = fc
(c) (d)
f = 2fc f = 4fc
Figure 6.22 Sound radiation from a panel with excitation by a plane wave at normal
incidence. (a) Below the critical frequency. (b) At the critical frequency. (c)
One octave above the critical frequency. (d) Two octaves above the critical
frequency. Dotted line arrows indicate resonant radiation.
1
θ r = arccos (6.60)
σθ
Sound radiation from plates 185
a θr
a θ
Figure 6.23 Angle of incidence θ and main direction of sound radiation θr for forced
vibrations in a finite plate with characteristic dimension a.
10
ka ∞
32
5
10 lg σθ (dB)
θr
–5
0° 15° 30° 45° 60° 75° 90°
Angle of incidence (θ)
Figure 6.24 Equivalent angle of radiation from a square panel is 70° when the angle of
incidence is 85° and ka = 32.
186 Sound Insulation in Buildings
Figure 6.24 shows the principle of this conversion for the angle of incidence
θ = 85° and ka = 32. The equivalent angle of radiation is found to be around
θr = 70°.
The simulations shown in Figure 6.25 are four examples of radiation
from resonant vibrations in a plate that is excited by an incident plane
wave. The normal modes radiate sound equally to both sides of a plate, and
the radiation is symmetric around the normal to the plate. In Figure 6.25a,
the mode with modal number 1 is excited, and the radiation is nearly omni-
directional. With higher modal numbers, the radiation is more directive,
but with a clear refraction effect (Figure 6.25b and c). In these simulations,
the resonant radiation is strong and the radiation from the forced bending
wave is not visible.
(a) (b)
(c) (d)
REFERENCES
The total mechanical energy in a vibration system is the sum of the kinetic
and the potential energies. In a plate with normal modes, the kinetic and
the potential energies are on average equal. An explanation of this is that
189
190 Sound Insulation in Buildings
at resonance the kinetic and the potential energies are per definition equal,
and the average response magnitude is mainly determined by the response
at the resonances. Since it is common to describe mechanical vibrations by
the velocity of the surface, i.e. the kinetic part of the energy, we calculate
the total energy E as twice the kinetic energy, as in:
E = mS v 2 (7.1)
where m is the mass per unit area of the plate, S is the area of the plate and
v 2 is the mean square of the velocity averaged over the area of the plate.
For a sound field in a room, the energy is commonly described by the
sound pressure, i.e. the potential part of the energy. The total energy E in
the room can be expressed as twice the potential energy, as follows:
p 2
E= V (7.2)
ρ c2
where V is the volume of the room, p is the mean square of the sound
2
pressure averaged over the volume of the room, ρ is the density of air and c
is the speed of sound in air.
The loss factor η of a resonant system is defined as the dissipated energy
in a time period of vibration divided by 2π and the total mechanical energy:
Pd T0 P
η= = d (7.3)
2π E ω E
where Pd is the dissipated power (energy per time unit), T0 is the time period
of the vibration, E is the current total energy in the system and ω = 2πf = 2π/
T0 is the angular frequency.
In a room, the energy losses are usually described by the equivalent
absorption area A or by the reverberation time T. If the volume is V and the
sound field in the room can be assumed to be approximately diffuse, the
energy density in the stationary sound field is E / V = 4Pin / cA, where Pin is
the sound power emitted in the room. Energy balance at stationary condi-
tions means that Pin = Pd and, thus, the loss factor of a room is:
Pd cA cA
η= ⋅ = (7.4)
ω 4Pd V 4ω V
The reverberation time is given by Sabine’s equation T = 55V /cA, and the
loss factor of a room can also be expressed as:
55 2.2
η= ≅ (7.5)
4ω T fT
Statistical energy analysis, SEA 191
Two resonant systems are said to be coupled if energy can be transferred from
the normal modes in one system to the normal modes in other system. This
requires two conditions to be fulfilled, a physical one and an acoustical one:
P12′
η12 = (7.6)
ω E1
where P12′ is the power (energy per time unit) transferred from system 1 to
system 2, E1 is the current total energy in system 1 and ω is the angular
frequency.
Example. The sound power radiated from resonant modes in a plate
(system 1) into a room (system 2) can be derived. The radiated sound power
from one side of the plate with velocity of vibration v 1, area S1 and radia-
tion efficiency σ 1,res is (Equation 6.24):
Insertion of this and the total energy (Equation 7.1) into the definition
yields:
ρc
η12 = σ 1,res (7.8)
ω m1
192 Sound Insulation in Buildings
This shows that the energy transfer to the room is strongest at low frequen-
cies and for low mass per unit area if the radiation efficiency is constant.
Two coupled resonant systems are considered (Figure 7.1). Each system
may have a source that supplies the system with mechanical or acous-
tic energy, P1i and P2i, respectively. The dissipated energy is P1d and P2d ,
respectively.
The net energy transfer per unit time from system 1 to system 2 can be
written as:
However, the consideration of the energy balance between the two systems
should be made on the basis of the normal modes that carry the energy.
It is the energy per mode in each system that is important for the energy
balance, as it is the modes that are the degree of freedom of the system. In
system 1, the average energy per mode Em1 is:
E1 E ∆f
Em1 = = 1⋅ (7.10)
∆N1 ∆f ∆N1
where Δf is the frequency band under consideration and ΔN1 is the number
of modes in system 1 within this frequency band. The ratio ΔN1/Δf is called
the modal density of system 1, i.e. the number of modes per Hz around the
centre frequency f.
The energy balance equation for the two systems can now be expressed
in terms of the average energy per mode:
Room Plate
P1, in P2, in
P΄12
P΄21
E1 E2
P1, d P2, d
Figure 7.1 Block diagram to show the energy transfer between two coupled resonant
systems.
Statistical energy analysis, SEA 193
The idea is that each normal mode in a resonant system can be considered a
degree of freedom, and it is a general physical principle that energy transfer
between coupled systems will always be in the direction from the system
with higher energy per degree of freedom to the system with lower energy
per degree of freedom. This means that in the case of energy balance,
Em1 = Em2 , the net power flow P12 is zero and thus Equation 7.11 yields,
∆N1
η21 = η12 (7.12)
∆N2
This is the very important reciprocity relation for coupled resonant sys-
tems. It says that it is possible to derive the coupling loss factor in one
direction if the coupling loss factor is known in the opposite direction and
if the modal densities of the two systems are also known. As an example,
it is well known how a plate radiates sound into a room, and with the
reciprocity relation, it will then be possible to estimate how much energy is
transferred into a plate from the sound field in a room.
The energy balance equation can be expressed by only one coupling loss
factor using Equation 7.12:
These assumptions are the reason for calling the method statistical. None
of the assumptions are strictly correct, but on average they may be the best
available estimate. Later, it will be shown that the results from the SEA
method may also be used without the statistical part in case of very low
modal density in one of the systems. The modal densities of some typical
systems are collected in Table 7.1.
194 Sound Insulation in Buildings
Table 7.1 Statistical modal density in different one-, two- and three-dimensional
systems
System Modal density, ΔNi/Δf
Room, 3D modes only 4π
Volume V Vf 2
c3
More accurate, see Equation 4.13
Narrow cavity, 2D modes only 2π
Area S = lx ly Sf
c2
Tube, both ends either closed or open, 1D modes only 2
Length lx lx
c
Thin plate, bending waves π
Area S, critical frequency fc Sfc
c2
More accurate, see Equation 6.9
Thick plate, shear waves 2π
Area S, speed of shear waves cs Sf
cs2
More accurate, see Equation 6.10
Beam, longitudinal waves 2
Length lx, speed of sound cm lx
cm
7.3 RESONANT TRANSMISSION THROUGH
A SINGLE WALL
P΄12 P΄23
P΄21 P΄32
E1 E2 E3
Figure 7.2 Block diagram to show the energy transfer between three coupled resonant
systems.
Statistical energy analysis, SEA 195
p12 S2 EV S
R = 10 lg 2 = 10 lg 1 3 2 (7.14)
p3 A3 E3V1A3
where S 2 is the area of the wall and A 3 is the absorption area of the receiv-
ing room.
The SEA method implies to set up the energy balance equations for each
of the systems, i.e. three in this case:
System 1 is excited by a stationary sound source, and the equation for that
system will only be used if it is relevant to analyse the results as a function
of the power of the source. In the present situation, this is not the case; it is
sufficient to find the ratio between the energy of systems 1 and 3, which is
independent of the power of the source. So, it will be sufficient to continue
with the equations from systems 2 and 3, i.e. the systems without a source.
The loss factors (Equation 7.3) and the coupling loss factors (Equation 7.6)
are inserted in Equations 7.16 and 7.17:
Here, the total loss factors have been introduced in systems 2 and 3. The
loss factor of the wall is combined from internal losses in the solid material
and boundary losses (both contained in η2) and radiation losses to either
side (these are the coupling loss factors from the wall to each of the rooms,
η21 and η23). The total loss factor is the one that can be measured experi-
mentally since it is not possible to avoid the radiation losses. The total loss
factor of the receiving room combines the absorption of sound from the
wall (η32) and the absorption from all other surfaces in the room (η3,d).
Again, the total loss factor is the one that can be measured experimentally.
196 Sound Insulation in Buildings
E1 η2,totη3,tot η32
= − (7.20)
E3 η12η23 η12
The last term can be neglected because it must be close to unity and the
total ratio is expected to be >>1. In the first term, it is an advantage to
change the coupling loss factor η12 with η21, which is done by applying the
reciprocity relation (Equation 7.12):
leading to
2
E1 4π ∆f cA3 ω m2
≅ 3 V1f 2 ⋅ ⋅ η2,tot ⋅ ⋅ (7.22)
E3 c ∆N2 4ωV3 ρcσ 2,res
2
E1V3S2 ω m2 c 2σ 2,res
2
∆N2
Rr = 10 lg = 10 lg − 10 lg ⋅ (7.23)
E3V1A3 2ρc 2η2,tot S2 f ∆f
In this result, the first term is recognised as the mass law at normal inci-
dence R 0. In this form, Equation 7.23 is a generally valid expression for
thin or thick walls, having either high or low modal density. The modal
density can be inserted according to the type of wall (Table 7.1) (bending
waves or shear waves). If a thin wall is assumed (only bending waves and
high modal density), the resonant transmission loss is approximately:
πσ 2 f
Rr = R0 − 10 lg 2,res c ,2 (7.24)
2η2,tot f
represents the excitation of the plate from the source room (reciprocity),
while the second one represents the radiation into the receiving room.
The forced part of the transmission cannot be handled with SEA. However,
it will be more thoroughly studied in Chapter 8. The forced transmission
can be added as an independent transmission path connecting source room
and receiving room (Figure 7.3). This is equivalent of adding a direct field
to a diffuse field in room acoustics.
The total transmission through the wall is found by adding the sound
powers of forced and resonant transmission:
Here, the velocity of forced vibrations is v 2,for and the radiation efficiency
of the forced vibrations is σ2, for (Equation 6.40). This equation is a good
basis for a sound transmission model; forced and resonant transmission is
combined, and for both parts, the transmission process is divided into exci-
tation (generating the vibrations in the panel) and radiation.
The transmission loss including both forced and resonant transmission
is then:
R = 10 lg
Pin
Ptot
= 10 lg
Pin
Pfor + Pres
( )
= −10 lg 10−0.1Rf + 10−0.1Rr (7.26)
P12,res P23,res
E1 E2 E3
Figure 7.3 Block diagram to show the combination of forced and resonant transmission
through a wall.
198 Sound Insulation in Buildings
The acoustic excitation of a plate by sound incident from one side can be
derived from the previous example with one room on either side of the plate
or it can be simplified to one room and one plate only. If the room is system
1, the plate is system 2, and if the room is excited by a stationary source, we
only need to consider the energy balance equation for system 2:
In terms of loss factor and coupling loss factors, this leads to:
Here, it should be noted that in this particular case, the loss factor η2
includes the radiation loss from the back of the plate, since η21 represents
the radiation loss from the exposed side of the plate back to the source
room. Compare Equation 7.28 with the three-system model in the previous
section (Equation 7.18). In any case, it is more practical to introduce the
total loss factor of the plate. Then, the energy ratio is:
If the plate is excited from one side as a wall that is part of the surfaces of a
room, the coupling loss factor is as in Equation 7.8 except that the number-
ing is different here. Together with the modal density for room 1, this leads
to the result:
2
v 2,res c2 σ 2,res ∆N2
≅ ⋅ ⋅ (7.30)
(ω m2 ) 2f η2,tot S2 ∆f
2
p12
This is a general result for the acoustic excitation of the resonant modes
in a plate from one side, as a function of the modal density of the plate.
Assuming a thin plate with bending waves, this simplifies to:
2
v 2,res π fc ,2 σ 2,res (7.31)
≅ 2 ⋅
p12 (ω m2 ) 2f η2,tot
Statistical energy analysis, SEA 199
2
v 2,res π fc ,2σ 2,res (7.33)
≅ 2 ⋅
p12 (ω m2 ) f η2,tot
Room (1)
V1
Slab (2) S2 S1
S3
V3
Room (3)
Figure 7.4 Case of a large slab separating a source room and a receiving room. Other
rooms are not shown. All walls are assumed to be lightweight and the vibra-
tion energy equally distributed in the slab.
200 Sound Insulation in Buildings
room (2) has the volume V3 and the floor area S 3. The sound is transmitted
through a large slab with area S 2 that is greater than both S1 and S 3. We
also assume that the walls are lightweight and, thus, the junction attenu-
ation can be neglected. So, we have three coupled resonant systems as in
Figure 7.3, but now the area of the slab is different from the excitation area
S1 and the radiation area S 3.
The steady-state energies in the three systems are:
p12
E1 = V1 (7.34)
ρc2
E2 = m2S2 v 22 (7.35)
p 32
E3 = V3 (7.36)
ρc 2
ρ cS3σ
η23 = (7.37)
ω m2S2
∆N1 4π
≅ 3 V1f 2 (7.38)
∆f c
∆N2 π
≅ 2 S2 fc (7.39)
∆f c
In the same way as in Equation 7.22, we find the ratio of energy in source
and receiver rooms:
2
E1 2η2,tot f V1A3S2 ω m2 (7.41)
≅ ⋅ ⋅
E3 πσ 2 fc V3S1S3 2ρ c
Statistical energy analysis, SEA 201
2
EV S ω m2 πσ 2 fc S2
Rr = 10 lg 1 3 1 = 10 lg − 10 lg + 10 lg S3
E3V1A3 2ρ c 2η2,tot f
πσ 2 fc S2
Rr = R0 − 10 lg + 10 lg S3 (7.42)
2η2,tot f
Comparing with Equation 7.24, we see that the extended area of the slab
gives rise to an increased transmission loss. If the area of the slab is 10 times
the area of the ceiling radiating into the receiving room, the transmission
loss may be up to 10 dB higher than in the normal case where S 2 = S 3.
REFERENCES
M.J. Crocker, A.J. Price (1969). Sound transmission using statistical energy analysis.
Journal of Sound and Vibration 9, 469–486.
R.H. Lyon, G. Maidanik (1962). Power flow between linearly coupled oscillators.
Journal of Acoustical Society of America 34, 623–639.
Chapter 8
203
204 Sound Insulation in Buildings
(a) (b)
Pres
vres
vfor
Pfor
Figure 8.1 Principle of sound transmission through a panel. Excitation and radiation are
different for forced transmission (a) and resonant transmission (b).
For each transmission part, the physical process can be divided into excita-
tion and radiation. The excitation gives rise to the vibrations of the panel,
and the velocity is controlled by the impedance (mass and stiffness) of the
panel. The radiation is described by the radiation efficiency, which is differ-
ent for forced and resonant radiation (see Chapter 6).
ωm 2
R0 = 10 lg 1 + (dB) (8.2)
2ρ c
m f
R0 ≈ 20 lg + 20 lg − 2 dB (8.3)
1kg/m 2 100 Hz
Airborne sound transmission through single constructions 205
This simple relation between transmission loss, mass per unit area and
frequency is called the mass law, and is the basis of all the following
considerations on airborne sound insulation. The mass law is characterised
by a 6 dB increase every time m or f is doubled.
8.1.3 Wall impedance
The resistance of a wall or a plate against a sound wave is called the wall
impedance Zw or the transmission impedance. Zw is defined as the ratio
between pressure difference ∆p across the plate and the velocity v of the
vibrations of the plate (Equation 5.23):
∆p
Zw = (8.4)
v
When ∆p and v are out of phase, the wall impedance is a complex number,
which is normally the case. The wall impedance is determined from the
bending wave equation (see Equation 1.24). Assuming a harmonic time
dependency (with angular frequency ω = 2π f) of the pressure exciting the
plate, the bending wave equation is:
2
∂2 ∂2
B 2 + 2 v − mω 2 v = jω∆p (8.5)
∂x ∂y
2
∂2 ∂2
B 2 + 2 vˆ e− jkr sin θ − mω 2 vˆ e− jkr sin θ = jω∆pˆ e− jkr sin θ (8.7)
∂x ∂y
Zw =
∆pˆ
vˆ
=
1
jω ( )
B (k sin θ ) − mω 2 = ZB + Zm
4
(8.8)
206 Sound Insulation in Buildings
Zm = jω m (8.9)
This would be sufficient for thin panels. However, in order to derive results
applicable also to thick panels, it is necessary to consider shear waves in
addition to bending waves. The shear wave impedance can be found from
a wave equation for shear waves in a way similar to that for bending waves
and this leads to:
2
1 ω c
ZS = Gh sin θ = −jω m ⋅ S ⋅ sin2 θ (8.11)
jω c c
where G is the shear modulus, h is the thickness of the panel and c s is the
speed of shear waves in the panel.
Like a simple mechanical system, the wall impedance contains three
elements representing mass, stiffness and losses. The wall impedance
can be described by an electrical analogy, as shown in Figure 8.2. The
mass is represented by an inductor and the losses by a resistance r ≈ ηωm
(Equation 2.28). The inductor is L = m, and the capacitors are:
c4 c2
C1 = , C2 = 2
ω B
4
ω Gh
Bending Shear
C1 C2
1 Z Z
(C1 + C2 ) + r = Zm + B S + r
−1
Zw = jωL +
jω ZB + ZS
(c /c)2 (c /c)4
= jω m 1 − S 2 B 4
⋅ sin4 θ + ηω m (8.12)
(cS /c) + (cB /c)
c 2 c 4
−1
= jω m 1 − sin θ ⋅ + + ηω m
4
cS cB
In case of a material with c S ≫ c like glass or steel, the wall impedance can
be approximated by the more simple formula 5.35.
The imaginary part of the wall impedance (Equation 8.12) relative to the
mass impedance is displayed as a function of the frequency in Figure 8.3a
for grazing incidence and in Figure 8.3b for the angle of incidence 45°.
The wall impedance is zero when coincidence occurs. We notice that
coincidence at grazing incidence is not exactly at the critical frequency,
but slightly above fc for c S /c = 2. The minimum of the impedance is shifted
to very high frequencies for c S /c→ 1. If c S /c < 1, which can be the case in
sandwich panels, there is no possibility of coincidence, simply because the
speed of the transverse waves is subsonic. At the angle of incidence 45°, the
coincidence dip is shifted to higher frequencies by one octave for c S /c ≫ 1.
The wall impedance (Equation 8.12) at grazing incidence (θ = 90°) can
be used to derive the effective speed of free transverse vibrations, as shown
by Kurtze and Watters (1959). A plane sound wave with variable speed c
is assumed to propagate along the surface of the panel and excite a forced
transverse motion. The ratio of the pressure to the transverse velocity is
the wall impedance, which is a function of the speed of the exciting wave.
The speed of free transverse waves in the panel cB,eff is found by the mini-
mum energy condition, i.e. the impedance must be minimum that means
Im{Zw} = 0. This leads to an equation in c of the fourth degree:
The speed cB,eff equals the speed c of the exciting wave in this equation, and
according to Rindel (1994), the solution is:
4 −1/3
cB2 1 1 c 1 1 (8.14)
cB,eff = − + 1+ 4 S ≅ 3 + 3
cS 2 2 cB cB cS
208 Sound Insulation in Buildings
(a) 2
θ = 90°
1.5
|Zw / Zm|2
0.5
0
0.1 0.2 0.5 1 2 5 10
f/fc
10 2 1 0.5
(b) 2
θ = 45°
1.5
|Zw / Zm|2
0.5
0
0.1 0.2 0.5 1 2 5 10
f /fc
10 2 1 0.5
Figure 8.3 Wall impedance relative to the mass impedance as function of frequency for
different values of cS /c. (a) At the angle of incidence 90° and (b) at the angle
of incidence 45°. Solid line: cS /c = 10, dashed line: cS /c = 2, dotted line: cS /c = 1,
dashed–dotted line: cS /c = 0.5.
where the approximation is accurate within ±1 % (see also the dispersion
curve in Figure 1.4).
At low frequencies, the impedance of the plate is dominated by the reso-
nance at the first normal mode. If the size of the panel is not extremely
small, the cross-over frequency from bending waves to shear waves is
fs ≫ f11 and bending waves can be assumed at the first normal mode. Thus,
Airborne sound transmission through single constructions 209
the natural frequency of the first normal mode f 11 and the critical frequency
fc are closely related, as seen from Equation 6.7 with (m, n) = (1, 1):
c2 1 2 1 2
f11 = + (8.15)
4 fc lx ly
where lx and ly are the plate dimensions. Therefore, a high fc means a low f11
and vice versa. Some examples are given in Table 8.1. It is noted that the thick
plates (concrete and brick) have a much lower modal density than thin plates
(porous concrete, glass or gypsum), even if the latter have smaller dimensions.
At frequencies below and around the natural frequency of the first struc-
tural mode, f11, the stiffness comes not only from bending the plate but
also from the connections to the surroundings at the boundaries. As an
approximation, we assume the system to behave as a simple mechanical
system with the natural frequency f 11.
k
Zw = jω m − d + r
ω
(8.16)
2
f
= jω m 1 − 11 + ηω m
f
Combining with the above result (Equation 8.12) and using Equation 5.32 to
replace cB with fc, we get the wall impedance at the angle of sound incidence θ:
2 2 2
−1
f c f
Zw = jω m ⋅ 1 − ⋅ 1 − sin θ ⋅ + + ηω m (8.17)
11 4 c
f c S f
In analogy with Equation 6.38, we can take into account the finite size of
the panel by replacing the term sin 2 θ with the radiation efficiency:
−2
sin2 θ = 1 − cos2 θ → 1 − σ for,θ (8.18)
Table 8.1 C
alculated critical frequency, first natural frequency and number of
modes for examples of thin and thick plates
ΔN in One-Third
Octave
Plate
Dimensions (m) fc (Hz) f11 (Hz) 100 Hz 500 Hz
4 mm glass 1.2 × 1.3 3250 12 3 15
13 mm gypsum board 2.5 × 1.2 2615 10 5 24
75 mm porous concrete 2.5 × 4.0 707 9 4 22
180 mm concrete 2.5 × 4.0 106 62 0.7 3
250 mm brick 2.5 × 4.0 72 91 0.4 2
210 Sound Insulation in Buildings
where σfor,θ is the radiation efficiency of forced waves (Equation 6.39). This
can be used for relatively high frequencies where σfor,θ > 1, but for low
frequencies (small ka) coincidence is not possible, and sin 2 θ → 0.
For a diffuse sound field with random incidence, a good approximation
valid for ka > 3.84 ≈ 4 is obtained by the corresponding value for grazing
incidence, θ = 90° in Equation 6.39:
−2 3.84
sin2 θ → 1 − σ for,90 ≅ 1− (8.19)
ka
For ka < ca. 4, the plate is acoustically small and coincidence is not
possible. The reason is that the dimensions of the plate are too small
compared to the wavelength of the incident sound. So, instead of sound
propagation along the plate, diffraction effects mean that the transmit-
ted sound is refracted (see Figure 6.25; also Figure 6.10 for a graphical
presentation of the radiation efficiency as function of ka and the angle
of incidence).
U λBα Ucα
= 2 ( f ⋅ fc )
−1/ 2
ηborder = (8.20)
π 2S π S
2ρ c
ηrad = ⋅ σ res (8.21)
ωm
Airborne sound transmission through single constructions 211
1.000
cross
0.010
0.001
0.1 1 10
Ratio of plate thickness h2/h1
Figure 8.4 Absorption at the boundary where all plates are connected rigidly and made
of the same material but with different thickness. The absorption relates
to the plate with thickness h1, and h2 is the thickness of the crossing plate.
(Adapted from Craik, R.J.M., Appl. Acoust., 14, 347–359, 1981.)
Below the critical frequency, the radiation loss factor is small, but it can be
quite significant around the critical frequency. The total loss factor is:
Ucα 2ρ c
2 (
f ⋅ fc )
−1/ 2
ηtot = ηint + ηborder + ηrad = ηint + + ⋅ σ res (8.22)
π S ωm
When measuring the loss factor of a plate, it is the total loss factor that is
measured, as it is not possible to eliminate the radiation losses.
Examples of measured loss factors are shown in Figure 8.5. The wall
elements were 2.54 m high and 1.80 m wide and were connected in a
T-junction to a 180 mm clinker concrete slab. The loss factor generally
increases towards the lower frequencies. One of the walls was measured
with an elastic connection to the slab, and in this case the loss factor was
clearly higher than with the rigid connection, which may be due to viscous
losses in the elastic connection.
8.1.5 Forced transmission
Let the sound pressure of the incident, reflected and transmitted sound be
denoted as pi, pr and pt, respectively. Then, ∆p = pi + pr − pt ≅ 2pi and the
squared sound pressure in the source room in some distance from the wall
is p12 ≅ p i2 + p r2 ≅ 2p i2 because we assume nearly total reflection from the
panel, i.e. pi ≈ pr. The relation between velocity of the forced vibrations of
the plate vfor and the sound pressure p1 in the source room yields:
2
2 p12
v for ≅ 2 (8.23)
Zw
212 Sound Insulation in Buildings
0.1
Loss factor, η
0.01
0.001
50 100 200 500 1000 2000 5000
Frequency, Hz
Rigid connection Elastic connection
Figure 8.5 Measured loss factor as a function of frequency for two walls, which were
the lower part of T-junction to a slab of 180 mm clinker concrete (1700 kg/
m3). Full line: 150 mm porous concrete wall (660 kg/m3) with rigid connection
and dotted line: 100 mm porous concrete wall (660 kg/m3) with elastic
connection. (Adapted from Rindel, J.H., Appl. Acoust., 41, 97–111, 1994.)
Pfor = ρ cS v for
2
σ for (8.24)
where σfor is the radiation efficiency for forced bending waves (Equation
6.40). From Equation 5.5 and the above Equations 8.23 and 8.24, we get
the transmission loss for forced transmission:
2
p12 S 2
Zw ⋅ v for
Rf = 10 lg = 10 lg
4 ρ cPfor 8ρ c ⋅ Pfor
(8.25)
Z
= 20 lg w − 10 lg ( 2σ for ) (dB)
2ρ c
Airborne sound transmission through single constructions 213
Considering a diffuse sound field in the source room, the wall impedance
(Equation 8.17) is inserted with θ = 90° because this angle of incidence cor-
responds to the coincidence dip at random incidence. The forced transmis-
sion for random incidence yields:
−1 2 −1 2
f11 2 f 2 c 2 cf2c 2f 2
2 2
Rf = R0f += 10
R0lg+ 10
1lg− 1 − 11
⋅ 1 − ⋅ 1 − + + c 2
+ η + η
2
f f cS cS f f tot tot
−10lg−(10lg
2σ for()2(dB)
σ for ) (dB) (8.26)
The loss factor ηtot determines how sharp the dip is, either at coincidence or
at the first structural mode.
For thin plates (f11 ≪ fc) and at frequencies well below the critical fre-
quency, the forced transmission (Equation 8.26) can be approximated by
the transmission loss:
1
Rf ≅ R0 − 10lg (2σ for ) ≅ R0 − 5 (dB) f11 < f < fc (8.27)
2
Here, the last approximation is valid within ±2 dB for surfaces above 10 m 2
and frequencies above 100 Hz. However, for small areas and low frequen-
cies, the area has a significant influence on the transmission loss (Figure 8.6).
The reason for this area effect of sound transmission is the same as applied
to a loudspeaker; the diameter of the loudspeaker membrane determines
how well low frequencies can be radiated. The low-frequency limit is deter-
mined by the ratio between wavelength and diameter.
At frequencies above the critical frequency, the bending stiffness starts
to dominate the wall impedance and the transmission loss increases rapidly
with frequency:
50
40
30 R0
S = 1 m2
R, dB
S = 4 m2
20 S = 12 m2
S = 50 m2
S ∞
10
0
50 100 200 500 1000
Frequency, Hz
Figure 8.6 Influence of area on the transmission loss of a square panel with m = 25 kg/m2 .
(Adapted from Sato, H., J. Acoust. Soc. Jpn. 29, 509–516, 1973.)
8.1.6 Resonant transmission
As explained earlier, a plate that is excited by airborne sound will exhibit
forced vibrations and resonant vibrations, if the plate has any normal
modes in the frequency band of excitation. The resonant transmission
cannot be described in the same way as the forced transmission by
the wall impedance method, but instead the SEA can be applied (see
Chapter 7).
The transmission loss for resonant transmission alone Rr has been found
in Equation 7.23:
c 2σ res
2
∆N
Rr = R0 − 10lg ⋅ (dB) (8.30)
2ηtot Sf ∆f
the critical frequency, but it is the dominating transmission above the criti-
cal frequency where σres ≈ 1.
In the frequency range where bending waves can be assumed, the modal
density can be approximated by Equation 6.9 and the resonant transmis-
sion loss is:
πf σ 2 1
Rr ≅ R0 − 10 lg c res (dB) 4 f11 < f < f s (8.31)
2ηtot f 2
π c 2σ res
2
1
R ≅ R0 − 10 lg 2
(dB) f > f s (8.32)
ηtotcS 2
A slight change in the slope with frequency can be noted at fs /2. While R0
has a slope of 6 dB per octave, the resonant transmission increases with
9 dB per octave below fs /2 and continues with 6 dB per octave above fs /2.
As pointed out in Chapter 6, the statistical approximations for modal
density in a plate should not be applied at frequencies below 4 f 11.
( )
R = −10 lg 10−0.1Rf + 10−0.1Rr (8.33)
Using this with Equation 8.26 for forced transmission and Equation 8.30
for resonant transmission, this makes a general calculation model for
the random incidence transmission loss of homogeneous and isotropic
constructions.
In addition to the area of plate, angle of incidence, speed of sound and
frequency, the main parameters are
∆N
: modal density in the plate (Equation 6.9) bending or (Equation 6.10)
∆f shear
216 Sound Insulation in Buildings
We shall discuss this result for thin and thick plates, which behaves quite
differently from an acoustical point of view. An acoustically thin plate has
f11 ≪ fc , whereas an acoustically thick plate can be defined as one having
fc ≤ f 22 = 4 f11.
The critical frequency plays an important role for the transmission loss
of thin plates, and there is usually a local dip in the transmission loss curve
at or slightly above the critical frequency. Below the critical frequency,
the radiation from resonant vibration is weak and the forced transmis-
sion will normally be the dominating part of the transmission. However,
above the critical frequency, the resonant vibrations radiate efficiently and
dominate over the forced transmission. This behaviour is shown graphi-
cally in Figure 8.7. Between f 22 and ½ fc , the sound transmission is con-
trolled by the mass per unit area following the mass law, i.e. with a slope
of 6 dB per octave. It is noted that the increase of R by 6 dB per doubling
of the mass (the mass law) is only valid within a limited frequency range.
Increasing the thickness of the plate causes a lower critical frequency and a
higher modal resonance frequency, narrowing the mass law region.
Figures 8.8 and 8.9 show two examples of application of the calculation
model for an acoustically thin panel, namely, a 10 mm glass pane, but with
different dimensions. In the calculations, the internal loss factor is 0.01 and
Figure 8.7 Principle of sound transmission loss for an acoustically thin single construc-
tion as a function of frequency. The critical frequency is high, fc > f 22 . Increasing
the thickness of the plate causes a shift from the solid curve to the dotted
curve and the key frequencies move up and down as indicated by the arrows,
narrowing the mass law region.
Airborne sound transmission through single constructions 217
80
70
fc
60
Transmission loss, R, dB
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency, Hz
Figure 8.8 Measured and calculated transmission loss of a 10 mm glass pane (25 kg/m2).
Dimensions 1.6 m × 1.3 m. The solid line is measured, the dashed line is calcu-
lation with the complete model (Equation 8.33) and the dashed-dotted line is
calculation with a simplified model (Equation 8.27). (Adapted from Cops, A.,
Myncke, H., The Sound Insulation of Glass and Windows – Theory and Practice. IV
Jornadas G.A.L.A., Cordoba, 1971.)
absorption coefficient at the edges is 0.2. The coincidence dip at the critical
frequency (ca. 1300 Hz) is clearly seen in the large glass pane (Figure 8.8).
Here, the first structural modes are below f 22 = 90 Hz and thus of no impor-
tance. Below the critical frequency, the transmission loss is controlled by
forced transmission, and the agreement with the simplified equation 8.27
is good.
However, several changes can be observed in the case of a small pane
(Figure 8.9). Here, f11 = 90 Hz and f 22 = 360 Hz and the low frequencies are
strongly influenced by the structural modes. The coincidence dip is shallow
and shifted to a somewhat higher frequency and the loss factor is higher
due to increased influence of boundary losses. The simplified model does
not work in this case, but the agreement with the complete model (Equation
8.33) is very good. The area effect shown in Figure 8.6 is not so clear in
reality; it is to some extent overruled by dips at the lower structural modes.
An example of a wall with a somewhat lower critical frequency is a
100 mm thick porous concrete wall. The measured transmission loss is
218 Sound Insulation in Buildings
80
70
f11 fc fb
60
50
Transmission loss, R, dB
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency, Hz
Figure 8.9 Measured and calculated transmission loss of a 10 mm glass pane (25 kg/m2).
Dimensions 0.8 m × 0.65 m. The solid line is measured, the dashed line is calcu-
lation with the complete model (Equation 8.33) and the dashed-dotted line is
calculation with a simplified model (Equation 8.27). (Adapted from Cops, A.,
Myncke, H., The Sound Insulation of Glass and Windows – Theory and Practice. IV
Jornadas G.A.L.A., Cordoba, 1971.)
compared with the calculated one, using Equation 8.33 in Figure 8.10. In
the calculations, the internal loss factor is 0.001 and absorption coeffi-
cient at the edges is 0.1. The critical frequency is calculated to 374 Hz, but
the coincidence dip is located at a somewhat higher frequency, f b = 508 Hz
(Equation 6.55). The first structural modes are below f 22 = 67 Hz and thus
of no importance.
80
70
60
Transmission loss, R, dB
50
40
30
20
10
fc fb 1/2 fs
0
50 100 200 500 1000 2000 5000
Frequency, Hz
Aerated concrete Calculated R0, dB
Figure 8.10 Measured and calculated transmission loss of a 100 mm thick wall of porous
concrete. Dimensions are 3.7 m × 2.69 m. Mass per unit area is 130 kg/m2. (Data
from Lydteknisk Institut, Bygningsakustiske laboratoriemålinger, Building Acoustic
Laboratory Measurements, in Danish, Lyngby, Denmark, LL 871/68.)
Figure 8.11 Principle of sound transmission loss for an acoustically thick single construc-
tion as a function of frequency. The critical frequency is low, fc ≤ f 22 .
220 Sound Insulation in Buildings
8.31 and 8.32 above the frequency f 22 = 4 f11. The slope of the transmis-
sion loss curve changes from 9 dB per octave below ½ fs to 6 dB per octave
above ½ fs (Figure 8.11). At very high frequencies, the longitudinal waves
cause an upper limit of the transmission loss, although this region if very
seldom seen in practice. The cross-over frequency is f h as previously defined
(Equation 5.21).
At the lower frequencies, the sound transmission of thick plates is domi-
nated by single structural modes. In the example of a 180 mm concrete
wall, the first structural mode is at 106 Hz, and already below 400 Hz the
modal density is so low that the equations derived from SEA cannot be used
directly. A statistical approach is not valid when the modal density is low
i.e. below f 22 .
However, we may still use the general Equation 8.33 but with great cau-
tion. With low modal density, the SEA can be replaced by a modal energy
analysis. This implies that the first structural modes are calculated one
by one, and then the actual mode count ΔN in each frequency band is
observed. If one-third octave bands are used, Δf = 0.23 f and the actual
number of modes ΔN are inserted in Equation 8.30. In some frequency
bands ΔN = 0, which means that no resonant transmission is present in that
particular frequency band; only forced transmission remains. Other bands
may contain just one mode, ΔN = 1, e.g. the one-third octave band where
f11 is located. Since the first structural modes are much spread (there are
normally only two modes between f 11 and f 22 that are two octaves apart),
the transmission loss may exhibit great variations and pronounced dips
between f11 and f 22 .
The example in Figure 8.12 is a 185 mm hollow concrete slab with criti-
cal frequency around 88 Hz and the first natural frequencies are within
the important frequency range (f11 ≈ 67 Hz and f 22 ≈ 268 Hz). In the calcu-
lations, the internal loss factor is 0.01 and absorption coefficient at the
edges is 0.05. Above 250 Hz, the statistical modal density can be used and,
thus, the transmission loss can be estimated from Equations 8.31 and 8.32.
However, below 250 Hz there are very few resonant modes, so the one-
third octave bands 63 Hz, 160 Hz and 200 Hz each contain just one mode
(ΔN = 1) while the one-third octave bands 80 Hz, 100 Hz and 125 Hz have
none (ΔN = 0). Insertion in the general Equation 8.33 and using the sta-
tistical modal density above 200 Hz leads to the result shown in Figure
8.12. The dip at 160 Hz might look like a coincidence dip, but it is not. It
is actually the opposite, namely, the consequence of relatively high sound
insulation in the lower frequency range from 80 Hz to 125 Hz. In this fre-
quency range, the slab has no modes and, thus, the resonant transmission
is missing and the sound insulation at these frequencies is relatively high. At
frequencies above 160 Hz, the slope is close to 6 dB per octave and the tran-
sition to shear waves is around 840 Hz (calculated fs = 1680 Hz). The critical
frequency is also marked in the figure, but it is typical for an acoustically
thick construction that there is no dip at the critical frequency.
Airborne sound transmission through single constructions 221
80
70
60
50
Transmission loss, R, dB
40
30
20
0
50 100 200 500 1000 2000 5000
Frequency, Hz
Hollow concrete slab Calculated R0, dB
Figure 8.12 Measured and calculated transmission loss for a 185 mm hollow concrete
slab (density 1700 kg/m3). Dimensions are 3.37 m × 2.99 m. Mass per unit
area is 315 kg/m2 . (Data from Lydteknisk Institut, Bygningsakustiske laborato-
riemålinger, Building Acoustic Laboratory Measurements, in Danish, Lyngby,
Denmark, LL 709/74.)
At very low frequencies, e.g. under 100 Hz, the sound insulation is a com-
plicated interaction between the structure and the connected rooms.
Particularly, the dimensions of the receiving room have been shown to be
important (Gibbs and Maluski 2004). A pronounced dip in the transmis-
sion loss curve can be expected in the one-third octave band that contains
222 Sound Insulation in Buildings
80
70
60
50
Transmission loss, R, dB
40
30
20
f11 fc f21 f12 1/2fs
10
0
50 100 200 500 1000 2000 5000
Frequency, Hz
240 mm clinker concrete Calculated R0, dB
Figure 8.13 Measured and calculated transmission loss for a 240 mm clinker concrete
wall (density 1800 kg/m3). Dimensions are 3.70 m × 2.69 m. Mass per unit
area is 430 kg/m2 . (Data from Lydteknisk Institut, Bygningsakustiske laborato-
riemålinger, Building Acoustic Laboratory Measurements, in Danish, Lyngby,
Denmark, LL 1156/77.)
the frequency of the first axial room mode perpendicular to the partition
wall (or floor). If the room dimension perpendicular to the surface is called
L x, the dip will be at the frequency c/(2L x). Gibbs and Maluski (2004) have
found this dip to be on average
Heavyweight constructions are normally acoustically thick plates (fc < f 22)
and at frequencies below the critical frequency, the transmission loss is con-
trolled by forced transmission. Thus, the transmission loss below f 11 can be
approximated by the simple Equation 8.29. This means that theoretically
the transmission loss below f 11 increases towards lower frequencies with
12 dB per octave. The plate may behave like a stiff membrane, and the stiff-
ness comes from the fixations at the edges of the plate.
Airborne sound transmission through single constructions 223
80
70
60
Transmission loss, R, dB
50
40
30
20
10
f11
0
20 50 100 200 500 1000
Frequency, Hz
200 mm masonry walls Calculated Min. Max.
The results from the survey conducted by Gibbs and Maluski (2004)
included ~100 room pairs and showed a clear difference in the low-fre-
quency level difference for lightweight and heavyweight constructions. In
the frequency range from 20 Hz to 50 Hz, the lightweight walls had average
sound insulation around 20 dB to 25 dB decreasing towards lower frequen-
cies, whereas the heavyweight walls had average sound insulation around
40 dB to 60 dB increasing towards lower frequencies (Figures 8.14 and 8.15).
The explanation behind this finding must be that the frequency of the
first structural mode in lightweight walls is very low (f11 < 20 Hz), whereas
for heavyweight walls, this frequency is relatively high (f11 > 50 Hz). It is
quite remarkable how different the heavyweight walls and the lightweight
walls behave at the low frequencies. The measured data are averaged over
many different room dimensions and, thus, the above-mentioned dip due
to room depth has been averaged out. For the heavyweight walls, the struc-
tural modes have a strong influence and cause great variations in the fre-
quency range from 50 Hz to 125 Hz, but below 50 Hz the spread of the
measured data is surprisingly small.
224 Sound Insulation in Buildings
80
70
60
Transmission loss, R, dB
50
40
30
20
10
0
20 50 100 200 500 1000
Frequency, Hz
Lightweight walls Calculated Min. Max.
Some structures like drywalls and windows are often subdivided into
smaller areas, and this may influence the sound insulation in various
ways:
This means that the resonant transmission is changed, and most signifi-
cantly above the critical frequency, where the transmission loss is controlled
Airborne sound transmission through single constructions 225
80
70
60
Transmission loss, R, dB
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency, Hz
p12 S S
RE = 10 lg 2 ⋅ = L1 − L2 + 10 lg (dB) (8.34)
p2 A A
Figure 8.17 shows the parameters in Equation 8.34. The angle of incidence
does not go into the definition of external transmission loss.
Note that <p12> (and thus L1) is not really a spatial average since this is
the sound pressure in an open space. Instead, L1 is the sound pressure level
close to the wall including the energy reflection but excluding the interfer-
ence fluctuations.
It is also noted that RE is in reality the same as the apparent sound reduc-
tion index R’S as defined in ISO 16283-3 when the source is traffic noise [road,
rail or air traffic (see Equation 12.10)]. The only difference being that in the
ISO standard L1 is measured as the sound pressure level at the façade −3 dB.
The sound pressure level at a reflecting surface is known to be ~3 dB higher
than the average sound pressure level a few meters in front of the surface
(see Figures 3.7 and 3.8).
An advantage of the definition (Equation 8.34) is that information about
the direction of the incident sound is not necessary. It may have any angle of
incidence, or it may be a more complicated sound field, like sound from line
S L2
L1
θ
Figure 8.17 External transmission loss of a window with area S defined from the sound
pressure levels L1 and L 2 and the absorption area A in the receiving room.
The angle of incidence is θ.
Airborne sound transmission through single constructions 227
(e.g. a road). In the special case of a diffuse sound field on the source side,
the external transmission loss equals the normal transmission loss, R E = R.
2Pa
p12 = ρ c (8.35)
4π r 2
Thus, the external transmission loss can be expressed in terms of the sound
power of the outside point source and the sound pressure in the receiving
room:
P ρc S
a
RE = 10 lg ⋅ (dB) (8.36)
2π r 2 p22 A
The reciprocity principle states that in linear sound fields, the sound pres-
sure in one point due to an omnidirectional point source in another point
remains the same if the source and receiver are switched. Figure 8.18 shows
the reciprocity principle applied to the case of sound transmission through
a window. The room with absorption area A can be either the source or
the receiving room. The relation between the sound power of the source Pa
and the sound pressure p 2 remains the same when the source and receiver
switch positions.
(a)
A
S p2
Pa r
(b)
A
S Pa
p2 r
4Pa
p12 = ρ c (8.37)
A
P ρc S p12 S
a
RE = 10 lg ⋅ = 10 lg 2 ⋅ (dB) (8.38)
2π r p2 A
2 2 p2 8π r 2
This gives the following formula for calculation of the sound pressure level
L 2 in the outside receiver position when the indoor sound pressure level is
L1:
S
L2 = L1 − RE + 10 lg (dB) (8.39)
8π r 2
2
22 −1 2 2 −1 2
f11 2 f11 c
2
cfc fc 2
22
θ −⋅ sin
tot + ηtot
c θ ⋅ +c f + f + η
4 4 2
RE, f =R
= RR0E+, f 10 lg0 + 1 f 1 − ⋅f 1 −sin⋅ 1
10−lg
S S
σ for,θ )( 2(dB)
−10lg (−210lg σ for,θ ) (dB) (8.40)
The resonant transmission was derived by SEA in Equation 7.22. The result
contains twice the radiation efficiency for the resonant modes, one repre-
senting the excitation of the vibrations in plate and one representing the
sound radiation to the receiving room. In the case of the external transmis-
sion loss, these two radiation efficiencies must be treated separately because
the sound field is very different on either side. With transmission from the
outside to the inside, the first radiation efficiency is a function of the angle
Airborne sound transmission through single constructions 229
of incidence σres,θ (Equation 6.57) and the other one represents the radiation
into a room σres,d (Equation 6.51). Thus, the resonant external transmis-
sion loss at the angle of incidence θ can be calculated from the following
formula:
c 2σ res,θ σ res,d ∆N
RE,r = R0 − 10lg ⋅ (dB) (8.41)
2ηtot Sf ∆f
8.6 ORTHOTROPIC PLATES
Many building materials are orthotropic, i.e. they have different character-
istics in different directions as opposed to isotropic plates. In many wooden
plate materials, the bending stiffness is larger in the direction parallel with
the veins than perpendicular to the veins. Another example of an orthotropic
plate is a corrugated metal plate where a trapeze or waving profile involves a
very large difference on the bending stiffness in the two orthogonal directions.
As shown by Heckl (1960), it is possible to insert a geometric average
value Bxy of the two bending stiffnesses Bx and By as a good approximation
in many of the results already known from the isotropic plates:
Bxy = Bx By (8.42)
c2 m K
fcx = = cx (8.43)
2π Bx h
230 Sound Insulation in Buildings
(a)
60
θ = 0°
50
External transmission loss, RE, dB
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency, Hz
Measured Calculated
(b)
80
θ = 45°
70
60
External transmission loss, RE, dB
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency, Hz
Measured Calculated
Figure 8.19 Measured and calculated external transmission loss of a scale model window
made from 12 mm thick glass. (a) Angle of incidence 0°, (b) angle of incidence 45°
(Continued)
Airborne sound transmission through single constructions 231
(c)
80
θ = 75°
70
External transmission loss, RE, dB
60
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency, Hz
Measured Calculated
(d)
80
θ = 90°
70
External transmission loss, RE, dB
60
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency, Hz
Measured Calculated
Figure 8.19 (c) angle of incidence 75° and (d) angle of incidence 90°. Dotted line:
measured; full line: calculated.
232 Sound Insulation in Buildings
c2 m Kcy
fcy = = (8.44)
2π By h
where fcx is the critical frequency in the direction with the largest bending
stiffness, i.e. fcx < fcy, h is the plate thickness and Kcx and Kcy are material
constants. As a result of Equation 8.42, the critical frequency can in many
relations be replaced by:
However, this should be done with caution. For the natural frequencies of
the structural modes, Equation 6.7 changes to:
c 2 m2 n2
fmn = + (8.46)
4 fcxlx2 fcyly2
∆N π ⋅ S
≅ 2 fcx fcy (8.47)
∆f c
The two different critical frequencies involve the possibility of two dips in
the sound insulation and a relatively low sound insulation in the frequency
range between the dips. According to Heckl (1960), the radiation efficiency
in the frequency range fcx < f < fcy is determined by
2
1 fcx 4f
σ res ≅ ln f (8.48)
π2 fcy cx
Below the lower critical frequency (f < fcx), the transmission loss can be cal-
culated from the forced transmission (Equation 8.26) modified by introduc-
ing two critical frequencies instead of only one and simplified by neglecting
shear waves and the frequencies below f 11:
f
2
f
2
R ≅ Rf = R0 + 10 lg 1 − 1 − + ηtot 2
− 10lg ( 2σ for ) (8.49)
fcx fcy
Above the lower critical frequency fcx, the resonant transmission will nor-
mally be dominant, and the transmission loss is determined by inserting
Equations 8.47 and 8.48 in Equation 8.30:
1 f fcx 4f
4
R ≅ Rr = R0 − 10 lg 3
2π η
⋅ cx
f fcy f ( fcx < f < fcy ) (8.50)
ln
tot cx
Airborne sound transmission through single constructions 233
In some cases, the forced and the resonant transmission are both important
in this frequency range, and Equation 8.33 is used to find the total trans-
mission loss.
In the upper frequency range above the higher critical frequency, Equation
8.30 can be applied with the modal density (Equation 8.47) and simplified
with the approximation σres ≈ 1:
π fcx fcy
R ≅ Rr = R0 − 10 lg
2ηtot f
( f > fcy ) (8.51)
For some orthotropic plates, the interval between fcx and fcy is so small that
only one coincidence dip appears but the dip may cover a larger frequency
range than with isotropic plates. Other orthotropic plates may have a very
large interval between the critical frequencies; some corrugated plates can
have critical frequencies from below 200 Hz to above 20 kHz.
Figure 8.20 shows an example of measured and calculated trans-
mission loss of a corrugated steel plate. With reference to the sketch in
Figure 8.20, the dimensions of the corrugation are a = 120 mm, b = 120 mm,
c = 45 mm, d = 30 mm and h = 0.7 mm. The plate has a mass of 6.3 kg/m 2
and assumed internal loss factor of 0.001. The critical frequencies used for
80
70 c
h
60 a d b d
Transmission loss, R, dB
50 fcx
40
30
20
10
0
100 200 500 1000 2000 5000 10000
Frequency, Hz
Corrugated steel Calculated R0, dB
the calculations are 750 Hz and 18,500 Hz, respectively. Below the lower
critical frequency, the transmission is forced (Equation 8.48) and above that
frequency, the transmission is resonant (Equation 8.49). It is typical for a
corrugated plate that the transmission loss is far below the mass law.
8.7 SANDWICH PLATES
m = mc + 2mf = ρ c hc + 2ρ f hf (8.52)
(b) cB
(c) cS
(a) hf
(d) cBf
hc h
cBf
(e)
hf
Figure 8.21 (a) Section of sandwich panel, (b) bending wave, (c) shear wave, (d) bending
waves in the face plates and (e) dilatation resonance in elastic core.
Airborne sound transmission through single constructions 235
Here, index c refers to the core material and index f refers to the face plates.
According to Ver and Holmer (1971, p. 313), the bending stiffness of the
sandwich plate is:
1
Ef hf ( hf + hc ) (8.53)
2
B=
2
The bending wave is shown in Figure 8.21b. This has the speed of propaga-
tion cB determined by Equation 1.25:
B
cB = 2π f 4 (8.54)
m
One peculiarity of sandwich plates is that the bending stiffness is very high
even if it is a lightweight construction and, thus, the speed of the bending
wave is extremely high. Shear wave motion plays an important role in sand-
wich plates. In the shear waves, there is no compression or dilatation of the
face plates, and the movements in the core material are perpendicular to
the direction of propagation (Figure 8.21c). The propagation speed of the
shear wave is:
Gh
cS = (8.55)
m
where G is the shear modulus, h is the total thickness of the plate and m is
the total mass per unit area. The core material is enclosed such that it can-
not extend to the sides. The shear modulus is determined by the modulus of
elasticity Ec of the core material and the Poisson’s ratio μ:
Ec
G= (8.56)
2 (1 + µ )
At high frequencies (short wavelengths), the shear wave is influenced by the
bending stiffness of the face plates. So, if we divide the sandwich plate into
two parts along the dotted line in Figure 8.21d, the shear wave actually
becomes a bending wave in each of the face plates. The bending stiffness of
the face plate is (Equation 1.22):
Ef hf3
Bf = (8.57)
12(1 − µ 2 )
The propagation speed cBf is determined by Equation 1.25, but with half
the mass of the core material added to the mass of the face plate:
Bf 2Bf
cBf = 2π f 4 = 2π f 4 (8.58)
mf + m c /2 m
236 Sound Insulation in Buildings
The described waves (Figure 8.21b through d) will not appear separately
as pure bending or pure shear waves, but there is a gradual transition as a
function of the frequency. An electrical model of the wall impedance of a
sandwich panel is seen in Figure 8.22. The bending and shear stiffnesses
are in parallel as for a thick panel, but in addition the bending stiffness of
the face plates adds to the shear stiffness.
The inductor is L = m, and the capacitors are:
c4 c2 c4
C1 = , C2 = , C3 =
ω 4B ω 2Gh ω 4 Bf
The wall impedance is then calculated from:
−1
1 1
Zw = jωL + C1 + +r
jω 1 / C2 + 2 / C3
(8.59)
c4 4
cBf + c 2cS2
= jω m 1 − B4 ⋅ 4 4 + ηω m
2 2
c cB + cBf + c cS
cS2 6
c + c 4 − cS2c 2 − cBf
4
= 0 (8.60)
cB4
Bending Shear
C3
C1 C2
C3
cS = cB ⇒ cB,eff = c s
1
2
( )
5 − 1 ≅ cS ⋅ 0.7862 ≅ cS ⋅ 2−1/3
cS = cBf ⇒ cB,eff = cS
1
2
( )
5 + 1 ≅ cS ⋅ 1.2720 ≅ cS ⋅ 21/3
−1/3
1 1
cB,eff ≅ 3 + 3 3 (8.61)
c
B c S + cBf
CB
1000
Speed of propagation, m/s
1 fc
c = 344 m/s
Cs
CBf
100
log (frequency)
Figure 8.23 Dispersion curve for the effective bending wave speed in a sandwich
plate. Here, the shear wave speed is assumed to be less than the speed of
sound in the air. The critical frequency is where the effective bending wave
speed equals the speed of sound in the air. The crossing points 1 and 2 are
explained in the text.
238 Sound Insulation in Buildings
8.7.2 Resonant transmission
The dispersion function for the transverse waves also has consequences
for the natural frequencies of the structural modes. Assuming simply sup-
ported edges, the natural frequencies fmn are determined as in Equation 6.6,
but now using the wavelength λB,eff = cB,eff /f:
2 2 2 2
2 2fmn m n (8.62)
λ = c = l + l
B,eff B,eff x y
However, the solution for the natural frequencies is not simple because of
the dispersion function of cB,eff.
Figure 8.24 shows a geometrical representation of Equation 8.62, each
dot represents a structural mode and the frequency is represented by the
radius 2/λB,eff. The statistical number of modes below a certain frequency f
is the number of dots within the quarter circle with the radius correspond-
ing to that frequency:
Airborne sound transmission through single constructions 239
n/ly
n:
4
2/λB
3
m: 1 2 3 4 5 6 m/lx
1/4π ⋅ ( 2 / λB,eff )
2
f2
N= = π l xl y (8.63)
(1 / l x ) ⋅ (1 / l y ) cB2 ,eff
With the area S = lx · ly, the modal density is found by differentiation with
frequency:
∆N f f dc
= 2π S 2 1 − ⋅ B,eff (8.64)
∆f cB,eff cB,eff df
For pure bending waves, we have cB ~ √f, while the speed is constant for
shear waves.
The resonant transmission depends on the modal density and by inser-
tion in Equation 8.30, we get:
c 2σ res
2
∆N
Rr = R0 − 10 lg ⋅
2η Sf ∆f
(8.65)
2 2
πσ res c
= R0 − 10 lg ⋅ − ∆Rc (dB) (f > fc )
2ηtot cB,eff
where the correction term ΔRc is determined from the slope of the disper-
sion curve:
f dc
∆Rc = 10 lg 2 1 − ⋅ B,eff (dB) (8.66)
cB,eff d f
240 Sound Insulation in Buildings
This term takes values between 0 dB and 3 dB, the former referring to pure
bending waves and the latter referring to pure shear waves.
The sound radiation from the resonant modes depends on the speed of
the transverse wave relative to the speed of sound in air (see Equation 6.49).
This leads to the radiation efficiency above the critical frequency:
−1/2
2
c
σ res ≅ 1 − ( f > fc ) (8.67)
cB,eff
This is close to unity, σres ≈ 1 for cB,eff ≫ c. Insertion in Equation 8.65 gives
the following result for the transmission loss above the critical frequency:
2ηtot c
2
Rr ≅ R0 + 10 lg + 10 lg B,eff − 1 − ∆Rc (dB) ( f > fc ) (8.68)
π c
8.7.3 Forced transmission
The transmission loss for forced transmission through a sandwich panel
can be calculated as for a homogeneous panel (Equation 8.25). The coinci-
dence dip around the critical frequency can be quite pronounced if c S ≫ c.
However, some sandwich constructions have c S < c, and this means that the
coincidence dip is more shallow or disappears completely.
Inserting the wall impedance Equation 8.59 in Equation 8.25 yields the
transmission loss for forced transmission:
22 22 4 2 22
2
c⋅ B1 − ccBBf + cc2BfcS2+ c cS 2 + η 2
4
f11
44
= +
f11 − ⋅
R R 10 lg 1
Rf = R0f + 100lg 1 − ⋅ 1− 4 ⋅ 4 4 4 4 24 2 2+ η
cB ++ccBfcS + c cS2 tot
tot
f f c cBc+ cBf
σ for ()2(dB)
−10lg−( 210lg σ for ) (dB) (8.69)
8.7.4 Dilatational resonance
So far, the core material has been assumed to be incompressible. However,
for porous core materials like foam and mineral wool, the core is com-
pressible and dilatational resonances may occur (Figure 8.21e). The first
Airborne sound transmission through single constructions 241
80
f11 fc fdil
70
60
50
Transmission loss, R, dB
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency, Hz
Measured R0 Rf Rr
1 4Ect
fdil = (8.70)
2π hc ( 2mf + mc / 3)
where Ect is the compression modulus of the core material given by:
Ec
Ect = (8.71)
3 (1 − 2µ )
242 Sound Insulation in Buildings
2 2
f
Rf ,dil
= Rf + 10 lg 1 − 2
+ ηtot (dB) (8.72)
fdil
In summary, sandwich panels can be divided into two types. One type
has a core material that provides a shear wave speed less than the speed
of sound in air. Then, the critical frequency is high, and the transmission
loss is controlled by the forced transmission and the result can be close
to the mass law. However, the dilatational resonance may spoil the sound
insulation. The other type has a core with a high shear modulus and the
shear wave speed exceeds the speed of sound in air. The critical frequency
is low and the transmission loss is controlled by resonant transmission.
The mass law makes an upper limit for the obtainable sound insulation of
sandwich plates and even small alterations in the core material may lead
to major deteriorations of the sound insulation.
Another example of the transmission loss for a sandwich plate with a
relatively stiff but compressible core of mineral wool is given in Figure
8.26. Here, c S /c < 1 meaning that the free transverse vibrations propa-
gate at subsonic speed and, thus, the resonant transmission is negligible.
In such constructions, one or more dilatation resonances may occur as
the masses of the face plates oscillate with the core material as a spring
(Figure 8.21e). The fundamental dilatational resonance is clearly seen in
the example in Figure 8.26, butthere is also a dip at a higher-order dilata-
tional resonance. Although not included in the current calculation model,
this second dip can be associated with the thickness of the core hc being
equal to half a wavelength of longitudinal waves in the core material, the
Airborne sound transmission through single constructions 243
80
70
fdil
60
50
Transmission loss, R, dB
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency, Hz
Measured Calculated R0, dB
Table 8.2 Examples of typical data for core materials in sandwich panel
Young
Density, Modulus, Shear Modulus,
Material ρm (kg/m3) E (106 N/m2) G (106 N/m2)
Urethane foam 33–72 7–19 2.1–3.7
Polystyrene foam 33–42 9–26 8.5
Honeycomb, aluminium 45–123 482–368 –
Honeycomb, plastic 64–149 248–1360 –
Plastic foam 30 6.2 2.1
Glass wool, 120 0.34–0.48 –
fibres ≠ surface
Glass wool, fibres ┴ 120 <6.5 9
surface
Glass wool, fibres ┴ 80 <1.9 3.6
surface
Glass wool, fibres ┴ 45 <1.2 1.3
surface
Source: Bodlund, K. Luftlydisolering. En sammanstälning av tillämplig teori, (Airborne sound
insulation. A collection of applicable theory, in Swedish). Rapport R60:1980, Statens
råd för Byggnadsforskning, Stockholm, 1980.
REFERENCES
247
248 Sound Insulation in Buildings
m1 m2
fc1 fc2
d
R1 R2
ρd cd
P1 P2I
P2b
Figure 9.1 Sound transmission through a double construction can be divided into one con-
tribution via the cavity and another contribution via connections (sound bridges).
(Reproduced from Kristensen, J., Rindel, J.H., Bygningsakustik – Teori og praksis.
(Building acoustics – Theory and practice, in Danish). SBI-Anvisning 166. Danish
Building Research Institute, Hørsholm, Denmark, 1989. With permission.)
(
R = 10 lg cos(kd) − n ( K1 + K2 ) sin(kd)
2
(9.1)
1
2
+ (n + 1 / n)sin(kd) + ( K1 + K2 ) cos(kd) − 2nK1K2 sin(kd) + η02
2
where
ρc
n= is the ratio of impedance of the air outside the wall and the air
ρ d cd
in the cavity
ω m1
K1 = is the normalized wall impedance for plate 1
2ρ c
ω m2
K2 = is the normalized wall impedance for plate 2
2ρ c
Normally, there is no need to include all terms in Equation 9.1. For most
double constructions, only two terms (and the loss factor) are of importance:
It is seen from this that R will have a minimum if the two important terms
become numerically equal, which gives the following equation:
Airborne sound transmission through double constructions 249
ω d m1 + m2 ρ d cd
tan(kd) = tan = ⋅ (9.3)
cd m1m2 ω
1 ρ d cd2 1 1 (9.4)
f0 = +
2π d m1 m2
1000
f0 fd
100
Panel distance d (mm)
m1 = m2(kg/m2)
5
10
10
20
50
10
0
20
0
1
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 9.2 Resonance frequency f0 and the knee frequency fd as functions of the cavity
depth for a symmetrical double construction with air in the cavity. The knee
frequency involves only the cavity depth, whereas the resonance frequency also
depends on the mass per unit area of the two plates. The indicated example
is for d = 200 mm and m1 = m2 = 5 kg/m2 , which yields f0 = 85 Hz and fd = 275 Hz.
250 Sound Insulation in Buildings
cd
fd = (9.5)
2π d
Above this frequency, there are natural frequencies that are solutions to
Equation 9.3 as tan(kd) ≈ 0. These frequencies represent standing waves in
the cavity orthogonally on the two plates, and in principle this gives a num-
ber of dips in the frequency curve for the transmission loss. In practise,
however, these dips cannot be seen as the sound incidence not only occurs
orthogonally but also at diffuse sound incidence. For f > fd , the maximum
transmission loss is achieved by the insertion of cos (kd) ≈ 0 and sin (kd) ≈ 1
when calculating the transmission loss. These are good approximations
when the cavity is attenuated with absorbing material.
R = 20 lg ( K1 + K2 ) (f < f0 )
(1+ 2)
f
R ≅ R1 + R2 + 20 lg 2n ( f0 < f ≤ fd ) (9.7)
fd
R + R + 20 lg 2n
1 2 ( ) (f > fd )
80
70
fd it y
lim vit
p er ca
Up d
60 pe
dam
y y
rtl vit
Pa ca
Transmission loss, R (dB)
50 p ed
m
y da
kl
f0 ea
40 W
t
imi
er l
30 Low
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
m + m2 f
R(1+ 2) ≅ 20 lg 1 + 20 lg − 7 dB (9.8)
1 kg m −2 100 Hz
f
R1 + R2 + 20 lg + 6 dB ( f0 < f < fd )
R≅ fd (9.9)
R1 + R2 + 6 dB ( f ≥ fd )
d ⋅U
R ≅ R1 + R2 + 10 lg α (9.10)
S
80
70
it
100 mm im Empty
60
erl
p
50 mm Up
Transmission loss, R (dB)
50
40
t
imi
30
wer l
Lo
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 9.4 Measured transmission losses of a double wall with 6.4 mm and 3.2 mm
hardboard plates and a cavity of 160 mm without sound bridges. Curve 1: No
absorption in the cavity, curves 2 and 3: attenuated with 50 mm glass wool
and 100 mm glass wool, respectively. Upper and lower limits are indicated.
(Adapted from Sharp, B.H., Noise Control Eng., 11, 53–63, 1978.)
Airborne sound transmission through double constructions 253
where
α is the sound absorption coefficient of the boundary in the cavity
d is the depth of the cavity
U is the perimeter of the boundary
S is the area of the plates
0.5 for d ≤ 20 mm
α ≅ 10 mm (9.11)
for d > 20 mm
d
80
70
it
r lim
pe
60 Up
Empty
Transmission loss, R (dB)
50
150 mm
40
50 mm
t
imi
er l
30 Low
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 9.5 Measured transmission losses for a similar wall as in Figure 9.4, but with
frame-absorber of glass wool (48 kg/m3). The thickness of the frame-absorber
is either 150 mm or 50 mm. (Adapted from Sharp, B.H., Noise Control Eng.,
11, 53–63, 1978.)
254 Sound Insulation in Buildings
The use of Equation 9.10 is restricted to the range between lower and upper
limit as indicated in Figure 9.3. In Figure 9.4, some examples of measured
transmission losses can be seen with and without sound absorbing material
in the cavity. Corresponding results with a frame-absorber can be seen in
Figure 9.5.
P1
R = 10 lg = −10 lg 10− R /10 + 10− Rb /10 (9.12)
P2 + P2b
Rb = R(1+ 2) + ∆ Rm(9.13)
2
v12 Z1 + Z2
2
= (9.14)
v 2,0 Z1
V1,0
V1
V2 V2,0
Figure 9.6 Transmission of bending waves from plate 1 to plate 2 through a sound bridge.
(Reproduced from Kristensen, J., Rindel, J.H., Bygningsakustik – Teori og praksis.
(Building acoustics – Theory and practice, in Danish). SBI-Anvisning 166. Danish
Building Research Institute, Hørsholm, Denmark, 1989. With permission.)
2
v12+ 2 m1
= (9.15)
v12 m1 + m2
v12+ 2 ρ cS
∆ Rm = 10 lg
P2b
(9.16)
v 2 ρ cS m1 Z + Z2
= 10 lg 2,0 + 20 lg ⋅ 1
P 2b
(m + m
1 2) Z1
9.2.1 Point connections
The point excitation of a plate results in both a near-field radiation and a
resonant radiation, which is caused by the natural vibrations in plate 2. For
simplicity, the resonant radiation is disregarded here, but it will be dealt
with later. The sound power radiated from a number of equal excitation
points Np is (Cremer and Heckl, 1967, p. 471):
8c 2
P2b ≅ N p P2n = ρ cv2,0
2
N p (9.17)
π 3 fc22
256 Sound Insulation in Buildings
F 4c 2m
Zo = = 8 mB = (9.18)
vo π fc
Applying this for plates 1 and 2 (and inserted in Equation 9.16) yields:
Here, the last term can be combined with fc2 from the first term to yield an
auxiliary quantity fcp, which is a combination of the critical frequencies of
the two plates applying particularly for point connections:
m1fc 2 + m2 fc1
fcp = (9.20)
m1 + m2
and thus:
Sπ 3 fcp2
∆Rm, p ≅ 10 lg 2
(9.21)
N p 8c
∆ Rm, p ≥ 0 dB(9.23)
Airborne sound transmission through double constructions 257
80
70
fd it
lim
p er
60 Up
Transmission loss, R (dB)
50
ΔRm
f0
40
mit
er li
30 Low
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 9.7 Transmission loss for a double construction with rigid sound bridges.
9.2.2 Line connections
The sound transmission via line connections can be discussed analogously
to point connections. Instead of the point impedance, the line impedance is
used for each of the plates where the expression for a beam with excitation
in a point can be found (Cremer and Heckl, 1967, p. 273):
f (9.24)
Z = 2(1 + j)mcL
fc
2c
P2b ≅ P2n = ρ cv 2,0
2
L (9.25)
π fc 2
258 Sound Insulation in Buildings
30
Hz
000
=10
Correction for sound bridges ∆R (dB ) f c,p
Hz
00
20 50
Hz
00
20
Hz
00
10
10
z
0H
50
z
0H
20
0
0.1 0.2 0.5 1 2
Distance between connections e (m)
Figure 9.8 Influence of rigid point connections on the transmission loss. The connections
are assumed to have a square pattern with the centre distance e between the
points. The equivalent critical frequency fcp can normally be approximated by
the critical frequency of the thinner of the plates in the double construction.
2
m1 fc 2 + m2 fc1
fc = (9.26)
m1 + m2
the following result applies for the sound transmission through a stiff line
connection:
S π fc
∆Rm, ≅ 10 lg (9.27)
Lt 2c
When comparing this with the result for point connections (Equation 9.22),
a slightly lower dependency of the critical frequency (3 dB/octave instead
Airborne sound transmission through double constructions 259
1 f
Zr = (1 + j)mcLr (9.29)
2 fc
c
P2b ≅ P2n = ρ cv 2,0
2
Lr (9.30)
π fc 2
30
Correction for sound bridges ∆R (dB)
20
z
0H
1000
f c,l = z
0H
500
z
0H
200
10 z
0H
100
Hz
500
Hz
200
0
0.1 0.2 0.5 1 2 5
Distance between connections b (m)
Figure 9.9 Influence of rigid line connections on the transmission loss. The connections
are assumed to have the centre distance b between the lines. The equivalent
critical frequency fcl can normally be approximated by the critical frequency
of the thinner of the plates in the double construction.
260 Sound Insulation in Buildings
S π fc
∆Rm, r ≅ 10 lg (9.31)
Lr c
30
Correction for sound bridges ∆R (dB)
z
0000 H
f c,l =1
20 z
0H
500
z
0H
200
z
0H
100
Hz
10 500
Hz
200
0
0.1 0.2 0.5 1 2 5
Area to circumference ratio S/U (m)
Equation 2.57 and the dissipated power is expressed through the loss factor
by Equation 7.3:
2
v 2,0 Re {Z2 } = 2π fm2S v 22 η2 (9.32)
where η2 is the total loss factor for plate 2. The resonant radiated sound
power is:
2
v 2,0 Re {Z2 } ρ cσ 2
2
P2r = v 2,0 ρ cSσ 2 = (9.33)
2π fm2η2
The near-field radiation depends on the type of excitation; please refer to
the overview given in Table 9.1. For a point connection, the near-field radi-
ation (cf. Equation 9.17) is:
8c 2
P2n = ρ cv 2,0
2
(9.34)
π 3 fc22
The totally radiated sound power for a single point connection is:
P2b = P2n + P2r = P2nκ p(9.35)
where κp is inserted as a factor for the resonant radiation. For point connec-
tions, this becomes:
πσ 2 fc 2
κ p = 1+ (9.36)
4η2 f
Table 9.1 S ome quantities related to the excitation of a plate by point, line or
boundary connections
Point Line Boundary
Type of excitation
4c 2 m f f
Impedance Z0 2(1 + j)mcL 1
2 (1 + j)mcLr
π fc fc fc
π fc 1− j fc 1− j fc
Admittance Y0
4c 2 m 4mcL f mcLr f
Near-field 8c 2 2c c
ρ cv02 ⋅ Np ρ cv02 ⋅ L ρ cv02 ⋅ Lr
radiation P2n π 3 fc2 π fc π fc
Resonant radiation πσ fc σ fc σ fc
1+ 1+ 1+
factor κ 4η f 2η f 4η f
262 Sound Insulation in Buildings
σ2 fc 2
κ = 1+ (9.37)
2η2 f
σ2 fc 2
κr = 1+ (9.38)
4η2 f
9.2.5 Elastic connections
Two plates connected with a sound bridge as shown in Figure 9.6 are con-
sidered. The mechanical connection can be of point or line shape, and it is
presumed to be elastic. The sound bridge is characterized by the mechani-
cal impedance Zb:
F k
Zb = ≅ d (9.39)
v1,0 − v2,0 jω
Here, F is the force whereby the sound bridge affects the plates, v 1,0 and v 2,0
are the velocities of the two plates at the point of the sound bridge. Zb will
primarily depend on the dynamic stiffness kd of the sound bridge, whereas
the internal losses can often be disregarded. However, the latter does not
apply to sound bridges of viscous elastic materials.
A bending wave propagating in plate 1 with a velocity amplitude v 1 will
get a slightly reduced velocity v 1,0 opposite the sound bridge:
where Y1 is the mechanical admittance of the plate, i.e. the reciprocal of the
impedance Z1.
In plate 2, the velocity opposite the sound bridge will be determined by
the admittance Y2 of the plate.
Y2
P12 = v12 for Zb → ∞ (9.43)
(Y1 + Y2 )2
The effect of an elastic sound bridge will, thus, be described by a coupling
factor γ, which is the ratio between the transferred power via an elastic
sound bridge and a totally stiff sound bridge, respectively.
γ =
(Y1 + Y2 )2
Zb
2
2 (9.44)
1 + 2 Re {Zb }(Y1 + Y2 ) + Zb (Y1 + Y2 )
2
−1
2π f
2
γ ≅ 1 + (9.45)
kd (Y1 + Y2 )
In the latter formula, the internal losses of the sound bridge are disregarded,
as Equation 9.39 has been inserted.
9.2.6 Coupling factor
In the following, it is shown how the coupling factor γ can be determined
for three types of elastic sound bridges, point connections, line connections
and boundary connections. Indices p, l and r are used, respectively.
The calculations are based on Equation 9.45, and the dynamic stiffness
of the sound bridge is considered, but the internal losses are not. By insert-
ing the admittances as given in Table 9.1, it is found that the coupling fac-
tors can be written as:
(
γ p = 1 + ( f fkp ) )
2 −1
(9.46)
(
γ = 1 + ( f fk ) )
3 −1
(9.47)
( 3 −1
γ r = 1 + ( f fkr ) ) (9.48)
Here, some knee frequencies are inserted, below which the sound bridges
behave as stiff connections:
fcp
fkp = kdp (9.49)
8c 2mm
264 Sound Insulation in Buildings
2/3
1 fc
fk = kd′ (9.50)
2 2π cmm
2/3
1 2 fc
fkr = kdr
′ (9.51)
2 π cmm
The dynamic stiffness of the point connection is kdp and k′dl and k′dr are
the dynamic stiffness’s per unit length for line connection and boundary
connection, respectively. In these expressions, some auxiliary terms are
used, fcp (Equation 9.20) and fcl (Equation 9.26) and an equivalent mass per
unit area mm , which is determined by the mass per unit area of each of the
two plates:
m1m2
mm = (9.52)
m1 + m2
8c 2 2c
∆Rm = −10 lg 3 2 N pγ pκ p + ( Lγ κ + 12 Lrγ rκ r ) (9.53)
Sπ fcp Sπ fc
When this quantity has been determined, the transmission loss is found
from Equations 9.13 and 9.12. Examples of measured transmission loss
for light double walls with different types of sound bridges are shown in
Figure 9.11.
Heavy double constructions are especially sensitive to sound bridges,
as the critical frequency typically is rather low, which leads to a signifi-
cant resonant sound radiation. Figures 9.12 and 9.13 show examples of the
importance of sound bridges due to boundary coupling and point connec-
tions for double walls of clinker concrete.
Finally, Figure 9.14 shows an example of the use of secondary wall, i.e.
a double construction made by a massive wall and a light plate, which
reduces the sound radiation because of a high critical frequency. It is seen
that there is a good improvement of sound insulation even with rigid con-
nections, but the result is much better without sound bridges.
Airborne sound transmission through double constructions 265
80
70
60
Transmission loss, R (dB)
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 9.11 Measured transmission loss for double walls of 13 mm gypsum board with
106 mm cavity depth. Wall 1: No attenuation in the cavity, no sound bridges;
wall 2: 50 mm of mineral wool in the cavity, no sound bridges; wall 3: rigid line
connections of wooden laths, b = 400 mm; wall 4: elastic line connections of
steel laths, b = 400 mm. (Adapted from Northwood, T.D., Transmission loss
of plasterboard walls. Build. Res. Note No. 66. National Research Council
of Canada, Ottawa, ON, 1970.)
80
70 Leca, 150 mm
Plaster, 10 mm
60 Mineral wool, 50 mm
Leca, 100 mm
Transmission loss, R (dB)
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 9.12 Measured transmission loss for a single wall and a double wall with a rigid
boundary connection. Solid line: Single 150 mm clinker concrete wall with plas-
ter; dashed line: double wall, 150 mm and 100 mm clinker concrete on com-
mon concrete foundation. (Adapted from Vigran, T.E., Lydisolasjon i bygninger.
(Sound insulation in buildings, in Norwegian). Tapir, Trondheim, Norway, 1979.)
The results can be explained by the normal modes of the gypsum board
in the sections between the studs. The boundary condition also plays a
role here, and it may be something between simply supported and clamped
depending on the construction details. The natural frequency of the fun-
damental mode (m, n) = (1, 1) can be estimated from Equation 6.7 with an
extra boundary condition factor γ:
c2 1 2 1 2
f11 = γ + (9.53)
4 fc lx ly
Here, c is the speed of sound in air, fc is the critical frequency and the
dimensions of the plate section are lx and ly. The factor γ is unity for sup-
ported plates but for a clamped boundary, it is a complicated function of
dimensions and mode number, as shown by Timmel (1991). Typical values
Airborne sound transmission through double constructions 267
100
90
80
70
Transmission loss, R (dB)
60
50
40 Leca, 150 mm
Plaster, 10 mm
30 Mineral wool, 50 mm
Binder
20 Leca, 100 mm
Resilient filling
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 9.13 Measured transmission loss for a double wall of 150 mm and 100 mm clinker
concrete. Solid line: No boundary connections, Rw = 74 dB, dashed line:
With point connections, steel wire ties, e = 500 mm, Rw = 56 dB. (Adapted
from Vigran, T.E., Lydisolasjon i bygninger. (Sound insulation in buildings, in
Norwegian). Tapir, Trondheim, Norway, 1979.)
are in the range from 1.3 to 2, with the larger values for the fundamental
mode (1, 1).
For a gypsum board with critical frequency 2500 Hz and height 2700 mm,
the natural frequency f11 of the fundamental mode is calculated for some
examples of stud distances in Table 9.2. For the case of partly clamped
boundaries, the f 11 is very low at cc 600 mm and very high at cc 100 mm,
but at cc 350 mm, the frequency is exactly in the range from 125 Hz to 160
Hz where the severe dip in Figure 9.15 (solid line) is located. It is concluded
that stud distances around 300 mm to 400 mm should be avoided in this
type of wall.
268 Sound Insulation in Buildings
80
70
60
Transmission loss, R (dB)
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 9.14 Measured transmission loss for single and double walls. Solid line: 150 mm
clinker concrete wall with 10 mm plaster, Rw = 48 dB. Dashed line: Clinker
wall and a light secondary wall of 13 mm gypsum board with rigid wooden
line connections and 50 mm cavity with mineral wool, Rw = 54 dB. Dotted
line: The same, but secondary wall on steel lath only fixed in top and bottom,
75 mm cavity with mineral wool, Rw = 61 dB. (Adapted from Homb, A. et al.,
Lydisolerende konstruksjoner. (Sound insulation properties of windows, in
Norwegian). NBI-anvisning 28. Oslo, Norway, 1983, p. 61c.)
9.3.2 Examples
Dry walls can be designed within a wide range of sound insulation depend-
ing on the number of gypsum board layers, the cavity depth and whether the
studs are common to both sides or separated. Table 9.3 gives some examples
Airborne sound transmission through double constructions 269
80
70
60
Transmission loss, R (dB)
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Table 9.3 E
xpected transmission loss of drywalls with 13 mm gypsum boards, in
the laboratory (Rw and C 50–5000) and in the field (R′w)
Mineral
Wall Number Cavity wool
thickness of plate depth thickness Rw C50–5000 R′w
Type (mm) layers (mm) (mm) (dB) (dB) (dB)
Common 100 1 + 1 75 0 38–41 −1 ~35
studs 100 1 + 1 75 45 42–44 −3 ~40
125 2 + 2 75 45 48–50 −4 ~46
150 2 + 2 100 45 51–53 −4 ~47
175 2 + 2 125 90 56–58 −6 ~52
Separate 150 2 + 2 100 100 56–58 −5 ~51
studs 200 2 + 2 150 100 61–63 −6 ~54
275 1 + 1 250 250 60 −6 ~55
425 1 + 1 400 400 70 −4 ~60
Source: Byggforsk 524.325, Lydisolasjonsegenskaper til lette innervegger. (Sound insulation
properties of lightweight internal walls, in Norwegian). Byggforskserien, SINTEF
Building and Infrastructure, Oslo, Norway, 2000.
9.4.2 Examples
In Table 9.4, some examples of measured sound insulation of windows with
two and three layers of glass are given. The constructions are explained by
glass thickness in mm and distance between panes in mm. Laminated glass
is denoted with a slash (/), so, for example, 4/1.14/4 means two panes of
4 mm glass with an interlayer of 1.14 mm.
For evaluation of the sound insulation against road traffic noise, it is
necessary to include the spectrum adaptation term C tr. The data used for
the Table 9.4 did not include frequencies below 100 Hz, but if available, the
Airborne sound transmission through double constructions 271
80
70
60
Transmission loss, R (dB)
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Table 9.4 E
xamples of two- and three-layer windows, total thickness, expected
weighted sound reduction index (Rw) and practical sound insulation
against road traffic noise (Rw + C tr)
Total
thickness Rw + Ctr
Type (mm) Construction (mm) Rw (dB) (dB)
Two-layer 20 4 – 12 – 4 34 30
units 22 6 – 12 – 4 36 33
24 8 – 12 – 4 37 34
26 10 – 12 – 4 38 34
Two-layer 25 4 – 12 – 4/1.14/4 37 33
units with 35 6 – 20 – 4/0.76/4 41 36
laminated
glass 37 8 – 20 – 4/0.76/4 43 39
41 6/0.38/6 – 20 – 4/0.38/4 43 40
Three-layer 24 4–6–4–6–4 33 28
units 30 4–9–4–9–4 34 29
32 6 – 12 – 4 – 6 – 4 39 33
28 8–6–4–6–4 38 34
48 8 – 20 – 4 – 12 – 4 40 35
Three-layer 46 5 – 16 – 4 – 12 – 4/1.14/4 38 35
units with 51 8 – 18 – 4 – 12 – 4/0.2/4 43 39
laminated
glass 54 6/0.2/6 – 20 – 5 – 9 – 4/0.2/4 41 40
Single glass 67 5 – 42 – (4 – 12 – 4) 38 31
and 89 5 – 63 – (4 – 9 – 8) 40 35
two-layer
unit in 111 6 – 85 – (4 – 12 – 4) 43 40
separate 111 5 – 80 – (4 – 15 – 3/0.76/3) 45 38
sashes
Source: Byggforsk 533.109. Lydisolerende egenskaper for vinduer. (Sound insulation proper-
ties of windows, in Norwegian). Byggforskserien, SINTEF Building and Infrastructure,
Oslo, Norway, 2013.
REFERENCES
The impact sound pressure levels that can be obtained with homogeneous
massive plates can be described theoretically by the sound radiation from
point excitation of plates. However, supplementary attenuation of the
impact sound pressure level is normally necessary. This can be done by
soft resilient floor coverings or by elastically layered floating floors, which
can be considered double constructions. A suspended ceiling can also be
relevant in some cases.
The impact sound is the most important cause of complaints from the
dwellers in multi-storey houses, and the low frequencies below 100 Hz call
for special attention in the acoustic design.
The impact sound pressure level L n is a measurement for the sound trans-
mission into a room when the floor in another room is mechanically excited
by a standardized tapping machine:
A2
Ln = L2 + 10 lg (dB) (10.1)
A0
p 22 A2 4 ρcP2
Ln = 10 lg = 10 lg (dB) (10.2)
p02 A0 p02 A0
where p 22 is the mean sound pressure square in the receiving room, and P2 is the
sound power radiated into the receiving room. The formula in Equation 10.1 is
275
276 Sound Insulation in Buildings
the basis of measuring the impact sound pressure level in the laboratory as well
as in buildings. However, the symbol L′n is used in buildings in order to indicate
that the result may be influenced by flanking transmission.
The approximation comes by inserting the data from the tapping machine
and the acceleration of gravity, g = 9.81 m/s2. For less stiff structures like
wooden floors, the relation becomes more complicated and (Equation 10.3)
does not hold. For one-third octave bands, the relative bandwidth is 0.23
and the mean-square-force of the tapping machine on a heavy and stiff floor
is:
F 2 ≅ 0.91f (N 2 ) (10.4)
F 4c 2m
Z0 = ≅ 8 mB = (10.5)
v0 π fc
depends on the position of the point relative to the modal pattern of the
plate.
A very important relationship exists between the modal density ΔN/Δf
of a finite plate and the point impedance of the equivalent infinite plate. If
m is the mass per unit area of the plate and S is the area, the real part of
the point admittance (the reciprocal of the impedance) (see Cremer et al.,
1988) is:
1 1 ∆N
Re = ⋅ (10.6)
Z
0 4mS ∆f
The modal density of a homogeneous, isotropic plate depends on the thick-
ness. If the plate is thin compared to the wavelength, bending waves can
be assumed, but if the plate is thick, compared to the “wavelength” shear
waves will occur instead. The approximate modal density is:
π
Sfc (f < f s / 2) (bending)
∆N c 2 (10.7)
≅
∆f 2π Sf (f > f s / 2) (shear)
c s2
where fc is the critical frequency, fs = fc (cs /c) is the crossover frequency for
2
shear waves, and c s is the speed of shear waves (see Equation 1.30). Both
the general results in terms of the modal density as well as the bending wave
approximation will be stated in the following.
If φ is the phase angle between force and velocity in the excitation point,
the power input to the plate by the point force is (Equation 2.55):
1
Pi = 1
2 F ⋅ v0 ⋅ cos ϕ = v 02 Re {Z0 } = F 2 Re
Z0
F 2 ∆N F 2π fc (10.8)
= ⋅ ≅ 2
4mS ∆f 4c m
In steady state, the power Pd that is lost by internal losses, boundary losses
and radiation losses is expressed by the total loss factor η (using Equations
7.1 and 7.3):
Pd = 2π fmS v r2 η (10.9)
F 2 ∆f f
= 8π fm 2S 2η ≅ 8Sm 2c 2η (10.10)
2
v r ∆N fc
278 Sound Insulation in Buildings
The sound power P2 radiated into the receiving room depends on the radia-
tion efficiency σres and the average velocity (see Equation 6.24). From this
and Equation 10.10, the radiated sound power can be found as a function
of the input force F:
F 2 ρ cσ res ∆N F 2 ρσ res fc
P2 = ⋅ ≅ (10.11)
8π Sm 2η f ∆f 8m 2cη f
The sound power radiated from the near field around the excitation point
may also be included using the point impedance (Equation 10.5) together
with Equation 9.17, which leads to:
F 2 ρ F 2 ρ cσ res ∆N
P2 = + ⋅ (10.12)
2π m 2c 8π Sm 2η f ∆f
The first term is due to the near field radiation and the second term is the
resonant contribution from the natural modes in the plate.
4 ρ 2 F 2 1 c 2σ res ∆N
Ln = 10 lg 2 2
+ ⋅ (10.13)
p0 A0 m 2π 8π fSη ∆f
Insert the numerical constants and the force of the standardized tapping
machine per one-third octave, and the impact sound pressure level per one-
third octave is:
4 c 2σ res ∆N
Ln ≅ 82 − 20 lg m + 10 lg f + ⋅ (10.14)
π π Sη ∆f
where m shall have the unit kg/m 2 . The last term represents the near field
radiation plus the resonant radiation from the natural modes. When the
loss factor η is small, the contribution of the near field is often insignificant
compared to the resonant radiation. For massive floors with a low critical
frequency, the radiation efficiency is σres ≈ 1 because f > fc. Assuming bend-
ing waves, an approximation for the impact sound pressure level is:
η
Ln ≅ 82 − 10 lg m 2 ( fc < f < fs / 2) (10.15)
fc
Impact sound insulation 279
Above the frequency fs /2, the vibrations in the plate propagate as shear
waves and the approximation changes to:
η 2f
Ln ≅ 82 − 10 lg m 2 + 10 lg (f > fs / 2) (10.16)
fc fs
If the frequency dependency of the loss factor is disregarded, these approxi-
mations show that the impact sound pressure level for a massive floor is
independent of the frequency in the bending wave range, but increases with
frequency at higher frequencies where the shear waves take over.
The influence of the plate thickness is seen from the fact that m is pro-
portional to the thickness and fc and fs are inversely proportional to the
thickness. The impact sound pressure level L n decreases 9 dB per doubling
of the plate thickness in the bending wave range, but decreases 12 dB per
doubling of the plate thickness in the shear wave range. For a rough esti-
mate, the relationship is shown for concrete and lightweight concrete in
Figure 10.1. However, it is not realistic to use so thick plates that the impact
sound pressure level is acceptable for dwellings without further attenuation
of the impact sound.
100
Impact sound pressure level per 1/3 octave (dB)
90
Por
80 ous
con
cret
e
70
Con
cret
e
60
50
40
30
20
50 100 200 500 1000
Plate thickness h (mm)
Figure 10.1 Approximate impact sound pressure level in one-third octave band for
massive floor of concrete (2300 kg/m3) or of porous concrete (600 kg/m3)
assuming bending waves, a radiation factor σ = 1, and a loss factor η = 0.05.
280 Sound Insulation in Buildings
With a carpet or some other thin flooring, e.g. linoleum with soft underlay,
a reduction of the hammer force of the floor is obtained by decreasing the
speed gradually from maximum at the surface to zero when a certain pres-
sure is reached in the flooring, after which the hammer returns to the sur-
face level. The course corresponds to half a cycle for a mass-spring system
where the spring constant is determined by the modulus of elasticity of the
flooring, its thickness ht and the actuated area Sh (Figure 10.2).
If the mass of the hammer is called mh, the resonance frequency of the
system can be determined by:
1 Sh Et (10.17)
f0 =
2π mh ht
For the standardized tapping machine, mh = 0.500 kg and Sh = 700 mm 2 .
The diagram in Figure 10.3 shows a relationship between the resonance
frequency f 0 and the elastic properties of the flooring. For many elastic
materials, the dynamic modulus of elasticity is about double that deter-
mined by a static load.
As shown by Ver (1971), if the resonance frequency f 0 is known, and
f > f 0 the attenuation of the impact sound pressure level can be determined
approximately for f > f 0:
f
∆L = Ln,without − Ln,with ≅ 40 lg (10.18)
f0
For f ≤ f 0, the attenuation is ΔL ≅ 0 dB. The idealized frequency course
of the impact sound attenuation is illustrated in Figure 10.4. For a good
design, f 0 should be as low as possible and preferably under 90 Hz. In
Figure 10.5, some examples of measured impact sound attenuation of thin
flooring can be seen.
mh Sh Et
ht
Figure 10.2 Resilient floor covering with thickness ht and Young’s modulus Et. The ham-
mer of the tapping machine has the mass mh = 500 g and the area Sh = 700 mm2 .
(Reproduced from Kristensen, J., Rindel, J. H., Bygningsakustik – Teori og
praksis, SBI-Anvisning 166 (In Danish), Danish Building Research Institute,
Aalborg University Copenhagen, 1989. With permission.)
Impact sound insulation 281
1.E+11
1.E+10
Et/ht (N/m3)
1.E+09
1.E+08
100 200 500 1000
Resonance frequency f0 (Hz)
Figure 10.3 Diagram for estimation of the resonance frequency f0 for a soft floor cover-
ing with thickness ht and Young’s modulus Et.
40
30
20
ΔL, dB
10
–10
0.5 1 2 5 10
Relative frequency f/f0
Figure 10.4 Idealized frequency course of the impact sound attenuation of a soft floor
covering.
282 Sound Insulation in Buildings
70
60
50
Improvement, ΔLn (dB)
40
30
20
10
–10
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 10.5 Examples of impact sound attenuation for floor coverings. Solid line: Thin
linoleum. Dashed line: Vinyl on a layer of felt. Dashed–dotted line: Thick
and soft carpet.
c 2σ res ∆N
Ln ≅ 82 − 20 lg m + 10 lg ⋅ − ∆L (10.19)
π Sη ∆f
2
ω m c 2σ res
2
∆N
R ≅ Rr = 10 lg − 10 lg ⋅
2ρ c 2f S η ∆f
c 2σ res
2
∆N
( )
= 10 lg m 2 f 3 − 10 lg ⋅
π S η ∆f
− 44 (10.20)
Impact sound insulation 283
R + Ln = 38 + 30 lg f − 10 lg σ res − ∆L (10.21)
R + Ln = 38 + 30 lg f (10.22)
With a soft flooring or a floating floor that is locally reacting, the relation-
ship changes above the resonance frequency f 0 of the flooring:
f
R + Ln = 38 + 30 lg f0 − 10 lg
f0
(f > f0 ) (10.23)
(a) 80
f0
70
60
50
R΄and L΄n (dB)
40
30
20
10
f11
0
20 50 100 200 500 1000 2000 5000
Frequency (Hz)
R΄ L΄n
(b) 140
e
av
130 oct
per
B
f0 9d
120
Sum R΄+ L΄n (dB )
110
–3 d
B pe
r oct
ave
100
90
80
20 50 100 200 500 1000 2000 5000
Frequency (Hz)
Heavy Theory Resilient floor
Figure 10.6 (a) Measured airborne and impact sound insulation of typical heavy floor
constructions. Average of 10 measurements. The frequency of the lowest
structural mode f11 and the resonance frequency of the flooring f 0 are indi-
cated. (Below 50 Hz the results are not corrected for reverberation time.)
(b) The sum R′ + L′n for a massive floor. Thick line: The sum of the measured
results in a. Thin straight line: Idealized frequency course for a hard floor.
Dashed line: Floor with impact sound attenuation by flooring. (Data from
DELTA, Målerapport: Boligers lydisolation ved lave frekvenser. (In Danish). PDØ
870028, DELTA, Aarhus, Denmark, 1999.)
Impact sound insulation 285
Figure 10.7 Transmission of airborne and impact sound through floors with floating
floors. Due to flanking transmission, the attenuation of impact sound is bet-
ter than the attenuation of airborne sound. (Reproduced from Kristensen,
J., Rindel, J. H., Bygningsakustik – Teori og praksis, SBI-Anvisning 166 (In
Danish), Danish Building Research Institute, Hørsholm, Denmark, 1989.
With permission.)
1
Pi = F 2 Re (10.24)
Z
0,1 + Zh
where Z 0,1 is the point impedance of the floating floor slab. The impedance
Zh of the hammer is determined by its mass mh = 500 g:
Zh = jω mh (10.25)
2m1 S1 ∆f
f > fZ = ⋅ (10.26)
π mh ∆N1
286 Sound Insulation in Buildings
Zh (mh)
F
Z1 (m1 , fc1)
d Zd (em, cm)
Z2 (m2 , fC2)
Figure 10.8 Floating floor with excitation from a falling mass. Each element is character-
ized by a mechanical impedance. (Reproduced from Kristensen, J., Rindel, J.
H., Bygningsakustik – Teori og praksis, SBI-Anvisning 166 (In Danish), Danish
Building Research Institute, Hørsholm, Denmark, 1989. With permission.)
Here, the point admittance has been inserted from Equation 10.6. For fre-
quencies higher than f Z , an extra contribution ΔL Z for the attenuation of
the impact sound pressure is obtained because of the impedance of the
hammer:
f 2
∆LZ = 10 lg 1 + (10.27)
fZ
For f < f Z , the attenuation is ΔLZ ≅ 0 dB. However, for lightweight floors this
attenuation has some importance. For example the frequency f Z is ≅ 700 Hz
for a 22 mm chip board and ≅ 200 Hz for a 12 mm chip board. Above f Z ,
the attenuation increases by 6 dB per octave in the case of a thin slab per-
forming bending waves, changing to 12 dB per octave in the case of a thick
slab performing shear waves. For measurements in the laboratory of impact
sound attenuation of floating floors, the measurement standard ISO 10140-1
(Annex H.4.1) prescribes an evenly distributed load of 20 kg/m 2 to 25 kg/m2 ,
which will diminish the problem of the hammer impedance.
The attenuation of the impact sound pressure level of a floating floor can
be expressed by the ratio of the velocities in the floor in the two situations
shown in Figure 10.9a and b:
v2 a
∆L = 20 lg (10.28)
v2b
By means of the mechanical impedances Z1 and Z 2 for the two plates, the
result is:
Impact sound insulation 287
Figure 10.9 Point force excitation of a slab (a) and a floating floor on the same slab
(b). (Reproduced from Kristensen, J., Rindel, J. H., Bygningsakustik – Teori og
praksis, SBI-Anvisning 166 (In Danish), Danish Building Research Institute,
Hørsholm, Denmark, 1989. With permission.)
F
v2 a = (10.29)
Z2
F
v1b = (10.30)
Z1
The force F′, which is transferred through the elastic layer from plate 1 to
plate 2, can be expressed as follows:
F' Z FZd
v2b = ≅ v1b d = (10.32)
Z2 Z2 Z1Z2
1 1 1
f0 = k ''d + (10.34)
2π m1 m2
where k″d is the dynamic stiffness per unit area of the elastic layer. This
equation is the same as Equation 9.4 for a double wall, except that here the
dynamic stiffness is not primarily due to the air in the cavity. In many floor
constructions, m1 ≪ m2 and, thus, the latter can be removed from Equation
10.34.
The dynamic stiffness per unit area is a combined contribution from the
structure of the elastic material and from the air contained in the pores of
the elastic material:
ρ mcm2 ρ c 2
k ''d = k ''m + k ''a = + (10.35)
d qd
Here, d is the thickness of the elastic layer and q is the porosity of the mate-
rial, i.e. the ratio between the pore volume and the total volume. ρm and cm
are the density of the material and the speed of longitudinal sound waves,
respectively.
An elastic material is characterized by the dynamic modulus of elasticity
ρmcm2 , which is the same as the stiffness per unit area per thickness unit.
Typical values for a selection of elastic materials are stated in Table 10.1.
Except from light mineral wool, the contribution from the air is often of
minor importance. The dynamic stiffness can be determined experimen-
tally after ISO 9052-1 by measuring the resonance frequency with a load
of known mass.
the material. If the results are corrected for the contribution from the stiff-
ness of the air, the following empirical relationship is found:
where m1 is the static load. Because it is mainly the speed of sound cm that
can change with different loads, Equation 10.35 leads to the relationship:
cm ≅ constant ⋅ 4 m1 (10.37)
Table 10.2 gives the data for the same types of mineral wool, as in
Figure 10.10. In a special experimental setup, the sound speeds cm are mea-
sured and the calculated dynamic modulus of elasticity correspond to the
curves in Figure 10.10.
The speed of sound in mineral wool is approximately one-tenth of the
speed of sound in air, which is of significance for the frequency depen-
dency of the impact sound attenuation. The sound propagation in an elas-
tic material, such as mineral wool, will be accompanied by losses. As
shown by Cremer and Heckl (1967, p. 169), this can be formulated by a
complex sound speed whose imaginary quantity is determined by the loss
factor η:
1
1
Dynamic elasticity module (MN/m2)
0.5 2
0.2
0.1
0.1 0.2 0.5 1 2 5 10
Static load (kPa)
Figure 10.10 Dynamic elasticity as a function of the static load for mineral wool. The
curves 1, 2 and 3 are regression lines for the same three types of min-
eral wool described in Table 10.2. (Adapted from Gudmundsson, S., Sound
insulation improvements of floating floors. A study of parameters. Report
TVBA-3017, LTH, Lund, Sweden, 1984.)
290 Sound Insulation in Buildings
Table 10.2 T hree examples of mineral wool for floating floors (same as in
Figure 10.10)
Dynamic elasticity
Type (number) Density, ρm (kg/m3) Speed of sound, cm (m/s) (MN/m2)
Rockwool (1) 160 49 0.38
Rockwool (2) 100 35 0.12
Fibre glass (3) 50 22 0.02
Source: G
udmundsson, S., Sound insulation improvements of floating floors. A study of
parameters. Report TVBA-3017, LTH, Lund, Sweden, 1984, p. 159.
Note: The speed of sound is measured with a static load of 1 kPa.
η
cm → cm 1 + j (10.38)
2
A number of measuring results of the loss factor have been referred by
Gudmundsson (1984, p. 162), but there is much uncertainty of this param-
eter. For Rockwool loaded with 1 kPa, the loss factor has been measured to
η = 0.23–0.28. For heavier loads, loss factors have been measured (η = 0.10–
0.15 for loads of 4 kPa to 6 kPa). In order not to complicate the description
further, it is presumed in the following that the loss factor is so small that
it will not contribute to the impact sound attenuation.
Z1 = jω m1 (10.39)
where m1 is the mass of the floor per unit area. For the elastic layer, the
impedance is:
ρ mcm2
Zd = (10.40)
ω
sin d
cm
where the denominator goes towards 0 at frequencies where standing waves
in the elastic layer may occur. It will be practical to introduce a knee fre-
quency fd:
cm
fd = (10.41)
2π d
Impact sound insulation 291
At low frequencies, the spring effect is not affected by standing waves and
sin(ω d/cm) ≅ ω d/cm. At higher frequencies f > fd a useful simplification is to
set sin(ω d/cm) ≅ 1, leading to:
ρ mcm2 k ''m
ω d = ω ( f ≤ fd )
Zd ≅ (10.42)
ρ mcm ( f > fd )
At frequencies below the resonance frequency f 0, the construction does not
react as a double construction and the attenuation of the impact sound level
is insignificant. For floating floors, the expression (Equation 10.34) for the
resonance frequency (Equation 10.34) can often be simplified, as m1 < m2
and k″m > k″a.
1 ρ mcm2
f0 ≅
2π m1d (10.43)
ω 2m1d f
∆L = 20 lg
ρ mcm2
= 40 lg
f0
( f0 < f ≤ fd ) (10.44)
ω m1 f f
∆L = 20 lg
ρ mcm
= 40 lg d + 20 lg
f0 f0
(f > fd ) (10.45)
50
40
2
Improvement, ΔLn (dB)
1
30 3
fd2
20
10
fd3
fd1
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 10.11 Impact sound attenuation of floating floors with three different elastic
materials. All curves are averaged results from three floors, measured
3 months and 1 year after inauguration. All floors are built from 150 mm
concrete with 35 mm floating concrete floor. Curve 1: 15 mm mineral wool;
curve 2: 25 mm mineral wool; curve 3: 15 mm polystyrene foam plate.
(Adapted from Kuhl, W., Acustica, 51, 103–115, 1982.)
∆f 4c 2m1
Z0,1 = 4m1S ≅ (10.46)
∆N1 π fc1
The impedance for a spring element consisting of the active section of the
elastic layer can be written as:
km
Z0, d = (10.47)
ω
70
60
6 dB per octave
50
Improvement, ΔLn (dB)
40
30
20
10
–10
50 100 200 500 1000 2000 5000
Frequency (Hz)
MW type 1 MW type 2 MW type 3
π 2
km = km′′ Sa = km′′ λB1 (10.48)
16
ρ mcm2 π c 2
km = ⋅ (10.49)
d 16ffc1
128m1d 2 f
∆L = 20 lg f ≅ 40 lg ( f > f0 ) (10.50)
πρ mcm2 f0
where f 0 is the resonance frequency (Equation 10.43). The result for locally
reacting floors corresponds exactly to the previously found result for
294 Sound Insulation in Buildings
(a) (b) ½ λB
F F
Figure 10.13 Floating floors with a point force excitation. (a) Resonantly reacting floor.
(b) Locally reacting floor. (Reproduced from Kristensen, J., Rindel, J. H.,
Bygningsakustik – Teori og praksis, SBI-Anvisning 166 (In Danish), Danish
Building Research Institute, Hørsholm, Denmark, 1989. With permission.)
resonantly reacting floors below the knee frequency, but there is no knee
frequency for locally reacting floors. The slope of ΔL is 12 dB per octave in
the entire frequency range, which is in agreement with the model above for
the transmission through the elastic layer. The section of the active cylinder
is reduced by increasing frequency and, thus, the stiffness is inversely pro-
portional to the frequency (see Equation 10.49).
Examples of measured attenuation of impact sound pressure level for
locally reacting floors are seen in Figure 10.14. The types of applied mineral
70
60
50
12 dB per octave
Improvement, ΔLn (dB)
40
30
20
10
–10
50 100 200 500 1000 2000 5000
Frequency (Hz)
MW type 1 MW type 2 MW type 3
Figure 10.14 Measured impact sound attenuation of locally reacting floors made from
22 mm chipboard laid on different types of mineral wool, 50 mm thick. The
curves refer to the material numbers 1, 2 and 3, respectively, in Table 10.2.
The structural slab is 160 mm concrete. (Adapted from Gudmundsson, S.,
Sound insulation improvements of floating floors. A study of parameters.
Report TVBA-3017, LTH, Lund, Sweden, 1984. Fig. III.2.3.)
Impact sound insulation 295
wool are the same as used in Table 10.2 and Figure 10.10. The curves
increase by 12 dB per octave in agreement with Equation 10.50.
1 kdp ⋅ N p
f0 = (10.51)
2π m1S
The attenuation of the impact sound by point supported floors that are
presumed to be resonantly reacting has been analysed by Ver (1971). With
indices 1 and 2 referring to the flooring plate and the structural slab,
respectively, his result is:
Z mη Z f3
∆L = 10lg 0,1 + 1 1 + 2πη1 0,1 ⋅ 2 (10.52)
Z0,2 m2η2 kdp f0
Above a certain frequency, the last term dominates and the impact sound
attenuation increases by 9 dB per octave, if the loss factor of the floor-
ing plate η1 and the stiffness of the patches are frequency independent.
Inserting the resonance frequency (Equation 10.51) and the point imped-
ance (Equation 10.46) yields:
32π 3 m 2η ∆f f
∆L ≅ 10lg 2 ⋅ 1 1 ⋅ ⋅ f 3 ≅ 30lg (10.53)
k
dp N p ∆N 1 f0
Figure 10.15 Point supported floating floor with a point force excitation. The force is
transferred to the slab through a number of elastic patches. (Reproduced
from Kristensen, J., Rindel, J. H., Bygningsakustik – Teori og praksis, SBI-
Anvisning 166 (In Danish), Danish Building Research Institute, Hørsholm,
Denmark, 1989. With permission.)
296 Sound Insulation in Buildings
50
40
9 dB per octave
Improvement, ΔLn (dB)
30
20
10
–10
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 10.16 Measured impact sound attenuation of a point supported floor made from
22 mm wood on laths on 10 mm fibreboard pads. The structural slab is 200 mm
clinker concrete slab with 30 mm finish. Measured data from Figure 5.32.
Especially, the importance of a large loss factor and a low modal density
for the flooring plate should be noted. Thus, the flooring should preferably
be thick and with high damping.
An example of a flooring that performs in accordance with this model is
the traditional wooden floor on laths on resilient pads (Figure 10.16)
40
n=0
30
Improvement, ΔLn (dB)
20 n=1
n = 10
10
–10
50 100 200 500 1000 2000 5000
Frequency (Hz)
Compared to floating floors, the suspended ceilings have the advantage that
the resonance frequency can be significantly lower (preferably below 50 Hz)
because the gap between the slab and the suspended ceiling can be made
much wider. Still, the use of suspended ceilings for attenuation of impact
sound is not common practice.
The resonance frequency of a suspended ceiling is calculated as for an
asymmetric double construction with one heavy mass and one lighter mass:
1 ρc2
f0 ≅ (10.54)
2π m3d
Here, m3 is the mass per unit area of the suspended ceiling, d is the suspen-
sion, and ρc 2 is the dynamic elasticity of the air in the cavity.
Both perforated and closed gypsum board plates can provide attenuation
of impact sound. It is possible, at least to some extent, to combine a room
acoustic treatment of the room below with improved impact sound insula-
tion (Seidel and Hengst, 2017).
Table 10.3 O
verview of measurement results on floor constructions made from
massive wood
Construction Description Rw (dB) Rw,50 (dB) Ln,w (dB) Ln,w,50 (dB)
185 mm wood 44 43 82 75
115 mm wood 56 52 59 60
30 mm mineral
wool
185 mm wood
185 mm wood 62 59 50 57
150 mm
suspended
ceiling
2 × 13 mm
gypsum board
115 mm wood 63 59 41 53
30 mm mineral
wool
185 mm wood
150 mm
suspended
ceiling
2 × 13 mm
gypsum board
115 mm wood 60 57 50 53
30 mm mineral
wool
50 mm concrete
tiles (200 × 400)
mm
185 mm wood
(Continued)
Impact sound insulation 299
115 mm wood 63 61 33 48
30 mm mineral
wool
100 mm sand
185 mm wood
150 mm
suspended
ceiling
2 × 13 mm
gypsum board
90
80
70
60
R and Ln (dB)
50
40
30
20
10
50 100 200 500 1000 2000 5000
Frequency (Hz)
R Ln
Figure 10.18 Measured transmission loss and impact sound pressure level of a floor
of 185 mm massive wood. (Adapted from Eriksen, L., Ejlersen, C. V. J.,
Lydisolering – Etageadskillelser af massive træelementer. (In Danish). MSc
Thesis, Byg-DTU and Ørsted-DTU, Technical University of Denmark,
Lyngby, Denmark, 2003.)
80
Floating floor
70
60
50
R and Ln (dB)
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Ri Ln
Figure 10.19 Measured transmission loss and impact sound pressure level of a floor of
185 mm massive wood with a floating floor of 115 mm massive wood on
30 mm mineral wool. (Adapted from Eriksen, L., Ejlersen, C. V. J., Lydisolering –
Etageadskillelser af massive træelementer. (In Danish). MSc Thesis, Byg-DTU
and Ørsted-DTU, Technical University of Denmark, Lyngby, Denmark, 2003.)
80
Floating floor + sand
70
60
50
R and Ln (dB)
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
R Ln
Figure 10.20 Measured transmission loss and impact sound pressure level of a floor of
185 mm massive wood with a topping of 100 mm sand and a floating floor
of 115 mm massive wood on 30 mm mineral wool. (Adapted from Eriksen,
L., Ejlersen, C. V. J., Lydisolering – Etageadskillelser af massive træelementer.
(In Danish). MSc Thesis, Byg-DTU and Ørsted-DTU, Technical University
of Denmark, Lyngby, Denmark, 2003.)
302 Sound Insulation in Buildings
70
60
50
Improvement, ΔLn (dB)
40
30
20
10
–10
50 100 200 500 1000 2000 5000
Frequency (Hz)
Floating floor Suspended ceiling Floating floor + suspended ceiling Sum
80
70
Impact sound pressure level, L΄n (dB)
60
50
40
30
20
10
0
20 50 100 200 500 1000 2000 5000
Frequency (Hz)
Tapping machine Walking without shoes
Walking with shoes HT (ISO 226:2003)
Figure 10.22 Impact sound measured with the tapping machine compared with the
average sound pressure level measured with a live walker with or without
shoes. The floor is 250 mm concrete slab with 50 mm light concrete top-
ping, 17 mm resilient layer (polyurethane) and 65 mm floating concrete
floor. The hearing threshold for pure tones (HT) is shown for compari-
son. (Adapted from Wolf, M., Burkhart, C., Vermeidungsstrategien und
Ansätze einer Vermeidung des Estrichdrönens. Proceedings of DAGA 2016,
Aachen, Germany, 2016. Figure 3.)
also shows the hearing threshold (HT), although this is for pure tones
under special listening conditions. Nevertheless, this indicates that for
the actual example, the audible footfall noise is dominated by frequencies
below 100 Hz.
The relation between the spectrum from a live walker and from the tap-
ping machine depends to a great deal on the floor construction. Figure 10.23
shows the difference in spectrum for three different floors; a bare concrete
(difference increases monotonic with frequency), a floating top surface (the
difference is nearly constant around 20 dB above 100 Hz) and a floor with
carpet (maximum difference around 50 Hz to 200 Hz). A positive differ-
ence means that the tapping machine is louder than the live walker. These
results demonstrate that the tapping machine can be used effectively as a
testing device up to at least 500 Hz for all types of surface except carpet
(Warnock, 1998). With a carpeted floor, the difference curve is fairly close
to the others below 200 Hz. Since the impact sound from a carpeted floor
304 Sound Insulation in Buildings
60
50
Difference in impact sound pressure level (dB)
40
30
20
10
–10
–20
20 50 100 200 500 1000 2000 5000
Frequency (Hz)
Concrete surface Floating top surface Carpet
Figure 10.23 Difference between impact sound pressure levels measured with the
tapping machine and with a live walker with shoes, shown for three dif-
ferent floorings: Bare concrete, floating floor and carpet. (Adapted from
Warnock, A. C. C., Floor research at NRG Canada. Conference in Building
Acoustics “Acoustic Performance of Medium-Rise Timber Buildings”, Dublin,
Ireland, 1998. Figure 3.)
Complaints about impact noise in dwellings are very often related to the
thudding sound created by footfall. The low-frequency problems of impact
sound have been clearly related to lightweight, wood-frame building con-
structions (Blazier and DuPree, 1994). They found that the peak in the foot-
fall spectrum occurred at the fundamental natural frequency of the floor
construction, typically between 15 Hz and 30 Hz. The characteristic
difference between impact sound spectra for lightweight and heavyweight
Impact sound insulation 305
(
L50 −80 = 10 lg 100.1⋅L50 + 100.1⋅L63 + 100.1⋅L80 ) (10.55)
For the evaluation of this low-frequency impact sound, the following guide
(LoVerde and Dong, 2017) is suggested:
80
70
Impact sound pressure level, L΄n (dB)
60
50
40
30
20
10
0
20 50 100 200 500 1000 2000 5000
Frequency (Hz)
Heavy Light
80
70
Impact sound pressure level, L΄n (dB)
60
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Complaints Acceptable
Figure 10.25 Mean values of the impact sound pressure levels measured on site, from
10 cases of quality check and 12 cases of complaints. (Adapted from Rittig,
C., Übertragung tieffrequenter trittschallgeräusche auf massivdecken
mit schwimmendem estrich. Bachelor thesis, Building Physics, Technical
University, Stuttgart, Germany, 2013.)
Table 10.4 S ome characteristics of the average impact sound pressure levels in
cases of complaints and in cases of no complaints
L′n,w (dB) L′n,w,50 (dB) L50–80 (dB) Slope (dB/octave)
Complaints 42 50 62 −7
No complaints 37 41 53 −5
Note: Same data as in Figure 10.25.
308 Sound Insulation in Buildings
(
∆L0 = −10lg 1 + Q2 ≅ −20lg Q )
(10.56)
80
70
Impact sound pressure level, L΄n (dB)
60
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 10.26 Impact sound pressure level of a timber floor. Construction from top to
bottom: soft carpet, 22 mm chipboard, 48 mm × 198 mm wooden beams
with 200 mm mineral wool in the cavity, 48 mm × 48 mm lath with resil-
ient mounting of 2 × 13 mm gypsum board. (Adapted from Hagberg, K.,
Simmons, C., Bygg & teknik 1/97, 17–18, 1997.)
Impact sound insulation 309
80
60
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 10.27 Impact sound pressure level of a heavyweight floor with hard surface.
Construction from top to bottom: Clinker tiles on 250-mm concrete.
(Adapted from Hagberg, K., Simmons, C., Bygg & teknik 1/97, 17–18, 1997.)
80
70
Impact sound pressure level, L΄n (dB)
60
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 10.28 Impact sound pressure level of a heavyweight floor with floating floor.
Construction from top to bottom: 14 mm parquet, 3 mm ethafoam,
30 mm gypsum, 0.15 mm plastic foil, 15 mm mineral wool, 230 mm con-
crete. (Data from Multiconsult, Unpublished measurement report. Project
Number 103407, 2002.)
310 Sound Insulation in Buildings
Table 10.5 S ome characteristics of the impact sound pressure levels in cases of
complaints
Construction L′n,w (dB) L′n,w,50 (dB) L50–80 (dB) Slope (dB/octave)
Lightweight timber 44 57 72 −9
Concrete with hard surface 62 52 45 +5
Concrete with floating floor 42 56 71 −7
Note: Same data as in Figures 10.26 through 10.28.
2π m1f0
Q= (10.57)
rsd
where m1 is the mass per unit area of the floating floor, rs is the airflow
resistivity per unit length and d is the thickness of the resilient layer. Typical
values of Q for floating floors are between 2.5 and 10.
Unfortunately, it is difficult to get information about the flow resistance
from the manufacturers of resilient materials for floors; normally, only data for
the dynamic stiffness are available. The relevant measurement standards are
ISO 9052-1 for the dynamic stiffness and ISO 9053 for the airflow resistance.
REFERENCES
L.L. Beranek (1989). Balanced noise-criterion (NCB) curves. Journal of the Acoustical
Society of America 86(2), 650–664.
L.L. Beranek and I.L. Vér (1992). Noise and Vibration Control Engineering. John
Wiley & Sons, New York.
W.E. Blazier and R.B. DuPree (1994). Investigation of low-frequency footfall noise
in wood-frame, multifamily building construction. Journal of the Acoustical
Society of America 96 (3), 1521–1532.
L. Cremer and M. Heckl (1967). Körperschall. (Structure-borne sound, translated by
E.E. Ungar 1973). Springer, Berlin, Germany.
DELTA (1999). Målerapport: Boligers lydisolation ved lave frekvenser. (Measurement
report: Sound insulation of dwellings at low frequencies, in Danish). PDØ
870028, DELTA, Aarhus, Denmark.
L. Eriksen and C.V.J. Ejlersen (2003). Lydisolering—Etageadskillelser af massive
træelementer. (Sound insulation – floor constructions of massive wood slabs,
in Danish). MSc Thesis, Byg-DTU and Ørsted-DTU, Technical University of
Denmark, Lyngby, Denmark.
S. Gudmundson (1984). Sound insulation improvements of floating floors. A study
of parameters. Report TVBA-3017, LTH, Lund, Sweden.
Impact sound insulation 311
Flanking transmission
The flanking constructions and the junctions between the constructions are
of high importance for the sound insulation between rooms. The number of
transmission paths to be considered is often quite high, so there is need for
a computer program to handle this in a reasonable way. Computation pro-
grams for sound insulation normally have the drawback that their results
are the maximum values that can be achieved under conditions with perfect
workmanship and acoustically correct building details, which can seldom
be achieved in practice. Yet, computation programs are useful tools at the
design stage for sound insulation.
11.1 FLANKING TRANSMISSION
OF AIRBORNE SOUND
11.1.1 Transmission paths
In buildings with traditional heavy constructions of concrete or masonry,
about 50 % of the sound transmission between rooms with a common
dividing partition is flanking transmission. Of course, the flanking trans-
mission is 100 % between rooms without a common dividing partition.
Figure 11.1 shows the principles of flanking transmission on four differ-
ent types of building. It is important whether the constructions are heavy-
weight or lightweight because the sound radiation of free bending waves
is different. Heavyweight walls and floors have a low critical frequency
and, thus, the radiation efficiency for free bending waves is high (approxi-
mately unity). Figure 11.1a shows a building type with heavyweight walls
and floors like the precast concrete houses typical from the 1960s and the
1970s. Flanking is both vertical and horizontal, and this may have contrib-
uted to the bad reputation of sound insulation in houses from that time.
Type B with lightweight walls and heavyweight floors is not unusual today,
and flanking is mainly in the horizontal direction. Type C with heavy-
weight, load-bearing walls and lightweight floors is very common in older
buildings with masonry walls and traditional timber floors. These houses
313
314 Sound Insulation in Buildings
(a) (b)
(c) (d)
Figure 11.1 Direction of flanking transmission in four building types. (a) Type A – heavy-
weight walls and floors; (b) type B – light-weight walls and heavy-weight floors;
(c) type C – heavy-weight walls and light-weight floors; and (d) type D – light-
weight walls and floors.
11.1.2 Junction attenuation
One flanking transmission path involves two surfaces: one in the source
room and another in the receiving room. The pair of plates is connected
through a junction (Figure 11.3). The junction attenuation is defined as the
Figure 11.2 Airborne sound transmission horizontally. The transmission paths are shown
via one flanking surface (the floor) and directly through the wall. (Reproduced
from Kristensen, J. and Rindel, J. H., Bygningsakustik – Teori og praksis (Building
acoustics – Theory and practice, in Danish). SBI-Anvisning 166, Danish Building
Research Institute, Hørsholm, Denmark, 1989. With permission.)
Flanking transmission 315
P1,i P2,ij
Dv,ij vj
vi
Si σi Sj σj
P2,i
Figure 11.3 Simplified outline for flanking transmission from plate i to plate j at airborne
sound exciting. Dv,ij is the junction attenuation.
difference in velocity level on the plate i in the source room and that on the
plate j radiating into the receiving room:
v i2 1
Dv ,ij = 10 lg = 10 lg (11.1)
v 2j dij
P1,0
Rij = 10 lg (11.2)
P2,ij
where P1,0 is the incident sound power on the partition wall area S 0 and P2,ij
is the power radiated into the receiving room from the transmission path
under consideration. The reason for choosing the same reference area S 0
for all flanking transmission losses is that they can easily be added together
without further area corrections. This would not be the case if each flank-
ing transmission loss had the incident sound power P1,i on the flanking
surface Si as a reference.
316 Sound Insulation in Buildings
Under the assumption of diffuse sound field in the source room, the
flanking transmission loss can be written as (see Equation 8.25):
p12 S0
Rij = 10 lg (11.3)
4 ρ cP2,ij
where p1 is the mean square sound pressure in the source room and S 0 is
2
the area of the common surface between the rooms. In cases where the rooms
have no common separation area, S0 is set to a fixed reference value of 10 m2 .
With a diffuse sound field in the source room, the mean velocity of forced
vibrations in the excited plate i can be determined from Equations 8.23 and
5.35:
2 p12
v i2,for = (11.4)
(ω mi )2 (1 − ( f fc , i )
2 2
)
+ ηi2
The total vibrations in plate i are combined from forced and resonant excita-
tion. It can be assumed that both are transferred to the receiving plate j with the
same junction attenuation and, thus, the resulting velocity in plate j becomes:
(
v 2j = dij ⋅ v i2,for + v i2,res ) (11.6)
The forced vibrations are restricted to the excited plate i. In plate j, only
resonant vibrations are possible although the energy of the vibrations is fed
from forced and resonant vibrations in plate i. By means of Equations 11.4
through 11.6, the radiated sound power from plate j is:
ρ c 2 c 2
σ ∆N
i
= dij ⋅ p12 ⋅ ⋅ S jσ j ,res ⋅ + i ,res
⋅ (11.7)
( )
2
(ω mi )2
1 − ( f fc ,i ) + ηi2 2ηi Si f ∆f
2
S jσ j ,res 2 c 2σ i ,res ∆Ni
Rij = R0,i + Dv ,ij − 10lg − 10lg + ⋅ (11.8)
( )
2
S0 1 − ( f fc ,i ) + ηi2 2ηi Si f ∆f
2
where R0,i = 20lg (ω mi 2 ρc ) is the mass law at normal incidence for plate
i. From Equation 11.8, it is seen that a high flanking transmission loss can
be achieved when the plate being excited has a high transmission loss, i.e.
with heavy flanking constructions. Large junction attenuation or small
radiation efficiency for the radiating surface j may also contribute to a
high flanking transmission loss.
The transmission loss R of the direct transmission path can be written in
a form similar to Equation 11.8, using index 1 for this surface:
2σ 1,for c 2σ 1,res
2
∆N1
R = R0,1 − 10 lg + ⋅ (11.9)
( )
1 − ( f fc ,1 )2 + η12 2η1S1f ∆f
2
For this direct transmission, the radiation into the receiving room is dif-
ferent for the forced and resonant vibrations, represented by the different
radiation efficiencies σ 1,for and σ1,res. Here, the area is called S1, but it is
normally the same as the reference area S0.
R′ = 10 lg
P1,0
P2,0 + Σ P2,ij
(
= −10 lg 10− R / 10 + Σ 10− Rij / 10
) (11.10)
S0σ i ,res
Rij = Ri + Dv ,ij + 10 lg (11.11)
S jσ j ,res
Rij = Rji = 1
2 ( Rij + Rji ) (11.12)
Thus, the radiation efficiencies will disappear from the equation, and
S0
Rij = 12 Ri + 12 Rj + Dv ,ij + 10 lg (11.13)
Si S j
where
Dv ,ij = 1
2 ( Dij + Dji ) (11.14)
Figure 11.4 shows examples of transmission paths for impact sound pres-
sure horizontally and vertically. Horizontally, there will normally only be
two significant transmission paths, whereas vertically, there will be one
transmission path for each of the four flanking walls, in addition to the
direct transmission through the horizontal division.
p 2,2ij A2 4 ρcP2,ij
Ln,ij = 10 lg = 10 lg (11.15)
2
p A0
0 p02 A0
Flanking transmission 319
(a) (b)
Figure 11.4 Impact sound transmission paths. (a) Horizontal and (b) vertical. (Reproduced
from Kristensen, J. and Rindel, J. H., Bygningsakustik—Teori og praksis (Building
acoustics—Theory and practice, in Danish). SBI-Anvisning 166, Danish Building
Research Institute, Hørsholm, Denmark, 1989. With permission.)
F 2 Re {Yi } F 2 ∆Ni
v i2 = = ⋅ (11.16)
2π fmi Siηi 8π fmi Si ηi ∆f
2 2
where mi, Si and ηi are the mass per unit area, area and loss factor of plate
i, respectively. Yi is the mechanical point admittance of the plate, i.e. the
inverse of the point impedance Yi = 1/Zi. The modal density in plate i is ΔNi /
Δf, and this is a function of the frequency (different for bending waves and
shear waves).
The transmission of the vibration energy from plate i to plate j is described
with the junction attenuation Dv,ij (Equation 11.1). The radiated sound power
P2,ij is a function of the average velocity in plate j and the radiation efficiency:
F P2,ij
Dv,ij
vi vj
Zi Si mi ηi Sj σj
Figure 11.5 Simplified outline for flanking transmission from plate i to plate j by e xciting
with a tapping machine. Dv,ij is the junction attenuation and Zi is the mechani-
cal point impedance.
320 Sound Insulation in Buildings
v 2j F 2 ρc ∆Ni (11.17)
= ⋅ ⋅ ⋅ S jσ j ,res
v i2 8π fmi Si ηi ∆f
2 2
The impact sound pressure level for transmission path i → j can now be
determined from:
4 ρcP2,ij ρ 2c 2 v 2j F 2 ∆Ni
Ln,ij = 10 lg = ⋅ ⋅ σ
10 lg ⋅ S j j ,res (11.18)
p02 A0 p02 A0 v i2 2π fmi2 Si2ηi ∆f
c2 ∆Ni
Ln,ij = 82 − 20 lg mi − Dv ,ij + 10 lg 2 ⋅ ⋅ S jσ j ,res − ∆Lcover (11.19)
π
i i
S η ∆f
η S jσ j ,res
Ln,ij = 82 − 10 lg mi2 i − Dv ,ij + 10 lg − ∆Lcover (11.20)
fc ,i Si
2fS0
Rij + Ln,ij = 82 + R0,i − 20 lg mi + 10 lg − ∆Lcover (11.22)
π Siσ i ,res
As mi is also inherent in the mass law R0,i it cancels out and we get:
Siσ i ,res
Rij + Ln,ij = 38 + 30 lg f − 10 lg − ∆Lcover (11.23)
S0
This shows that the simple relation between the airborne sound insulation
and the impact sound pressure level for the total transmission (Equation
10.21) also applies to the individual transmission paths. The only differ-
ence is that the area correction, which is due to the flanking transmission
loss always refers to the area of the partition surface S 0. In case of vertical
transmission, S 0 = Si is the area of the floor in the source room.
11.3 JUNCTION ATTENUATION
11.3.1 Empirical model
One of the major problems when calculating sound insulation in buildings
is to provide data for the junction attenuation between the building parts.
For homogeneous massive plates, there is an extensive experimental basis,
as the data for cross junctions shown in Figure 11.6.
The junction attenuation is approximately 12 dB when the plates
have the same mass per unit area. Flanking transmission through a
structure intersected by a heavier structure is attenuated more than
12 dB and vice versa.
Pij (11.24)
γ ij =
Pinc
In Figures 11.7 and 11.8, the theoretical transmission coefficients are ren-
dered for different types of junctions. These are X-junctions or junctions
between three or two plates. Applying the numbering of Figure 11.9, the
corner joint only consists of plate 1 and plate 2, whereas the T-junction
misses either plate 3 or plate 4. The abscissa in Figures 11.7 and 11.8 is the
ratio:
m2 fc1 (11.25)
ψ =
m1fc 2
where m is the mass per unit area and fc is the critical frequency. The
index refers to the plates with numbers 1 and 2, respectively. In X- and
T-junctions, it is presumed that plate 3 equals plate 1, and plate 4 equals
plate 2.
60
st.d
d
50 a b
Junction attenuation Dv,ij, dB
c
40 transmission a b
transmission a c
30
20
10
0
0.1 0.2 0.4 0.6 0.8 1 2 4 6 8 10 20 40 60 80
Ratio of mass per unit area mcd / mab
0.01
0.02
Structural transmission loss, 10 lg (1/τij), dB
37 37 34 40
0.05
0.1
0.2
27 27 24 30
0.5
1
2
5
17 17 14 20 10
20
50
100
7 7 4 10 fc2
0 fc1
0 0
0
0.01 0.02 0.05 0.1 0.2 0.5 1 2 5 10 20 50 100
0.01 0.02 0.05 0.1 0.2 0.5 1 2 5 10 20 50 100
0.005 0.01 0.025 0.05 0.1 0.25 0.5 1 2.5 5 10 25 50
0.02 0.04 0.1 0.2 0.4 1 2 4 10 20 40 100 200
m f
ψ = 2 c1
m1 fc2
20 20
fc2
10 10 fc1
0 0
0.01 0.02 0.05 0.1 0.2 0.5 1 2 5 10 20 50 100
0.02 0.04 0.1 0.2 0.4 1 2 4 10 20 40 100 200
m f
ψ = 2 c1
m1 fc2
4 m4 = m2
m1 m3 = m1 m1 m3 = m1
1 3 1 3
2 m2 2 m2
2
f 1 1 2f
cB,eff = c s − + 1+ s (11.26)
fs 2 2 f
However, this is the so-called phase speed. In media with dispersion, the
group speed cg differs from the phase speed (Pedersen, 1995) and we get:
2
cB3 ,eff 2f
cg = 1+ s (11.27)
c s2 f
or approximately:
1 1
Pinc = ij mi v i2 c g ,i ≅ ij mi v i2 2cB,i (11.29)
π π
where cg,i is the group speed of the vibrations in plate i, and the approxima-
tion is valid for bending waves.
In the receiving plate j, the power transmitted from plate i must equal the
absorbed power under steady-state conditions Pij = Pj,abs and the steady-state
energy in the plate is Ej = mj Sj v 2j . Thus, we have the energy balance for
plate j:
where Sj is the area and ηj is the total loss factor for plate j. The structural
transmission coefficient yields:
Pij m j v 2j S jη j m j v 2j 2.2S j
γ ij = = 2π f ⋅
2
2
= 2π ⋅
2
2
⋅ (11.31)
Pinc mi v i ij c g ,i mi v i ij c g ,iTj
where cg,i is the group speed of transversal waves in plate i and ℓij is the
length of the edge that adjoins the two plates. In the last expression, the
structural reverberation time Tj in plate j is inserted from Equation 2.30.
This formula can be used as the basis for measurements of the structural
transmission coefficient.
The structural transmission coefficient can be expressed in dB and is
called the structural transmission loss R s,ij:
1 m lij
Rs,ij = 10 lg = Dv ,ij + 10 lg i + 10 lg (dB) (11.32)
γ ij mj aj
S jη j Sj
a j = 2π 2 f ⋅ ≅ 4.4 ⋅ π 2 (11.33)
c g ,i c g ,i ⋅ Tj
Insertion of the group speed (Equation 11.28) in Equation 11.33 yields the
structural transmission loss:
326 Sound Insulation in Buildings
Dv ,ij + 10 lg mi + 10 lg lij Tj f fc,i + 12 dB ( f < fs,i 4)
mj Sj
Rs,ij ≅ (11.34)
mi c s,i lij
Dv ,ij + 10 lg
mj
+ 10 lg Tj + 9 dB ( f > fs,i 4)
cS j
1 1 1 lij lij
Kij = (Rs,ij + Rs, ji ) = (Dv ,ij + Dv , ji ) + ⋅ 10 lg ⋅
2 2 2 ai a j
lij2
= Dv ,ij + 5lg (dB) (11.35)
ai ⋅ a j
Here, ai and aj are the equivalent absorption lengths of plates i and j, respec-
tively, and lij is the length of the connection.
As examples, we consider a rigid junction of homogeneous plates,
either as a cross-junction or a T-junction. See Figure 11.9 for definition
of K13 and K12 . Examples of the vibration reduction index suggested in
ISO 12354-1 are shown in Figures 11.10 and 11.11. These are assumed
frequency-independent.
50
40
Vibration reduction index, Kij, dB
30
20
10
–10
0.1 0.2 0.5 1 2 5 10 20 50
Mass ratio m2 / m1
K13 K12
Figure 11.10 Vibration reduction index for a rigid cross-junction. (Adapted from ISO
12354-1, Building acoustics—Estimation of acoustic performance of buildings
from the performance of elements—Part 1: Airborne sound insulation between
rooms, International Organization for Standardization, Geneva, Switzerland,
2017. Figure E1.)
50
40
Vibration reduction index, Kij, dB
30
20
10
–10
0.1 0.2 0.5 1 2 5 10 20 50
Mass ratio m2 / m1
K13 K12
Figure 11.11 Vibration reduction index for a rigid T-junction. (Adapted from ISO 12354-
1, Building acoustics—Estimation of acoustic performance of buildings from the
performance of elements—Part 1: Airborne sound insulation between rooms,
International Organization for Standardization, Geneva, Switzerland, 2017.
Figure E3.)
328 Sound Insulation in Buildings
With two elastic interlayers, e.g. one below and the other one above the
slab, the transmission loss becomes strongly frequency dependent, increas-
ing with frequency above a certain frequency, which according to Pedersen
(1995) can be estimated from:
−3/ 2
ρ1ρ 2 h1
−6
f1 = 2.5 ⋅ 10 d (Hz) (11.36)
G w
where
h1
ρ1
G d
ρ2 w
ρ1
h1
Figure 11.12 Junction with two identical thin elastic interlayers. (Adapted from Pedersen,
D. B., Appl. Acoust., 46, 285–305, 1995. With permission from Elsevier.)
where R s,ij,rigid is the structural transmission loss for the equivalent rigid
junction, n is the number of resilient layers in the junction (normally 1 or 2)
(Pedersen, 1995) and ΔR s is:
0 dB (f < f1)
∆Rs ≅ f (11.38)
10 lg (dB) (f > f1)
f1
With two elastic interlayers, the transmission loss increases by 6 dB per
octave above the frequency f1. In the case of only one interlayer, the improve-
ment of the transmission loss will be only about half the value for two inter-
layers. An improvement by 9 dB is obtained with either of these measures:
11.4 EXAMPLE
If the measured sound insulation is less than expected, it can be very dif-
ficult to point at the reason. However, the measures needed for an improve-
ment depend very much on which transmission path is most important. An
330 Sound Insulation in Buildings
80
70
Structural transmission loss Rs ,ij, dB
60
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency, Hz
Measured Corrected for Mindlin waves Simple estimate
Figure 11.13 Frequency dependency of the structural transmission loss with two elastic
interlayers. Measurement results without and with correction for group
speed of Mindlin waves (Equation 11.28). (Adapted from Pedersen, D. B.,
Appl. Acoust., 46, 285–305, 1995, Figure 4. With permission from Elsevier.)
75 mm mineral wool
80
f0
70
60
50
R' and L'n, dB
40
30
20
fc
10
0
50 100 200 500 1000 2000 5000
Frequency, Hz
R΄ before R΄ after L΄n
paths that only affect the airborne sound transmission and not the impact
sound transmission. In the current case, we see a clear deviation of the
sum curve at frequencies above 200 Hz and especially around 315 Hz. This
332 Sound Insulation in Buildings
140
f0
130
120
Sum R' + L'n, dB
110
100
90
fc
80
50 100 200 500 1000 2000 5000
Frequency, Hz
Figure 11.16 Sum of measured R′ + L′ n before and after mounting of gypsum board pan-
els on the internal walls. Theoretical curves from Figure 10.6b are shown
for comparison.
REFERENCES
R. Breuwer and J.C. Tucker. 1976. Resilient mounting systems in buildings. Applied
Acoustics 9, 77–101.
R.J.M. Craik. 1981. Damping of building structures. Applied Acoustics 14, 347–359.
E. Gerretsen. 1979. Calculation of the sound transmission between dwellings by
partitions and flanking structures. Applied Acoustics 12, 413–433.
Flanking transmission 333
Measurement methods
12.1.1 Microphone
The sound pressure level in a room is measured with a calibrated sound
level meter or a calibrated measurement microphone connected to an anal-
yser. The microphone must be selected to have a flat frequency response
in the frequency range of the measurement, and the electrical noise floor
must be sufficiently low to allow correct measurements. A microphone
with a big diaphragm, e.g. a 1 inch microphone. has a high sensitivity and
335
336 Sound Insulation in Buildings
an upper frequency limit around 8 kHz, which is sufficient for most cases
in building acoustics. However, a smaller ½ inch microphone is often pre-
ferred although the noise floor is higher, because it is less expensive and
easier to handle.
The frequency response should be flat for diffuse field incidence when
measuring in a room, i.e. a pressure microphone should be used, not a free
field microphone.
1 Tm
L = 10 lg
Tm ⋅ p02 ∫
0
p2 dt dB (12.1)
where
p is the sound pressure as a function of time (Pa).
p 0 = 20 μPa is the reference sound pressure.
Tm is the integration time (s).
( )
L = 10 lg 10Lsb /10 − 10Lb /10 dB (12.2)
where
When the difference is 4 dB, the correction is 2.2 dB. If the difference is less
than 4 dB, the correction 2.2 dB is applied and it must be clearly marked
that the result is influenced by background noise. For sound insulation
measurements, the background noise level should be at least 6 dB below the
measured sound pressure level (see Section 12.2.1).
338 Sound Insulation in Buildings
Table 12.2 R
elation between the Schroeder limiting frequency and the room
volume when the reverberation time is either 0.5 s or 0.8 s
V (m3)
fg (Hz) T = 0.5 s T = 0.8 s λ/2 (m)
100 200 320 1.72
125 125 200 1.38
160 80 125 1.08
200 50 80 0.86
250 32 50 0.69
315 20 32 0.55
400 13 20 0.43
500 8 13 0.34
Note: The minimum distance between uncorrelated microphone positions in a diffuse sound
level is well defined, whereas the centre region should be avoided because
extremely low sound pressure levels occur along the nodal lines. Since statisti-
cal methods do not apply at these low frequencies, it is impossible to avoid
correlation between the positions, and in contrast to the high-frequency range,
increasing the number of positions does not affect the measurement uncertainty.
Measuring the sound pressure level from building service equipment in
accordance with ISO 16032, three microphone positions are used; one in
a corner and two distributed in the reverberant field. The corner position
should be about 0.5 m from the nearest surfaces. Initial measurements are
made in all corners using the C-weighting filter, and the corner with the
highest C-weighted sound pressure level is chosen as position 1. Positions
2 and 3 shall be at least 1.5 m from position 1 and from any sound source.
The distance from room surfaces shall be at least 0.75 m. The result of mea-
sured sound pressure levels in the three microphone positions is calculated
as the energetic average using Equation 1.34.
T
LnT = L − 10 lg
T0
(dB ) (12.3)
A0 c ⋅ A0 ⋅ T
Ln = L − 10 lg
A
= L − 10 lg
55.3 ⋅ V
(dB ) (12.4)
where
12.2.1 Generation of sound
A loudspeaker is used to generate the sound, which must be at a sufficiently
high level in order to get a sound pressure level in the receiving room with-
out the influence of background noise. The directivity of the loudspeaker
should be close to omnidirectional, and the frequency response should
cover the relevant measurement range, which is at least from 100 Hz to
3150 Hz, but preferably from 50 Hz to 5000 Hz in one-third octave bands.
The traditional measurement technique uses stochastic noise (white or
pink), which can either be emitted as a broad-band signal or as a band-
limited signal with a bandwidth of at least one-third octave. The latter is, of
course, more time consuming, but has the advantage of allowing a higher
sound pressure level to be generated without overloading the loudspeaker.
Airborne sound insulation can also be measured using sound intensity, see
ISO 15186 (Part 1 to 3) and Machimbarrena and Jacobsen (1999).
More advanced measurement techniques use deterministic sound sig-
nals, either MLS (maximum-length-sequence) signal or swept-sine signals
in combination with either correlation signal processing or deconvolution
(see ISO 18233). The resulting impulse responses are squared and inte-
grated in order to get the stationary sound pressure level. Compared to the
traditional method, both methods have the advantage of suppressing the
influence of background noise. The swept-sine method has a number of
advantages over the MLS method:
The crest factor of a sound signal is defined as the ratio between the peak
value and the root-mean-square (RMS) value of the signal. A harmonic signal
that is composed of N sine functions has the crest factor √(2N), and a single
sine function has the crest factor √2 ≈ 1.414. White noise and pink noise have
a much higher crest factor, which means a higher risk that the loudspeaker
can be overloaded.
While distortion from the loudspeaker is not really a problem with the tra-
ditional measurement technique, it is fatal to the MLS technique. With the
swept-sine technique, some (but not all) of the harmonic distortion is separated
from the impulse response and does not influence the result. Both methods
require time-invariant systems, i.e. nothing must move in the rooms during the
measurements; even air movements may cause errors in the measurements and,
of course, moving microphones are out of question. While time invariance
is mandatory with MLS measurements, the swept-sine is much more robust
Measurement methods 341
Table 12.3 I mproved SNR as function of sweep duration using the swept-sine
technique
Duration of sweep (s) 60 336 762
Duration per one-third octave (s) 2.9 16 32
Improved SNR (dB) 14 21 24
to this. Some years ago, when the computer capacity was more restricted, the
MLS technique had the advantage of requiring less memory, but since this is no
longer an issue, the many advantages of the swept-sine technique means that
the MLS technique is obsolete for building acoustic applications.
With the swept-sine technique, the advantage of improving the signal-
to-noise ratio (SNR) is connected to the length of the sweep signal. Only a
single sweep is used in each measurement position, which is in contrast to
the MLS technique, which uses the averaging of several repeated sequences.
If we compare measurements with the averaging time Fast (0.125 s), the
improved SNR with different sweep durations are shown in Table 12.3.
As an example, the sound level difference between two rooms has been
measured with different techniques and the results are shown in Figure 12.1.
The background noise in one-third octaves was around 43 dB at 125 Hz
to 200 Hz, decreasing towards lower and higher frequencies, being around
30 dB at 50 Hz to 80 Hz and around 25 dB at 2500 Hz to 5000 Hz. The influ-
ence of the background noise is worst using the broadband excitation and
parallel analysis. Some improvement was obtained using band-limited noise
excitation because the loudspeaker could be set to a level about 10 dB higher.
The best result was obtained with the swept-sine; the excitation signal could
be increased another 3 dB, but most importantly, the influence of the back-
ground noise in the receiving room was efficiently suppressed. The differ-
ence between the methods is clearly seen at medium and high frequencies.
The use of swept sine for measuring airborne sound insulation has been
further studied by Satoh et al. (2011). Using very long sweeps (10 min per
octave) and a technique for subtraction of background noise, they could
measure sound pressure levels 30 dB below the background noise with less
than 1 dB error. Løvstad et al. (2014) describe an example of the practical
application of the swept-sine technique for a rather extreme case where the
sound insulation was 95 dB at 500 Hz.
80
70
60
Level difference, D (dB)
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Figure 12.1 Level difference between two rooms measured with different techniques.
Solid line: Swept-sine excitation. Dashed line: Serial excitation with noise
in one-third octave bands. Dotted line: Parallel excitation with pink noise.
(Adapted from Norsonic, Swept sine measurement technique using the Nor121
analyser. Application Note, Ed. 3, 10.06, Norsonic, Lierskogen, Norway,
2006. Figure 3.)
1 m
DnT = −10 lg
m ∑
j =1
10− DnT , j /10 ( dB ) (12.5)
m
R′ = −10 lg
1
∑ 10− Rj /10 (dB) (12.6)
′
m
j =1
where
m is the number of loudspeaker positions.
DnT,j is the standardized level difference for loudspeaker position j.
Rj′ is the apparent sound reduction index for loudspeaker position j.
The microphones shall be calibrated as for normal sound level measure-
ments before the measurements, although strictly speaking, it is sufficient
that the same microphone is used in the source room and in the receiving
room, since the measurement result is taken from the level difference. In
each room, the spatially averaged sound pressure level must be measured,
and this can be done in three different ways:
≥ 0.5 m
≥ 1.0 m ≥ 0.7 m
≥ 0.5 m ≥ 0.5 m
≥ 1.0 m
≥ 1.0 m
≥ 0.5 m
Figure 12.2 Sketch of a source position and microphone positions in the source room
with minimum distances indicated.
344 Sound Insulation in Buildings
r ≥ 0.7 m
θ ≥ 10°
≥ 1.0 m
≥ 0.5 m ≥ 0.5 m
220° rotation
m Microphone
.9
–0
m
0.3
2.0 m
0.5 m
Floor level
Figure 12.4 Manual scanning with the microphone mounted on a rod and following a cylin-
drical path. (Modified from ISO 16283-1, Acoustics – Field measurements of sound
insulation in buildings and of building elements – Part 1: Airborne sound insulation,
International Organization for Standardization, Geneva, Switzerland. Figure 1c.)
12.2.3 Low-frequency procedure
According to ISO 16283-1, a special low-frequency procedure shall be used
for the 50 Hz, 63 Hz and 80 Hz one-third octave bands when the room
volume is smaller than 25 m3. This applies to both the source room and
the receiving room. However, it would make sense to apply this method
more generally to the frequencies that are below the Schroeder limiting fre-
quency. The principle is that measurements made in the central zone of the
room (the default procedure) are combined with the highest sound pressure
level that occurs in a room corner. The microphone position in the corner
is at a distance of 0.3 m to 0.4 m from each room boundary (Figure 12.5).
dz
dy
dx
Figure 12.5 Position of a microphone near a corner. The distances dx, dy and dz should
all be within 0.3 m to 0.4 m.
346 Sound Insulation in Buildings
1 Lcorner
2
L
LLF = 10 lg ⋅ 10 10 + ⋅ 1010 (dB) (12.7)
3 3
LLF is determined for each loudspeaker position and the resulting DnT or R′
values for 50 Hz, 63 Hz and 80 Hz are averaged as for the default procedure.
However, the reverberation time is measured in the octave band 63 Hz, and
the result is applied in all three one-third octave bands 50 Hz, 63 Hz and
80 Hz. The reason for this pragmatic method is that the modal density may
be very low below 100 Hz, especially in small volumes and, thus, measure-
ments in one-third octave bands may be unreliable or even impossible.
T
D2m,nT = L1, 2m − L2 + 10 lg (dB) (12.8)
T0
Measurement methods 347
where
L1,2m is the sound pressure level 2.0 m in front of the façade.
L 2 is the spatially averaged sound pressure level in the receiving room.
T is the reverberation time in the receiving room (s).
T0 = 0.5 s is the reference reverberation time.
Equation 12.8 is used both for traffic noise and for loudspeaker measure-
ments. Using a loudspeaker, the position of the loudspeaker is in a distance
of r ≥ 7 m from the centre of the test specimen or at least D = 5 m in front of
the façade (Figure 12.6). The angle of incidence shall be 45° ± 5° and the
loudspeaker can be placed either on the ground (preferred) or as high over
the ground as possible.
S
′ ° = L1,s − L2 + 10lg
R45 − 1.5dB (12.9)
A
where
L1,s is the sound pressure level on the test specimen.
L 2 is the spatially averaged sound pressure level in the receiving room.
θ = 45° ± 5°
Figure 12.6 Position of the loudspeaker in front of the façade. For the global method,
r ≥ 7 m and D ≥ 5 m. For the element method, r ≥ 5 m and D ≥ 3.5 m.
348 Sound Insulation in Buildings
(a) ≤ 10 mm (b) ≤ 3 mm
Figure 12.7 Position of the microphone on the test surface for the measurement
of L1,s. (a) Microphone with diaphragm perpendicular to the surface. (b)
Microphone with diaphragm parallel with the surface.
If, for practical reasons, the element loudspeaker method cannot be used,
it is possible to use road traffic as a noise source. Then, it is assumed that
the angles of incidence cover a wide range and the formula for the apparent
sound reduction index is:
S
′ s = L1, s − L2 + 10 lg
Rtr, − 3 dB (12.10)
A
beams or ribs. If the floor under test is larger than 20 m 2 , the number of
tapping machine positions may be increased, and if the area of the floor
is larger than 50 m 2 , the number of tapping machine positions should be
eight.
The measurement of spatially averaged sound pressure level and the
position of microphones in the receiving room follow the same rules as
for airborne sound insulation (see Section 12.2). If fixed microphone posi-
tions are used, the number of positions shall equal the number of tapping
machine positions (or a multiple of the number of tapping machine posi-
tions). Measurements shall be made in at least two different microphone
positions for each tapping machine position, if the number of tapping
machine positions is less than six. If six or more tapping machine positions
are used, it is sufficient to measure in one microphone position for each
tapping machine position.
Measurements of impact sound insulation should include the low fre-
quencies down to the 50 Hz one-third octave band. For frequencies below
100 Hz, the low-frequency procedure using corner positions apply (see
Section 12.2). Averaging times for the measurements and correction for
background noise are also the same as for airborne sound insulation.
For each tapping machine position, the result is calculated either as the
standardized impact sound pressure level (Equation 12.3) or as the normal-
ized impact sound pressure level (Equation 12.4). Finally, the averaging
over all tapping machine positions is made as the energetic average using
one of these formulas:
1 m
L′ nT = 10 lg
m ∑10
j =1
Ln′ T , j /10
(dB) (12.11)
1 m
∑
′
where
m is the number of tapping machine positions.
Ln′ T , j
is the standardized impact sound pressure level for tapping
machine position j.
Ln,′ j is the normalized impact sound pressure level for tapping machine
position j.
The microphone shall be calibrated as for normal sound level measure-
ments before the measurements. This is particularly important for the mea-
surement of impact sound pressure level, as it is the actual sound pressure
level that is measured, not the difference between two levels as for airborne
sound insulation.
350 Sound Insulation in Buildings
12.5 REVERBERATION TIME
12.5.1 Measurement positions
For an engineering grade measurement, the number of source-micro-
phone combinations should be at least six, e.g. two source positions
combined with three microphone positions. However, when the result is
used only as a correction term to other engineering-level measurements,
only one source position and three microphone positions are required.
Also, under these conditions, a rotating microphone boom may be used
instead of multiple microphone positions, but only for the interrupted
noise method.
ISO 3382-2 recommends that in small domestic rooms, one source posi-
tion should be in a corner. Microphone positions should be at least half a
wavelength apart and a quarter wavelength from the boundaries. If a rotat-
ing microphone boom is used, the radius of the boom shall be at least 0.7 m.
where f is the centre frequency of the filter and Tdet is the reverberation
time of the detector, assuming exponential averaging of the decaying sound
pressure. This implies a lower limit for measurable reverberation times,
which is 0.7 s at 100 Hz and 1.4 s at 50 Hz (Table 12.4). However, it was
suggested by Jacobsen and Rindel (1987) to overcome these problems by
using a time-reversal technique, either by recording the decay in the room
and playing it backwards through the filters or by constructing the filters
with a time-reversed impulse response meaning a slow built-up and fast
decay. With this technique, the minimum requirements for the reverbera-
tion time are changed to:
Thus, the lower limit at 100 Hz one-third octave band is reduced to 0.18 s
and 0.35 s at 50 Hz.
Some instrument manufacturers have overcome the problem by the use
of reverse filtering, i.e. the one-third octave band filters have been specially
crafted to provide short reverberation times and still comply with require-
ments for the frequency response. Thus, the lower limit for measurement
of reverberation time in one-third octave bands can be around 0.1 s in the
entire frequency range from 50 Hz to 5000 Hz.
Another problem in small volumes is that the modal density is very low. In
the one-third octave bands below 100 Hz, there may be only a few modes,
or none. With two or three modes within a band, the decay curve may suf-
fer from strong wavy behaviour due to interference, or double slope due to
very different attenuation of the modes. In either case, the evaluation of a
reverberation time is difficult and associated with a huge uncertainty. A
pragmatic solution to this problem was suggested by Hopkins and Turner
(2005), namely, to measure the reverberation time in the 63 Hz octave band
and apply this result for the 50 Hz, 63 Hz and 80 Hz one-third octave bands.
This idea has been adapted in the ISO 16283 series for sound insulation
measurements.
Finally, it is a problem with commonly used sound sources for building
acoustic measurements that the sound power in the 63 Hz octave is not
sufficient for measuring the reverberation time. The signal-to-noise level
should be at least 35 dB and the background noise is often quite high at
low frequencies. The best solution is probably to use the swept-sine tech-
nique in combination with a loudspeaker that has good performance at low
frequencies.
Table 12.4 M inimum measurable reverberation time (in s) using normal one-
third octave and octave band filters and either forward or time
reversed analysis
Forward Backward
f, Hz One-third octave Octave One-third octave Octave
50 1.40 0.35
63 1.11 0.36 0.28 0.09
80 0.88 0.22
100 0.70 0.18
125 0.56 0.18 0.14 0.04
160 0.44 0.11
200 0.35 0.09
Measurement methods 353
12.5.5 Spatial averaging
The spatial averaging of reverberation time can be made in two different
ways. The reverberation times are evaluated in each pair of source-receiver
combinations, and the arithmetic average of reverberation times is calcu-
lated. Alternatively, the decay curves can be averaged, synchronizing the
beginnings and averaging the sound pressure squared in each time slot dur-
ing the decay. The sound power of the source shall be kept the same for
all measurements. The reverberation time is calculated from the ensemble-
averaged decay curve. This is the preferred method of spatial averaging in
ISO 3382-2.
12.6 VIBRATIONS
12.6.1 Calibration
Before any vibration measurement, it is important to calibrate the equip-
ment. Some calibrators give a specified RMS acceleration while other cali-
brators give a specified peak acceleration. An example is the B&K type
354 Sound Insulation in Buildings
a 9.81
La = 20 lg = 20 lg −6
≅ 136.8 dB re 1 µm/s2 (12.17)
a0 2 ⋅ 10
a a
v= = (12.18)
ω 2π f
Using the reference velocity v 0 = 10 −9 m/s and f = 80 Hz, the corresponding
velocity level of the calibration signal yields:
v a
Lv = 20 lg = La + 20 lg 0
v0 ω ⋅ v0
= La + 60 − 20 lgω ≅ 142.8 dB re 1 nm/s (12.19)
12.6.2 Measurement positions
Vibration measurements can be made both in vertical and horizontal direc-
tions. In buildings, the highest vibration levels are usually the vertical vibra-
tions on the floor. The vibration measurements should be made in positions
where the highest vibration level is obtained, which is in the middle of the
floor, or between the primary beams supporting the floor (Figure 12.8).
Additional measurement positions should be chosen more than 1 m from
the walls and from other positions.
The mounting of the accelerometer can be with bees’ wax on hard sur-
faces or using a device, as shown in Figure 12.9, in the case of carpets or
similar soft floor coverings.
12.6.3 Frequency weighting
Vibration measurements can either be performed in one-third octave
bands in the frequency range from 1 Hz to 80 Hz or using the standard-
ized frequency weighting curve (Figure 2.18). It is possible to calculate the
weighted acceleration level or the weighted velocity level from the one-third
Measurement methods 355
Primary
beam
Secondary
beams or
joists
Structural wall
or primary beams
60°)
=3
°( °
120
0
12
3×
130
0
15
15°
Figure 12.9
E xample of mounting device for measurement of vertical vibration on
carpet-covered floors. The weight of the device plus attached transducer
should be about 2.5 kg. Dimensions are in mm. (Reproduced from Jakobsen,
J., Measurement of vibration and shock in buildings for the evaluation of annoy-
ance. Nordtest project 618–86. Technical Report No. 139, Danish Acoustical
Institute, Denmark, 1987. With permission.)
356 Sound Insulation in Buildings
octave band spectrum using an equation similar to Equation 1.16 for the
A-weighted sound pressure level.
12.6.4 Integration time
The instantaneous maximum value of the acceleration level can be mea-
sured with an exponential averaging time constant of 1 s (“slow”) or with
a linear averaging over 2 s.
12.6.5 Noise floor
It is always recommended to ensure that the measured vibration values are
not influenced from the background noise of the measurement system. If
the accelerometer is placed on a heavy non-vibrating object, the apparent
vibration level should be at least 10 dB below the level of the real measure-
ments. If possible, baseline measurements should be obtained to verify the
background noise level.
12.6.6 Averaging of results
Whether it is relevant to average the measured results depends on the situa-
tion and the purpose of the measurements. In some cases, the spatial aver-
age over the measurement positions is calculated as an energetic average
(Equation 1.34). In other cases, the result must be reported for each mea-
surement position.
The Norwegian standard NS 8176 deals with vibrations in buildings due
to land, based transport, and the statistical maximum velocity must be cal-
culated in each position from at least 15 events of passing vehicles or trains.
According to this standard, the average maximum velocity is calculated
as the arithmetic average of the frequency-weighted maximum velocities
(not the squared velocities). Finally, the statistical maximum value (95 per-
centile) is calculated as the arithmetic average plus 1.8 times the standard
deviation of the frequency-weighted maximum velocities.
VDV = 4
∫a
0
4
w dt (m/s1.75) (12.20)
Measurement methods 357
12.7 LOSS FACTOR
2.2
ηtotal = (12.21)
f Ts
Table 12.5 Maximum measurable loss factors using normal one-third octave and
octave band filters and either forward or time-reversed analysis
Forward Backward
One-third octave Octave One-third octave Octave
0.03 0.10 0.13 0.39
358 Sound Insulation in Buildings
Br
ηtotal = (12.22)
f0
12.8 RADIATION EFFICIENCY
∑
n
vi2
Lv = 10 lg i =1
2
(dB) (12.23)
nv 0
∑
m
pi2
Lp = 10 lg i =1
2
(dB) (12.24)
mp 0
S
10 lg σ = Lp − Lv − 10 lg + 28 (dB) (12.25)
A
Measurement methods 359
where S is the area of the radiating surface and A is the equivalent absorp-
tion area of the receiving room. This formula can be derived from Equations
6.24 and 4.29.
12.9 JUNCTION ATTENUATION
2.2 π2 S j
aj = (m) (12.27)
f
Ts, j c
fref
where
Sj is the area of element j (m 2).
Ts,j is the structural reverberation time of element j (s).
c = 344 m/s is the speed of sound in air.
f is the frequency (Hz).
fref = 1000 Hz is a reference frequency.
360 Sound Insulation in Buildings
REFERENCES
J. Jakobsen (1987). Measurement of vibration and shock in buildings for the evalu-
ation of annoyance. Nordtest project 618–86. Technical Report No. 139,
Danish Acoustical Institute, Denmark.
J. Jakobsen (2001). Danish guidelines on environmental low frequency noise, infra-
sound and vibration. Journal of Low Frequency Noise, Vibration and Active
Control 20 (3), 141–148.
A. Løvstad, J.H. Rindel, V. Støen, A.T. Windsor, A. Negård (2014). Measurement and
simulation of high sound insulation and identification of sound transmission
paths in complex building geometries. Proceedings of Forum Acusticum 2014,
September 2014, Krakow, Poland.
M. Machimbarrena and F. Jacobsen (1999). Is there a systematic disagreement
between intensity-based and pressure-based sound transmission loss measure-
ments? Building Acoustics 6(2), 101–111.
NT ACOU 033 (1980). Bulkheads: Sound radiation efficiency – Laboratory test,
Nordtest, Helsinki, Finland.
NT ACOU 090 (1994). Building structures, junctions: Transmission of vibrations –
Field measurements, Nordtest, Helsinki, Finland.
NT ACOU 102 (1999). Building elements – Façade elements and façades: Field mea-
surements of airborne sound insulation – Loudspeaker method using MLS
noise signals, Nordtest, Helsinki, Finland.
Norsonic (2006). Swept sine measurement technique using the Nor121 analyser.
Application Note, Ed. 3, 10.06, Norsonic, Lierskogen, Norway.
NS 8176 (2005). Vibration and shock. Measurement of vibrations in buildings from
landbased transport and guidance to evaluation of effects on human beings,
2nd edition, Norwegian Standard, Oslo, Norway.
T. Poulsen and F.R. Mortensen (2002). Laboratory evaluation of annoyance of low
frequency noise. Working Report No. 1, Danish Environmental Protection
Agency, Copenhagen, Denmark.
F. Satoh, M. Sano, Y. Hayashi, J. Hirano, S. Sakamoto (2011). Sound insulation mea-
surement using 10 minute swept-sine signal. Proceedings of InterNoise 2011,
4–7 September 2011, Osaka, Japan.
C. Simmons (1999). Measurement of sound pressure levels at low frequencies in
rooms. Comparison of available methods and standards with respect to micro-
phone positions. Acustica – Acta Acustica 85, 88–100.
J. Takala and M. Kylliäinen (2013). Room acoustics and background noise levels in
furnished Finnish dwellings. Proceedings of InterNoise 2013, 15–18 September
2013, Innsbruck, Austria.
M.L.S. Vercammen (1992). Low-frequency noise limits. Journal of Low Frequency
Noise and Vibration 11, 7–13.
T. Yang, P. Liu, J. Yin (2014). Research on assessment of fourth power vibration dose
value in environmental vibration caused by metro. Proceedings of ICSV 21,
13–17 July 2014, Beijing, China.
Chapter 13
Measurement uncertainty
The sound pressure level in a room varies from one point to another, even in
an ideal diffuse sound field. These variations are particularly large when the
excitation is a pure tone or narrow band noise like a one-third octave band
and, thus, some kind of spatial averaging is necessary in order to get a reason-
ably accurate measurement result. Measurements at low frequencies in small
rooms need special consideration. This chapter presents an overview of the
theoretical background and methods for estimation of uncertainty in sound
pressure level measurements and reverberation time measurements in a room.
Chapter 4, Introduction to Room Acoustics, is a basis for this chapter.
363
364 Sound Insulation in Buildings
13.1.3 Confidence levels
Assuming a Gaussian distribution, the measurement result is an estimate
of the true result, which will be within a range of <z> ± k(g)·σ(z), where σ(z)
is the standard deviation and k(g) is a coverage factor that depends on the
confidence level, g (Table 13.1). This is called a two-sided test and is applied
when a measurement result is reported without comparison with any target
value. For example, with 95 % probability, the true result is within a range
of ±1.96 times the standard deviation (the expanded uncertainty U).
When a measured result is compared with a requirement, the one-sided
test applies. If, for example, a 90 % confidence level is chosen, the coverage
factor for a one-sided test is k = 1.28. This means that the limit zmax is not
Measurement uncertainty 365
1.0
0.8
0.6
Probability
0.4
0.2
0.0
–3 –2 –1 0 1 2 3
Measurand, z
w(z) P(z) 1 – P(z)
Table 13.1 Confidence levels g and the associated coverage factors k for one- and
two-sided tests
Coverage factor, k 1.00 1.28 1.65 1.96 2.58 3.29
Confidence level for 68 80 90 95 99 99.9
two-sided test, g (%)
Confidence level for 84 90 95 97.5 99.5 99.95
one-sided test, g (%)
(p ) = (p )
2
2
σ 2 2 − p2
ε R2 = 2 2 (13.2)
p2 p2
where σ (p2) is the standard deviation of the squared sound pressure and
<p2> denotes the spatial average. The sound pressure is the RMS value of a
time averaging, see Section 13.3.3 about sufficient measurement time.
(a) (b)
125 Hz sine 125 Hz noise
90 90
Sound pressure level (dB)
80 80
70 70
60 60
50 50
40 40
0 100 200 300 400 0 100 200 300 400
Distance (cm) Distance (cm)
(c) 125 Hz sine (d) 125 Hz noise
50 55 60 65 70 75 80 50 55 60 65 70 75 80
T
f g = 2000 (Hz) (13.3)
V
where T is the reverberation time in (s) and V is the room volume in (m3).
If the excitation of a room is with a band-limited noise signal (Figure 13.2b)
and the bandwidth is B (Hz), the relative spatial variance (Lubman, 1968)
is approximately:
2 1
ε R2 = arctan(Z) − 2 ln(1 + Z 2 ) (13.4)
Z Z
where Z = B/Br is the ratio between the bandwidth B of the noise signal and
the bandwidth Br of a typical room mode. With the reverberation time T,
the mode bandwidth is Br ≈ 2.2/T (see Equation 4.19). If the bandwidth B
is not very small, a good approximation for the relative spatial variance is:
−1 −1
B BT
ε R2 ≅ 1 + = 1 + (13.5)
π Br 6.9
where T is the reverberation time. It is seen that the variance decreases with
increasing bandwidth of the noise signal and with increasing reverberation
time; thus, a wideband noise signal in a highly reverberant room yields very
small spatial variations of the squared sound pressure. The approxima-
tion (Equation 13.5) is valid within 2.5 % for BT > 20. Below that range,
Equation 13.4 is used.
In practice, most measurements are performed with a wide band noise
signal and a series of band-limiting filters after the microphone. However,
it does not matter where in the measurement chain the band-limiting filter
is located; the above equation is still a valid approximation. The most com-
mon filters are the octave band (B = 0.71 f) and the one-third octave band
(B = 0.23 f), where f is the centre frequency of the band.
In the extreme case of a pure tone (bandwidth B = 0 Hz), we have εR 2 = 1,
and this corresponds to a standard deviation on the sound pressure level
of 5.6 dB (Schroeder, 1969). Thus, the sound pressure level for a pure tone
excitation in a room will vary within a range of approximately 22 dB (with
95 % probability assuming Gaussian distribution). However, these results
refer to using the sound pressure level as the measurand, i.e. when the mean
value and the standard deviation are determined directly from the dB val-
ues. This introduces a bias error in the mean value compared to the correct
result calculated from the squared sound pressures. The bias error and the
368 Sound Insulation in Buildings
In Table 13.2 the bias error is 2.5 dB for a pure tone (B = 0) but decreases
with increasing bandwidth. For broadband noise excitation (BT > 200), the
bias error becomes negligible. The standard deviation of the sound pressure
level L from the mean value <L> is 5.6 dB for a pure tone in a diffuse sound
field, as first found by Doak (1959).
∆N
Beff = B ⋅ ⋅ Br ⋅ κ = ∆N (B) ⋅ Br ⋅ κ (13.7)
∆f
where ΔN(B) is the number of modes within the bandwidth B. The overlap
factor κ (≤1) is included in order to take account of the fact that the modes
have some degree of overlap. Even if the modal overlap index is M < 3, there
is still more than one mode within the bandwidth of the modes (M > 1) in a
certain frequency range extending at least one octave below fg.
So, the relative spatial variance is calculated with Z = B eff /Br = ΔN(B)·κ
in Equation 13.4. Using instead the approximation (Equation 13.5) yields:
Measurement uncertainty 369
−1
∆ N (B) ⋅ κ
ε R2 ≅ 1 + for f <f g (13.8)
π
The overlap factor κ has been estimated by comparison with empirical data
like those in Figure 13.3, and the result is κ = 0.37 = π/8.5. The suggested
approximation for the relative spatial variance below the limiting frequency is:
−1
∆ N (B)
ε R2 ≅ 1 + for f < f g (13.9)
8.5
where ΔN(B) is the number of room modes within the measurement band-
width B. This approximation was first suggested by (Rindel, 1981).
Figure 13.3 shows a comparison of measured spatial standard deviations
in a room compared with the estimations from Equation 13.9 below the
limiting frequency and Equation 13.5 above the limiting frequency. The
two theoretical curves cross at the limiting frequency.
In the common case of measurements in one-third octave bands, Equations
13.5 and 13.9 can be modified with insertion of B = 0.23 f and the modal
density (Equation 4.13). Using the approximation (Equation 4.14), the rela-
tive spatial standard deviation yields, below and above fg, respectively:
−1/ 2
V ⋅f3
εR ≅ 1 + for f < f g (13.10)
2.9 ⋅ c 3
−1/ 2
T ⋅f
εR ≅ 1 + for f ≥ f g (13.11)
30
These are displayed graphically in Figure 13.4. The relative spatial stan-
dard deviation is the higher one in the two Equations 13.10 and 13.11. At
frequencies below the Schroeder limiting frequency, the room volume is the
important parameter, whereas at frequencies above the limiting frequency,
the reverberation time is the important parameter. The approximation
above the limiting frequency (Equation 13.11) is good, and that under the
limiting frequency (Equation 13.10) is also quite good, because the error
by using the simple Equation 13.5, instead of Equation 13.4, to some
extent, compensates for the error due to the approximate modal density
(Equation 4.14) instead of the more correct one (Equation 4.13).
If the volume and reverberation time are known, it is possible to estimate
the relative spatial standard deviation for the curves in Figure 13.4. The
curve with the highest standard deviation is always the valid one, and the
cross-over is at the Schroeder limiting frequency. It should be noted that
reducing the reverberation time in a room makes things worse because the
standard deviation is increased at frequencies above the limiting frequency.
370 Sound Insulation in Buildings
Measured
Relative spatial standard deviation
0.8
Below fg
Above fg
0.6
0.4
0.2
0
100 200 500 1000 2000
One-third octave center frequency (Hz)
Figure 13.3 M
easured and estimated relative spatial standard deviation in one-third
octave bands for a 70 m3 room with reverberation time approximately 1.0 s.
The Schroeder limiting frequency is 240 Hz. (Measured data from Jacobsen,
F., The diffuse sound field. Statistical considerations concerning the rever-
berant field in the steady state. Report No. 27, The Acoustics Laboratory,
Technical University of Denmark, 1979.)
1.0
25 V(m3), below fg
100
Relative spatial standard deviation
0.8
400
1600
0.6
T(s), above fg
10 5 2 1 0.5
0.4
0.2
0.0
50 100 200 500 1000 2000 5000
One-third octave center frequency (Hz)
Figure 13.4 Curves for estimation of the relative spatial standard deviation on one-third
octave band measurements in a room. At low frequencies, the volume is the
determining parameter, at high frequencies it is the reverberation time. For a
given volume and reverberation time, the cross-over is at the Schroeder limiting
frequency.
Measurement uncertainty 371
• The modal density is very low; in some one-third octave bands, there
may be a single mode: in other bands, the sound pressure rides on the
skirts of the frequency response of one or two modes with natural
frequencies outside the band.
• The range of variation of the sound pressure level, i.e. the difference
between the maximum and minimum curves, is high in the region
above the first axial mode (1,0,0), but decreases towards zero when
the frequency drops towards the cavity mode (0,0,0).
50
40
30
SPL (dB)
20
10 Pos (0, 0, 0)
Pos (lx/4, ly/4, lz/4)
Pos (lx/2, ly/2, lz/2)
0
20 40 60 80 100 200
Frequency (Hz)
These observations lead to the result shown in Figure 13.6. The frequency
range is divided into three, and the corresponding limiting frequencies are
f 2,0,0 (the frequency of the second axial mode in the longest room dimen-
sion) and fg (the Schroeder limiting frequency).
The course of the standard deviation in region 1 has been determined on
the basis of a number of measured data from some cases: see examples in
Figures 13.7 and 13.8. The empirical equation is:
lx ⋅ f
εR ≅ for f < f2,0,0 (13.12)
c
1
Relative spatial standard deviation (εR)
Below f (2,0,0)
0.8 Between f (2,0,0) and fg
Above fg
0.6
0.4
0.2
f2,0,0 fg
0
10 20 50 100 200 500 1000 2000 5000
One-third octave centre frequency (Hz)
Figure 13.6 C
ourse of the relative spatial standard deviation, and division into three
frequency regions.
Measurement uncertainty 373
p2 + σ (p2 )
σ (Lp ) ≅ 10 lg = 10lg(1 + ε R )
p2
5
Standard deviation of SPL in 1/3 octave bands (dB)
Measured
4 Prediction
0
20 50 100 200 500 1000 2000 5000
One-third octave centre frequency (Hz)
5
Standard deviation of SPL in 1/3 octave bands (dB)
Measured
4 Predicted
0
20 50 100 200 500 1000 2000 5000
One-third octave centre frequency (Hz)
Figure 13.8 Measured and estimated spatial standard deviation in a 31 m3 bedroom (3.8
m × 3.3 m × 2.5 m, T = 0.2 s). (Measured data from Simmons, C., Noise Control
Eng. J., 60, 405–420, 2012.)
was 0.7 m. In this example, the volume is small and reverberation time is very
short. The Schroeder limiting frequency is 160 Hz and f2,0,0 = 90 Hz, so region
2 is quite narrow in this case. The peak is noticed around 90 Hz.
w (z) = ez− e
z
(13.14)
where
ln (10)
z= (L − L0 ) ≅ 4.34 (L − L0 ) (13.15)
10
Here, L is the sound pressure level in an arbitrary point and L 0 is the sound
pressure level corresponding to the spatial average of the squared sound
pressure, p2 .
The Poisson distribution is an asymmetric function (Figure 13.9). This
implies that the spatial averaged sound pressure level (L) is 2.5 dB lower than
L 0. The probability for a measured sound pressure level L to be less than L 0
is 63 %, whereas the probability to exceed L 0 is only about half of that, 37 %.
Other characteristic measures that can be extracted from the cumulative
probability function of the Poisson distribution are that the 50 % (median)
level is L 0 − 1.6 dB, the 68 % confidence interval is (−7.6 dB; 2.7 dB) and the
1
0.9
0.8
0.7
0.6
Probability
0.5
0.4
0.3
0.2
0.1
0
–20 –15 –10 –5 0 5 10
L–L0 (dB)
Figure 13.9 Probability functions for the sound pressure level in a diffuse sound field.
(Adapted from Pierce 1989, Figures 6.16.)
376 Sound Insulation in Buildings
x y z
p2 = p02 ⋅ cos2 π nx ⋅ cos2 π ny ⋅ cos2 π nz (13.16)
lx l y l z
It follows that in an oblique mode (all n ≠ 0), the spatial average covering
the entire volume is:
lx ly
x y
p 2 2
0
∫
0
2
lx ∫
= p ⋅ cos π nx d x ⋅ cos2 π ny d y
0
ly
lz
z 1
∫
⋅ cos2 π nz d z = ⋅ p02
0
lz 8
(13.17)
This implies that in each of the eight corners, the sound pressure level is
10 lg (8) = 9 dB higher than the sound pressure level L 0 of spatial averaged
squared sound pressure. In a similar way, it is found that in the tangential
modes (two n ≠ 0), the highest sound pressure level is 6 dB above the room
average and in the axial modes (one n ≠ 0), the highest sound pressure level
is 3 dB above the room average.
The probability density functions of the various types of room modes
have been derived by Waterhouse (1970). The cumulative probability
curves for the various types of room modes, oblique, tangential, axial and
the cavity mode, are shown in Figure 13.10.
The statistics of the oblique and tangential modes have some similari-
ties; about 40 % of all positions in the volume have sound pressure levels
below L 0 and about 60 % of all positions have a higher sound pressure
level. The median level is 1.5 dB to 1.7 dB higher than L 0. These are the
most important modes in the frequency range between f 2,0,0 and fg (region
2 in Figure 13.6). Although the sound field is not diffuse because of insuf-
ficient modal overlap, the sound field is still three-dimensional and several
modes will occur together within the one-third octave bands. Thus, it is
impossible to predict where the minima will occur.
In the extreme low frequency range – below, the first room mode
(1,0,0) – the sound in the room excites the first room mode as well as the
Measurement uncertainty 377
1
0.9 Oblique
0.8 Tangential
Axial
Cumulative probability
0.7
Cavity
0.6
0.5
0.4
0.3
0.2
0.1
0
–20 –15 –10 –5 0 5 10
L–L0 (dB)
Figure 13.10 Cumulative probability functions for the sound pressure level in various
types of room modes.
13.3.2 Zones to be avoided
Below the Schroeder limiting frequency, the sound field in a room is not
stochastic and it is possible to choose measurement positions that minimize
378 Sound Insulation in Buildings
the measurement uncertainty. Figure 13.11 shows the sound pressure level
along the length of a room; the L 0 level occurs in the distance lx /4 from the
walls, and in distances greater than lx /3, the sound pressure level drops more
than 3 dB below L 0. As indicated in Figure 13.11, the centre region should
6
Zone to be avoided
3
L–L0 (dB)
–3
–6
–9
–12
0 1/4 1/3 1/2 2/3 3/4 1
x / lx
Figure 13.11 S patial variation of sound pressure level in a room when the first axial mode
is excited. Measurements in the central zone more than 1/3 room dimen-
sion from the walls should be avoided.
Measurement uncertainty 379
(a)
6
Zone to be avoided
3
L –L0 (dB)
–3
0 0.25 0.5 0.75 1 1.25 1.5
x/λ
(b)
9
Zone to be avoided
6
3
L –L0 (dB)
–3
–6
–9
–12
0 0.25 0.5 0.75 1 1.25 1.5
r/λ
Figure 13.12 Spatial variation of sound pressure level in a near room boundaries in a dif-
fuse sound field excited by a pure tone. (a) Near a wall a zone < 0.2 λ should
be avoided. (b) Near a corner a zone < 0.7 λ should be avoided.
380 Sound Insulation in Buildings
near a wall and a corner are shown in Figure 13.12. The size of the interfer-
ence zones that should be avoided is 0.2 λ from the wall and 0.7 λ from the
corner. Measurements are preferably made in the central zone of the room,
but in practice it is often impossible to avoid the interference zones unless
the room is big.
2
ε t2 = (13.18)
Bt
where t is the measurement time (or the integration time). The relationship
is also shown in Figure 13.13 in terms of the expanded uncertainty in dB.
The conversion to dB is used (Equation 13.13).
U (Lp )
σ (Lp ) = 4.34 ⋅ ε t = (13.19)
k(g)
1000
2 min
100 0.2 dB
1 min
0.5 dB
1 dB
10
20 50 100 200 500 1000 2000 5000 10000
One-third octave center frequency (Hz)
2
2 4.34 ⋅ k(g)
t= (13.20)
B U (Lp )
In Figure 13.13, two-sided test and k(g) = 1.96 for 95 % confidence level is
assumed (Table 13.1).
The total relative spatial variance is:
ε 2 = ε t2 + ε R2 (13.21)
2
t 0.29 ⋅ T + (13.22)
B
where T is the reverberation time of the room and B is the bandwidth of the
measurement. As the last term typically is not significant, a simple rule of
thumb is that the integration time in each position should be at least three
times the reverberation time of the room.
13.4 SPATIAL AVERAGING
13.4.1 Correlation coefficient
The correlation coefficient R between the sound pressures p1 and p2 taken
in two different positions is defined as:
p1 ⋅ p2
R= (13.23)
p12 ⋅ p22
382 Sound Insulation in Buildings
sin(kx)
R= (13.24)
kx
where k = 2π f/c is the wave number and x is the distance between the two
positions. The shape of the function is shown in Figure 1.4. It is noted that
R = 0 for kx = π, and R is relatively small for kx > π.
For practical applications, it can be concluded that
λ
R ≅ 0 for x ≥ (13.25)
2
ε R2
ε2 = (13.26)
N
If for any reason, the distance between the microphone positions cannot
fulfil the minimum requirement (Equation 13.25), it is possible to calcu-
late the equivalent number of uncorrelated positions according to Lubman
(1969):
Measurement uncertainty 383
N
Neq = (i ≠ j) (13.27)
1+
1
N ∑ ∑ R (kx )
i j
2
ij
Here, N is the actual number of microphone positions and xij denotes the
distance between positions i and j. So, all possible pair combinations are
considered in the denominator and R can be calculated from Equation
13.24 if the sound field can be assumed a diffuse sound field. It is seen that
in general Neq ≤ N. It is also noted that Neq need not be a natural number.
Figure 13.14 shows examples of calculated equivalent number of mea-
surement positions as a function of the frequency and for different distances
between the actual positions. At high frequencies, it appears that Neq ≈ N,
but at low frequencies, the positions are not fully independent, and Neq < N.
With the positions equally distributed on a circle with radius r as in the
example in Figure 13.14, the equivalent number of uncorrelated positions
can be estimated by this empirical formula:
3 −1/3
4π
Neq ≅ min {N , Nmax } =min N , 1 + rf (13.28)
c
where f is the centre frequency of the measurement band.
Neq
N = 12
10
N=9
N=6
N=3
0
50 100 200 400 800 1600 Hz 3150
1 V
rmax = − 0.7 m (13.29)
2 1.25 ⋅ h
λ
1 for L≤
2
Neq ≅ (13.30)
2L λ
for L>
λ 2
30
10
Neq
Neq = 1+2L/λ
Circular path
1
0.1 0.3 1 3 10
L/λ
13.5 MEASUREMENT UNCERTAINTY ON
REVERBERATION TIME MEASUREMENTS
13.5.1 General
The uncertainty of decay rate measurements has been studied by Davy
et al. (1979) and Davy (1980). The decay rate d in dB per second is related
to the reverberation time T in seconds by d = 60/T.
386 Sound Insulation in Buildings
σ (T ) 1 ε s2 (d) ε e2 (d) (13.31)
= +
T N d2 n d2
σ (T ) 1+ H / n
=G (13.32)
T N BT
where
T is the reverberation time, in seconds.
σ (T) is the standard deviation of T, in seconds.
G and H are constants that depend on the evaluation range.
n is the number of decays measured in each position.
N is the number of independent measurement positions.
B is the bandwidth, in Hz.
Table 13.6 M inimum number of positions N and decays n for each of the three
quality levels of measurements according to ISO 3382-2
Survey Engineering Precision
No. of positions, N 2 6 12
No. of decays, n (interrupted noise method) 1 2 3
Equivalent no. of decays, n (integrated 10 10 10
impulse response method)
0.20
s(T20)
0.15 s(T30)
Standard deviation (s)
0.10
0.05
0.00
50 100 200 500 1000 2000 5000
One-third octave center frequency (Hz)
1 + (1.90 / n)
σ (T20 ) = 0.88 ⋅ T20 (13.33)
N BT20
This and the analogue result for σ(T30) are shown as functions of the fre-
quency in Figure 13.16 for the example of T20 = T30 = 1.0 s.
been used (see Davy, 1980). For the estimation of the standard deviation of
a measurement result, Equation 13.32 may be used with a value of n = 10.
13.6 MEASUREMENT UNCERTAINTY ON
SOUND INSULATION MEASUREMENTS
S ⋅T
R = Lp1 − Lp2 + 10 lg (dB) (13.34)
0.16 ⋅ V
where S is the area of the partition wall or floor (m 2) and V is the volume of
the receiver room (m3). It is noted that the uncertainty calculation will be
the same in case of DnT, as the difference is only whether the partition area
and the room volume are included or not.
The standard deviation of the last term is estimated from:
σ (T ) σ (T )
σ (10 lg T ) = 10 lg 1 + ≅ 4.34 ⋅ (dB) (13.35)
T T
ε R,1
σ (Lp1) = 4.34 ⋅ (dB) (13.36)
Neq,1
and similarly for the receiver room. Thus, the standard deviation of the
reduction index can be estimated from:
2
σ (T )
σ (R) = σ 2 (Lp1) + σ 2 (Lp2 ) + 4.34 ⋅ (dB) (13.37)
T
Similarly, the standard deviation of the impact sound pressure level can be
estimated from:
Measurement uncertainty 389
2
σ (T )
σ (Ln,T ) = σ 2 (Lp2 ) + 4.34 ⋅ (dB) (13.38)
T
13.6.3 Measurement uncertainty on
single-number quantities
Wittstock (2007, 2009) studied the statistical distribution of measured air-
borne sound insulation, and he found that the distribution function of the
transmission loss R′ is very close to the Gaussian distribution, much closer
than the distribution function of the transmission coefficient τ. So, it makes
sense to apply the statistical methods for uncertainty directly to the trans-
mission loss in dB.
390 Sound Insulation in Buildings
4
Standard deviation σ(R) (dB)
0
50 100 200 500 1000 2000 5000
Centre frequency, one-third octave bands (Hz)
Figure 13.17 Calculated standard deviations in one-third octave bands for the example
case of measured airborne sound insulation.
′ (13.39)
Rw′ − k ⋅ u ≥ Rw,req
Measurement uncertainty 391
Table 13.7 C
alculated standard deviations in one-third octave bands for each of
the terms in Equation 13.37
Frequency (Hz) σ(Lp1) (dB) σ(Lp2) (dB) σ(10 log T) (dB) σ(R) (dB)
50 1.0 3.3 0.9 3.6
63 0.9 4.1 0.8 4.3
80 0.7 3.9 0.7 4.0
100 0.7 3.5 0.6 3.6
125 0.6 3.0 0.6 3.1
160 0.5 2.3 0.5 2.5
200 0.5 2.1 0.5 2.2
250 0.5 1.9 0.4 2.0
315 0.4 1.8 0.4 1.8
400 0.4 1.6 0.3 1.6
500 0.3 1.4 0.3 1.5
630 0.3 1.3 0.3 1.3
800 0.3 1.1 0.2 1.2
1000 0.2 1.0 0.2 1.1
1250 0.2 0.9 0.2 1.0
1600 0.2 0.8 0.2 0.9
2000 0.2 0.7 0.1 0.8
2500 0.1 0.7 0.1 0.7
3150 0.1 0.6 0.1 0.6
4000 0.1 0.5 0.1 0.5
5000 0.1 0.5 0.1 0.5
Note: Input data according to the example.
REFERENCES
J.L. Davy, I.P. Dunn, P. Dubout (1979). The Variance of Decay rates in Reverberation
Rooms. Acustica 43, 12–25.
J.L. Davy (1980). The variance of impulse decays. Acustica 44, 51–56.
P.E. Doak (1959). Fluctuations of the sound pressure level in rooms when the receiver
position is varied. Acustica 9, 1–9.
GUM (1995). Guide to the Expression of Uncertainties in Measurement, 2nd
Edition, International Organization for Standardization, Geneva, Switzerland.
K. Hagberg and P. Thorsson (2010). Uncertainties in standard impact sound measure-
ment and evaluation procedure applied to light weight structures. Proceedings
of ICA 2010, August 23–27, 2010, Sydney, Australia.
ISO 3382-2 (2008). Acoustics – Measurement of room acoustic parameters – Part
2: Reverberation time in ordinary rooms, International Organization for
Standardization, Geneva, Switzerland.
ISO 12999-1 (2014). Acoustics – Determination and application of measurement
uncertainties in building acoustics – Part 1: Sound insulation, International
Organization for Standardization, Geneva, Switzerland.
ISO 18233 (2006). Acoustics – Application of new measurement methods in building
and room acoustics, International Organization for Standardization, Geneva,
Switzerland.
F. Jacobsen (1979). The diffuse sound field. Statistical considerations concerning the
reverberant field in the steady state. Report No. 27. The Acoustics Laboratory.
Technical University of Denmark.
H. Kuttruff (1979). Room Acoustics, 2nd Edition, Applied Science Publishers,
London.
D. Lubman (1968). Fluctuations of sound with position in a reverberant room.
Journal Acoustical Society America 44, 1491–1502.
D. Lubman (1969). Spatial averaging in a diffuse sound field. Journal Acoustical
Society America 46, 532–534.
D. Lubman (1971). Spatial averaging in sound power measurements. Journal of
Sound Vibrations 16, 43–58.
D. Lubman (1974). Precision of reverberant sound power measurements. Journal of
Acoustical Society America 56, 523–533.
D. Lubman, R.V. Waterhouse, C.S. Chien (1973). Effectiveness of continuous spa-
tial averaging in a diffuse sound field. Journal Acoustical Society America 53,
650–659.
M. Machimbarrena, C.R.A. Monteiro, S. Pedersoli, R. Johansson, S. Smith (2015).
Uncertainty determination of in situ airborne sound insulation measurements.
Applied Acoustics 89, 199–210.
Measurement uncertainty 393
14.1.1 Noise descriptors
International standard ISO 1996-1 defines noise descriptors applied to
environmental noise, for example, for noise regulations and noise map-
ping. They are all some kind of time-averaged, A-weighted sound pressure
level in the free field. The latter means that even if the sound is measured or
calculated close to a building, the result is corrected for the possible influ-
ence of reflections from that building.
If we do not take into account the distribution over day and night, the
sound energy is integrated over 24 h and we get Lp,A,24 h. However, preferred
descriptors do take into account this distribution. The day-evening-night
level L den is defined as:
1
(
24 h tday ⋅ 10
0.1 Lday,12
+ tevening
Lden = 10 lg , dB (14.1)
⋅100.1 (Levening,4 +5 dB) + t night ⋅ 100.1 (Lnight,8 +10 dB) )
where tday, tevening, tnight are the number of hours of the periods (24 h in total),
and L day, L evening and L night are the A-weighted, time-averaged sound pres-
sure levels in the same periods. Note the penalty of 5 dB in the evening and
395
396 Sound Insulation in Buildings
Table 14.1 Relation between some noise descriptors valid for road
traffic noise with typical example of traffic distribution
78 % (day, 12 h), 11 % (evening, 3 h) and 11 % (night, 9 h)
Level (dB) Difference (dB)
Lp,A,24 h 55 0
Lday,12 h 57 2
Lday,15 h 57 2
Levening,3 h 54 −1
Lnight,9 h 50 −5
Lden 58 3
Ldn 58 3
10 dB in the night. The definition of the three periods may differ from one
country to another; for example, the EU member states use 3 h evening and
9 h night periods. Sometimes, the evening period is not used at all, and we
get the day-night level L dn defined as:
1
Ldn = 10 lg
24 h
( )
tday ⋅ 100.1 Lday,15 + t night ⋅ 100.1 (Lnight,9 +10 dB) dB (14.2)
For road traffic noise having typical traffic distribution, the relation
between the noise descriptors is shown in Table 14.1.
14.1.2 Sleep disturbance
The World Health Organization has identified neighbour noise as a health
problem based on the findings from a large survey in eight European cities
from 2002 to 2003, the Large Analysis and Review of European hous-
ing and health Status (LARES) project (WHO, 2007). Information was
collected in 3382 households from 8539 people by interviews and health
questionnaire. Of all responding residents, 24 % reported that noise expo-
sure at night was a main reason for sleep disturbance. Traffic noise was
the dominant cause of sleep disturbance, closely followed by noise from
neighbouring flats (Figure 14.1).
While traffic noise has since long been known to be a serious source of
annoyance, the LARES investigation showed that also chronic annoyance
due to neighbour noise is associated with hypertension, depression, and
migraine (Maschke and Niemann, 2007). They concluded that neighbour
noise annoyance is a highly underestimated risk factor for healthy housing.
Noise from neighbours may affect people in different ways. Figure 14.2
illustrates how some people may react with curiosity, changing to annoy-
ance and anger and, in severe cases, ending with hatred and other similar
Noise effects and subjective evaluation of sound insulation 397
50
45
Percentage with sleep disturbance
40
35
30
25
20
15
10
0
Traffic Neighbour flat Parking Surrounding Animals Staircase use
areas
Figure 14.1 Sleep disturbance in adult residents and main sources of disturbing noise.
Total for eight European cities (n = 8539). (Adapted from WHO, Large Analysis
and Review of European Housing and Health Status (LARES). Preliminary over-
view, WHO Regional Office for Europe, Copenhagen, Denmark, 2007, Figure
37. With permission.)
reactions. Other people may react with irritation, growing to tension and
depression. Noise annoyance is related to health (Niemann et al., 2006):
annoyance → negative emotional reaction → neuro-vegetative-hormonal
regulatory disturbances → illness.
398 Sound Insulation in Buildings
14.1.3 Dose-effect relationships
Risk ratio (RR) and odds ratio (OR) are used in statistics, for instance,
to calculate dose-effect relationships. With reference to Table 14.2, the
investigated people are divided into groups (e.g. based on different level
of noise exposure), and the data are collected to show how many exhibit
a certain effect or not (e.g. stroke). Group 0 is a reference group, in which
the dose-effect relationship is assumed to be negligible. In each group, the
risk of being affected is a number between 0 and 1, and the RR is the ratio
between the risk in that group and the risk in the reference group. For
example, RR = 1.05 means that there is an increased risk of 5 % compared
to the reference group.
The odds for being affected in a group are simply the ratio between the
number being affected and the number not affected. The OR in a group is
then the odds in that group divided by the odds in the reference group. If
the effect is sufficiently rare (Yes << No), the risk in all groups may be much
less than unity and then RR ≈ OR.
Annoyance data for 8.6 million people living in London between 2003
and 2010 were analysed in relation to road traffic noise exposure in noisy
areas, L day (07–19) > 60 dB relative to quiet areas L day < 50 dB (Halonen
et al., 2015). Deaths were 4 % more common among adults and elderly
in noisy areas compared with quiet areas (Table 14.3). Adults living in
noisy areas were 5 % more likely to be admitted to hospital for stroke
compared to those living in quiet areas and went up to 9 % in the elderly
population.
In addition to annoyance, sleep disturbance and health problems, traffic
noise may have negative effects on children’s behaviour. In a large study, in
Denmark, 46 904 children with behavioural problems at the age of 7 years
were identified, and address from birth to 7 years was used to model the
traffic noise exposure. It was found that road traffic noise in early childhood
Table 14.2 Calculation of RR and OR for being affected (symptom, disease etc.)
when belonging to a certain group (e.g. based on level of noise exposure)
Effect
Yes No Risk Odds RR OR
Y0 Y0
Group 0 (reference) Y0 N0 1 1
(Y0 + N0 ) N0
Y1 Y1 Y1(Y0 + N0 ) Y1N0
Group 1 Y1 N1
(Y1 + N1) N1 Y0 (Y1 + N1) Y0 N1
Y2 Y2 Y2 (Y0 + N0 ) Y2 N0
Group 2 Y2 N2
(Y2 + N2 ) N2 Y0 (Y2 + N2 ) Y0 N2
Noise effects and subjective evaluation of sound insulation 399
isk ratio of death or stroke due to road traffic noise L day > 60 dB
Table 14.3 R
relative to L day < 50 dB
Risk ratio Confidence
(RR) interval 95 %
Death Adults (>25) 1.04 (1.00–1.07)
Elderly (>75) 1.04 (1.00–1.08)
Stroke Adults (>25) 1.05 (1.02–1.09)
Elderly (>75) 1.09 (1.04–1.14)
Source: Halonen, J.I. et al., Eur. Heart J., 36(39), 2653–2661, 2015.
1.6
1.5
1.4
1.3
Odds ratio
1.2
1.1
0.9
0.8
50 55 60 65
Exposure to road traffic noise, Lden (dB)
Figure 14.3 Relation between exposure to road traffic noise at childhood and abnor-
mal scores on total difficulties shown as odds ratio with 95 % confidence
intervals in four noise exposure categories. The reference category is Lden
< 50 dB. (Reproduced from Hjortebjerg, D. et al., Environ. Health Perspect.,
124(2), 228–234, 2016. With permission.)
400 Sound Insulation in Buildings
barriers; but, in fact, such measures for noise protection of housing areas with
high noise exposure are common in Denmark since the 1970s. In the act of
environmental protection, it was stated that the number of dwellings exposed
to noise levels more than 65 dB from road traffic should be reduced to a mini-
mum of 50 000 before 2010. The typical attenuation of road traffic noise due
to a noise barrier is 8 dB to 15 dB. If the result in the highest noise group in
Figure 14.3 is shifted downwards by 12 dB, it fits precisely on the curve.
14.2.1 Acoustical comfort
The concept of acoustical comfort in relation to buildings was first used by
Cummins (1978). Acoustical comfort is characterised as follows (Rindel
and Rasmussen, 1995):
35
30
Percentage annoyed (moderately + very + extremely)
25
20
15
10
0
Speech, etc. Music, etc. Footfall Traffic noise Noise (all
sources)
40
35
30
25
Percentage
20
15
10
0
Speech, etc. Music, etc. Footfall
Annoyed (moderately + very + extremely) Concerned to disturb neighbours
Figure 14.5 Percentage of people being annoyed by various noise sources in their home,
and percentage of people being concerned that they may disturb their neigh-
bours by various activities (n = 702). (Adapted from Løvstad et al., Sound
quality in dwellings in Norway – A socio-acoustic investigation. Proceedings
of BNAM 2016, Stockholm, Sweden, 2016.)
(Figure 14.5). More than 15 % were concerned (to some extent or very
much) about disturbing their neighbours due to footfall noise. These find-
ings clearly demonstrate that the constraints dwellers put on their own
activities due to insufficient sound insulation are at least as important as
the annoyance from noise. Noise from neighbours is like a troll with two
heads; the fear of disturbing the neighbour is half the problem.
has a significant correlation with the measured sound insulation, but not
with the actual noise level at the neighbours. Bradley (1983) states:
‘This is strong evidence that in this study at least the inadequacy
of the party wall was a source of social disruption in that neigh-
bors were thought to be inconsiderate when it was really the
party wall that was at fault.’
100
80
Percentage dissatisfied
60
40
20
0
40 45 50 55 60 65
Vertical sound insulation, R΄w (dB )
Figure 14.6 Proportion of residents in multi-storey flats who rate the sound insulation
as poor, presented as a function of the measured airborne sound insulation
through the separating floors. Average for 16 housing estates and regression
line (n = 471). (Weeber, R. et al., Schallschutz in Mehrfamilienhäusern aus der Sicht
der Bewohner (Sound insulation in multi-family houses from the point of view
of the resident, in German). F 2049, IRB Verlag, Stuttgart, Germany, 1986.)
404 Sound Insulation in Buildings
14.3 DOSE-RESPONSE FUNCTIONS
14.3.1 Statistical methods
Two different statistical models are usually considered, probit or logit. The
probit model is based on the normal distribution (the Gaussian d istribution:
see also Chapter 13). However, when cumulative probability functions
are evaluated and adapted to observed data, it is a disadvantage that these
functions are in integral forms. Instead, the logit model is based on slightly
different distributions often assumed in social surveys. The principle behind
the logit model is to consider the data to be analysed in proportions, preferably
with equal uncertainty attached to each proportion. The logit transforma-
tion achieves this. It is defined by Altman (1991, p. 145) as follows:
p
t = logit(p) = ln (14.3)
1 − p
et
⇒ p= (14.4)
1 + et
where p is a proportion and t is the logit. The logit transformation stretches out
the lower and upper parts of the distribution in the same way as in the Gaussian
distribution (Table 14.4). They both belong to the group of sigmoid functions
(S-shaped curves), which are used to model dose-response relationships.
The logistic function has the probability density function:
z− µ
−
e s
w(z) = (14.5)
z− µ 2
−
s 1 + e s
where the mean value is μ and s is a parameter proportional to the standard
deviation σ. The cumulative distribution function is:
Table 14.4 E
ffect of logit transformation
of a proportion p
p logit(p)
0.01 −4.60
0.05 −2.94
0.10 −2.20
0.25 −1.10
0.50 0.00
0.75 1.10
0.90 2.20
0.95 2.94
0.99 4.60
406 Sound Insulation in Buildings
0.5 0.8
0.4
0.3 0.6
w(z)
0.2
P(z)
0.1 0.4
0.0
–3 –2 –1 0 1 2 3
Logistic Gauss 0.2
Measurand, z
0.0
–3 –2 –1 0 1 2 3
Logistic Gauss
Measurand, z
1 1 1 z − µ (14.6)
P(z) = = + tanh
−
z− µ
2 2 2s
1+ e s
1 2 2
This is a symmetric distribution with the variance σ = s π . The func-
2
3
tions are shown in Figure 14.7, and it is seen that the cumulative distribution
is very close to that of a Gaussian distribution. An advantage of the logistic
function is that the hyperbolic tangent can be used for calculations, whereas
the Gaussian cumulative distribution can only be expressed in integral form.
With the logistic function, the probability for a random sample to be
within the range μ ± σ is 72 %, i.e. somewhat higher than with the Gaussian
distribution (68 %). The probability to be within an extended range of
μ ± 2σ is 95 % with both distributions (Table 14.5).
The proportion of the distribution in the extreme ranges is called
the tails of the distribution. The proportion outside ±3 standard devia-
tions is 0.9 % in the logistic distribution, but only 0.3 % in the Gaussian
Table 14.5 C
omparison of the logistic and the Gaussian distributions
Range ±1 σ ±2 σ ±3 σ
Logit (%) 72.0 94.8 99.1
Gaussian (%) 68.3 95.4 99.7
Note: Probability of being within a range of ±1, ±2 or ±3 standard deviations.
Noise effects and subjective evaluation of sound insulation 407
distribution (Table 14.5). Thus, the logistic distribution has heavier tails
than the Gaussian distribution, and this makes it more robust to inaccura-
cies in the model or to errors in the data. This is another advantage of the
logit model compared to the probit model.
90
Percentage annoyed (>2 complaints of 18)
80
70
60
50
40
30
20
10
0
45 50 55 60 65 70 75
Lp,A,24 h (dB)
level of the houses. It was thought to be very important for the later statisti-
cal analysis that there was a sufficient spread in the noise exposure, and that
there were none in the middle range. Noise measurements were made with
24 h integration time and the microphone on a building façade. Free field
Lp,A,24 h results were derived by subtraction of 5 dB due to the microphone
position. This is the noise parameter shown on the abscissa in Figure 14.8.
In total, 960 personal interviews were made among the dwellers in August
and September 1972. The purpose was to find out how traffic noise might
influence the psychical and physical health of the inhabitants, but this purpose
was hidden. The investigation contained in total 75 questions, out of which 18
were related to annoyance due to road traffic noise, e.g. if it was difficult to fall
asleep in the evening, or if they did not open the windows because of the noise.
In a preliminary analysis of the results, it was found that a good indicator of
noise annoyance could be the number of complaints among the 18 relevant
questions and the median of all questionnaires was two complaints out of 18.
Thus, the annoyance parameter on the ordinate shown in Figure 14.8 is the
percentage in each housing area that gave more than two complaints.
The annoyance data were averaged for each housing area (on average 34
interviewed persons in each area). A regression analysis was made using
the results from the 28 housing areas, as plotted in Figure 14.8. A sigmoid
function was assumed, and the arcsine function was used as an approxima-
tion. The curve shows, for example, that if the free field Lp,A,24 h increases
from 55 dB to 65 dB, the percentage annoyed increases statistically from
14 % to 55 %. In the middle range between 20 % and 80 % the slope of the
dose-response curve is ~4 % points per dB.
The dose-response relationships can also be derived from laboratory
experiments. The advantages are that the parameters can be chosen to cover
a wide range of exposure, which is important for a good statistical analysis.
The connection between exposure and respondent can be controlled, and
there is little risk of confounders that could lead to errors. The drawbacks
of laboratory experiments include lack of realism, only short-time expo-
sure is possible, and the number of test persons is normally more limited
than in a field investigation. Figure 14.9 is an example of a dose-response
relationship found in a laboratory experiment, where a recording of road
traffic noise was used at seven different levels in 5 dB steps from 40 dB to
75 dB. The test persons were (among many other things) asked whether the
noise would be acceptable if they worked in an office with that noise. Probit
analysis was used to estimate the dose-response curve. The results show
that in an office, 10 % would be dissatisfied if the noise level exceeds 45 dB.
The average slope of the curve in the middle range is 4 % points per dB.
14.3.3 Socio-acoustic surveys
In a socio-acoustic survey, the questions asked about noise problems are very
important for the quality of the results. The way in which such questions are
Noise effects and subjective evaluation of sound insulation 409
100
80
Percentage dissatisfied
60
40
20
0
35 40 45 50 55 60 65 70 75 80
Noise level, Lp,A,T (dB )
asked and the wording of the possible answers should follow the recommenda-
tions in Technical Specification ISO/TS 15666. Questions and possible answers
about how much the person is bothered/disturbed/annoyed can be like this:
“Thinking about the last 12 months, when you are at home, how much
does noise from traffic annoy you? Not at all – Slightly – Moderately –
Very – Extremely?” A numerical rating scale may also be used, e.g. with
numbers from 0 to 10, where 0 means not at all and 10 means extremely.
In addition, there must be a possibility for answering “do not know”.
Researchers in this field do not agree on using a numerical scale. However,
ISO/TS 15666 is considered the best practice agreed on. The standard also
contains an Annex B with recommended wordings of the questions in sev-
eral languages. The words used for the answers have been balanced in such
a way that the rating between the words is about 25 %.
An example is shown in Figure 14.10, which is one of the results from
a socio-acoustic survey (Amundsen et al., 2011). From the replies of 738
people in combination with known traffic noise exposure, the four curves
in Figure 14.10 were derived by logit analysis. Five regions are indicated cor-
responding to the distribution of the replies from the interviews. For example
at a noise level of 75 dB, 10 % were not annoyed, 20 % slightly annoyed, 30 %
moderately annoyed, 31 % very annoyed, and 9 % extremely annoyed (Table
14.6). The often used outdoor noise limit Lp,A,24 h = 55 dB corresponds approxi-
mately to 60 % not annoyed and 15 % annoyed (moderately + very + extremely).
Each of the four curves have the same shape, since a logistic distribution was
410 Sound Insulation in Buildings
100
1
90
80
2
Cumulative proportion (%)
70
60
3
50
40
30
4
20
10
5
0
40 45 50 55 60 65 70 75 80
Lp,A,24 hOutdoor (dB)
Figure 14.10 Cumulative annoyance curves for road traffic noise in a dwelling as a func-
tion of the outdoor A-weighted sound pressure level. The curves are the
result of logit analysis (n = 738). Region 1: Not annoyed. Region 2: Slightly
annoyed. Region 3: Moderately annoyed. Region 4: Very annoyed. Region 5:
Extremely annoyed.
Table 14.6 A
nnoyance due to road traffic noise heard in the dwelling at selected
levels of the outdoor noise
Outdoor sound pressure level Lp,A, 24 h
Degree of annoyance 45 dB (%) 55 dB (%) 65 dB (%) 75 dB (%)
1 Not annoyed 83 59 29 10
2 Slightly annoyed 12 56 32 20
3 Moderately annoyed 3 10 23 30
4 Very annoyed 1 4 13 31
5 Extremely annoyed 0 1 3 9
Source: Amundsen, A.H. et al., J. Acoust. Soc. Am., 129, 1381–1389, 2011.
Note: Data from logit analysis in Figure 14.10.
assumed, and the curves are shifted along the dB-axis by approximately 10 %,
10 % and 15 %, respectively. The slope of the curves in the middle range is
~3 % points per dB.
In the same investigation, the indoor noise level due to road traffic noise
was estimated and the result of the analysis is seen in Figure 14.11. The
often used indoor limit of Lp,A,24 h = 30 dB corresponds approximately to
40 % not annoyed and 30 % annoyed (moderately + very + extremely).
Noise effects and subjective evaluation of sound insulation 411
100
90
80
70
Cumulative proportion (%)
60
50
40
30
20
10
0
15 20 25 30 35 40 45 50
Lp,A,24 h Indoor (dB)
Slightly annoyed Moderately annoyed
Very annoyed Extremely annoyed
Figure 14.11 Cumulative annoyance curves for road traffic noise in a dwelling as a func-
tion of the indoor A-weighted sound pressure level. The curves are the
result of logit analysis (n = 738). (Adapted from Amundsen, A.H. et al., J.
Acoust. Soc. Am., 129, 1381–1389, 2011.)
(
pHA = 100 ⋅ exp − 100.1 (Ldn −Lct +5.3 dB) ) , % (14.7)
−0,3
of the community tolerance level for different traffic noise sources is given
in Table 14.7, together with the standard deviations. The 95 % prediction
intervals can be estimated as ±2 standard deviations. The dose-response
curves for aircraft, road traffic, and railroad are shown in Figure 14.12.
Because of the different character of the noise sources (especially different
variation with time), the annoyance score is quite different for the same
day-night level. Aircraft is more annoying than road traffic (correspond-
ing to +5 dB), while railroad is less annoying (corresponding to −9.5 dB).
However, in cases where noise from railroad is accompanied by strong
Table 14.7 Community tolerance level L ct (mean and standard deviation) for
different traffic noise sources
Railroad (low Railroad (high
Aircraft Road traffic vibration) vibration)
Mean Lct, dB 73.3 78.3 87.8 75.8
Standard deviation, dB 7.1 5.1 3.5 4.2
Difference from road 5.0 0.0 −9.5 2.5
traffic, dB
Source: ISO 1996-1, Acoustics – Description, measurement and assessment of environmental
noise – Part 1: Basic quantities and assessment procedures. International Organization
for Standardization, Geneva, Switzerland, 2016.
60
50
Percentage highly annoyed
40
30
20
10
0
45 50 55 60 65 70 75
Ldn (dB)
Aircraft Road traffic Railroad (low vibration)
Figure 14.12 Dose-response curves for aircraft, road traffic and railroad (low vibration).
(Adapted from ISO 1996-1, Acoustics – Description, measurement and assess-
ment of environmental noise – Part 1: Basic quantities and assessment procedures.
International Organization for Standardization, Geneva, Switzerland, 2016.)
Noise effects and subjective evaluation of sound insulation 413
vibrations, the annoyance score is much higher, between that for aircraft
and road traffic (not shown in Figure 14.12).
It is interesting to compare the curve for road traffic in Figure 14.12 with the
previous Figures 14.8 and 14.10. First, we recall that for a typical day-night dis-
tribution of traffic, the outdoor Lp,A,24 h = 55 dB corresponds to Ldn = 58 dB (Table
14.1). Figure 14.12 then shows 6 % highly annoyed from road traffic. In compar-
ison, Figure 14.8 shows 15 % annoyed. Figure 14.10 also shows 15 % annoyed
(moderately + very + extremely) but only 5 % highly annoyed (very + extremely);
the latter being close to the 6 % highly annoyed in Figure 14.12. This demon-
strates clearly the importance of how we define ‘annoyed’ or ‘highly annoyed’.
In future socio-acoustic investigations, this problem may be avoided if
the method in ISO/TS 15666 for such surveys is applied.
Table 14.8 Data from measured airborne sound insulation and degree of
annoyance due to speech and music
Horizontal Vertical
R′w R′w,50 DnT,w DnT,w,50 R′w R′w,50 DnT,w DnT,w,50
Measurements
Number, N 355 346 296 296 394 354 366 349
Min, dB 46 45 45 44 50 50 52 51
Max, dB 64 63 65 64 69 68 70 69
Mean, dB 56.8 53.8 58.6 55.4 61.6 58.7 61.0 57.8
Standard dev., dB 3.0 2.8 4.0 3.0 4.3 2.6 4.2 2.4
Annoyed due to 19.9 20.1 20.0 19.9 22.7 22.0 21.9 21.8
speech etc., (%)
Annoyed due to 27.8 26.4 25.4 24.8 27.3 25.4 25.4 26.5
music with bass
(%)
Source: Rindel, J.H. et al., Dose-response curves for satisfactory sound insulation between
dwellings, Proceedings of ICBEN 2017, Zürich, Switzerland, 2017.
Note: Percentage annoyed includes all levels (extremely, very, moderately, slightly).
414 Sound Insulation in Buildings
The upper part of Table 14.8 shows the number of measurements, the
range of variation, the mean value, and the standard deviation. It is noted
that airborne sound insulation is generally better in the vertical direction
than in the horizontal direction; mean values of R′w and R′w,50 are about
5 dB better for floors than for walls. This suggests that annoyance from
speech or music should be related mostly to the horizontal sound insula-
tion. If this is correct, it means that the dose-response results for airborne
sound insulation in the vertical direction cannot be reliable. In the subjec-
tive response from the questionnaire, there is no distinction between hori-
zontal or vertical direction of the annoying sound.
The degree of annoyance from speech is less than that from music. It is
also noted that the annoyance from speech seems to be higher in the vertical
direction than in the horizontal direction. However, this is simply a conse-
quence of the fact, that the average sound insulation of the floors is about
5 dB better than that of the walls. The degree of annoyance (all levels) corre-
sponding to the mean value of sound insulation is on average 26 % for music
and 21 % for speech. The cumulative annoyance curves were quite similar
for speech and music, and only the latter are shown here in Figure 14.13.
100
90
80
Cumulative proportion (%)
70
60
50
40
30
20
10
0
45 50 55 60 65
R΄w Horizontal (dB)
Slightly annoyed Moderately annoyed
Very annoyed Extremely annoyed
Figure 14.13 Cumulative annoyance curves due to music with bass as a function of the
normalised level difference between dwellings in the horizontal direction.
The curves are the result of logit analysis (n = 702). (Adapted from Rindel,
J.H. et al., Aiming at satisfactory sound conditions in dwellings – The use of
dose-response curves. Proceedings of BNAM 2016, Stockholm, Sweden, 2016.)
Noise effects and subjective evaluation of sound insulation 415
100
90
80
Cumulative proportion (%)
70
60
50
40
30
20
10
0
40 45 50 55 60
L΄n,w + CI,50–2500 (dB)
Slightly annoyed Moderately annoyed
Very annoyed Extremely annoyed
Figure 14.14 Cumulative annoyance curves due to footfall noise from above as func-
tion of the normalized impact sound pressure level including the 50 Hz
adaptation term. The curves are the result of logit analysis (n = 702).
(Adapted from Rindel, J.H. et al., Aiming at satisfactory sound conditions
in dwellings – The use of dose-response curves. Proceedings of BNAM 2016,
Stockholm, Sweden, 2016.)
416 Sound Insulation in Buildings
14.3.7 Vibrations
A survey on vibrations in dwellings due to traffic (road and rail) was carried
out in Norway (Klæboe et al., 2003; Turunen-Rise et al., 2003). About 1500
people were interviewed and one question was: “Can you in your dwelling
notice shaking or vibration caused by <source>? If yes, is the shaking/these
vibrations highly annoying, somewhat annoying, a little annoying or not
annoying for you?”
Measurements of vibrations were performed in the dwellings and out-
side on the ground. A semi-empirical model was used to determine the
statistical maximum RMS vibration velocity using the Wm weighting filter
(ISO 2631-2) (see Figure 2.18). The statistical maximum is defined as the
95 percentile, i.e. the value not exceeded in 95 % of the time. The results
of the logit analysis using the velocity level 20 lg (v/vref) dB are shown in
Figure 14.15.
The curves show that about 7 % will be highly annoyed at a weighted vibra-
tion velocity of 0.3 mm/s. At a velocity of 0.5 mm/s, 10 % will be highly
annoyed and about 40 % will annoyed slightly + moderately + highly.
90
80
Cumulative proportion (%)
70
60
50
40
30
20
10
0
90 100 110 120 130 140
Velocity level, Lv,w,95, dB re 1 nm/s
Perceives vibrations Slightly annoyed
Moderately annoyed Highly annoyed
Figure 14.15 Cumulative annoyance curves for vibrations in the dwelling due to road and
rail traffic. The curves are the result of logit analysis (n = 1427). (Adapted
from Turunen-Rise, I.H. et al., Appl. Acoust., 64, 71–87, 2003.)
Noise effects and subjective evaluation of sound insulation 417
Table 14.9 Results of laboratory experiment with airborne sound and simulated
constructions with different sound insulation below 160 Hz
Music, percent
Type of construction R′w (dB) R′w + C50–3150 (dB) Lp,A,T (dB) annoyed
Light 56 49 48 98
Light-medium 56 53 43 90
Medium 57 55 37 80
Medium-heavy 57 56 36 83
Heavy 57 56 35 83
Source: Mortensen, F.R., Subjective Evaluation of Noise from Neighbours with Focus on Low
Frequencies. Main report. Publication no. 53, Department of Acoustic Technology,
Technical University of Denmark, 1999.
Table 14.10 R
esults of laboratory experiment with impact sound and simulated
constructions with different sound insulation below 125 Hz
Walker, Children,
Type of L′n,w L′n,w + CI,50– Walker, Children, percent percent
construction (dB) 2500 (dB) Lp,A,T (dB) Lp,A,T (dB) annoyed annoyed
Light 55 62 39 41 71 81
Light-medium 55 58 32 35 51 78
Medium 55 56 27 30 36 47
Medium-heavy 55 55 25 28 28 51
Heavy 54 54 24 25 20 47
Source: Mortensen, F.R., Subjective Evaluation of Noise from Neighbours with Focus on Low
Frequencies. Main report. Publication no. 53, Department of Acoustic Technology,
Technical University of Denmark, 1999.
From the results of this investigation, it was concluded that the spectrum
adaptation terms in ISO 717-1 and ISO 717-2 with extended frequency
range down to 50 Hz give a better correlation between subjective and objec-
tive evaluation of sound insulation.
While the importance of the extended frequency range down to 50 Hz
is clear in relation to impact sound, and other researchers have come to a
similar conclusion (Warnock, 1998; Späh et al., 2014), the case of airborne
sound insulation is not so clear. Only when the sound is music with strong
contents of low frequencies, the 50 Hz adaptation term gives an improved
correlation with subjective impressions; but for other kinds of noise from
neighbours like speech, TV, etc., there is no need to include the frequencies
below 100 Hz for airborne sound insulation (Kylliäinen et al., 2016).
The spectrum and the frequency range are very different in music and
speech. A comprehensive investigation based on listening tests using
speech and music in combination with 20 partition walls with quite dif-
ferent acoustic characteristic was performed (Park et al., 2008; Park and
Noise effects and subjective evaluation of sound insulation 419
Bradley, 2009a,b). Three music samples were selected for the tests, all of
them being potentially annoying but of different musical styles. It was
found that the optimum range of included frequencies in the single-number
quantity is different for music sounds and speech sounds. Some measures
were strongly related to ratings of speech sounds and others to ratings of
music sounds, but none were highly successful for both types of sounds.
The best correlation with annoyance from music sounds was a parameter
defined as the A-weighted sound pressure level difference with pink noise,
63 Hz to 6300 Hz (R 2 = 0.978). For speech sounds, the best correlation
with annoyance was a parameter defined as the A-weighted sound pres-
sure level difference with pink noise, 200 Hz to 6300 Hz (R 2 = 0.977). It
was also found that the signal-to-noise ratio was well related to annoy-
ance ratings and audibility ratings of transmitted music sounds. Almost
all subjects found that sounds were just audible at signal-to-noise levels as
low as −10 dB. The thump of the rhythmic beating, which is typical in loud
music, may explain this finding.
From the point of view that sound insulation requirements are meant to
ensure a reasonable degree of protection against all typical sounds from
neighbours, the one-third octave bands 50 Hz to 80 Hz are important and
should be included in the evaluation of airborne sound insulation in dwell-
ings (Rindel, 2017).
Table 14.11 Acoustic design criteria for dwellings and the corresponding
expected percentage of people finding conditions satisfactory (not annoyed) or
not satisfactory (annoyed to some degree)
Satisfied
90 80 70 60 50 40 30 20
(not annoyed), %
Annoyed
(moderately +
3 5–8 9–12 15–18 20–25 27–35 37–45 50–55
very +
extremely), %
Airborne R′w, dB 64 59 56 53.5 51 48.5 46 43
sound R′w,50, dB 62 57 54 51.5 49 46.5 44 41
insulation
DnT,w, dB 66 61 58 55.5 53 50.5 48 45
DnT,w,50, dB 64 59 56 53.5 51 48.5 46 43
Impact sound L′n,w,50, dB 42.5 47.5 50.5 53 55.5 58 60.5 63.5
level L′nT,w,50, dB 39 44 47 49.5 52 54.5 57 60
Noise indoor, Lp,A,24 h, dB 16 21 24 26.5 29 31.5 34 37
road traffic
Noise outdoor, Lp,A,24 h, dB 47 52 55 57.5 60 62.5 65 68
road traffic Lden, dB 50 55 58 60.5 63 65.5 68 71
Noise effects and subjective evaluation of sound insulation 421
REFERENCES
ISO 717 (1996). Acoustics – Rating of sound insulation in buildings and of building
elements. Part 1: Airborne sound insulation. Part 2: Impact sound insulation.
International Organization for Standardization, Geneva, Switzerland.
R. Klæboe, I.H. Turunen-Rise, L. Hårvik, C. Madshus (2003). Vibration in dwell-
ings from road and rail traffic – Part II: exposure-effect relationships based on
ordinal logit and logistic regression models. Applied Acoustics 64, 89–109.
J. Kragh (1977). Analysis of the Correlation between Reactions to Road Traffic Noise
and Physical Indices of the Noise. Report No. 5 (In Danish). The Acoustical
Laboratory, The Danish Academy of Technical Sciences, Lyngby, Denmark.
M. Kylliäinen, J. Takala, D. Oliva, V. Hongisto (2016). Justification of standard-
ized level differences in rating of airborne sound insulation between dwellings.
Applied Acoustics 102, 12–18.
J. Lang and H. Muellner (2013). The importance of music as sound source in resi-
dential buildings. Proceedings of Inter Noise 2013, 15–18 September 2013,
Innsbruck, Austria.
A. Løvstad, J.H. Rindel, C.O. Høsøien, I. Milford, R. Klæboe (2016). Sound qual-
ity in dwellings in Norway – A socio-acoustic investigation. Proceedings of
BNAM 2016, 20–22 June 2016, Stockholm, Sweden.
C. Maschke and H. Niemann (2007). Health effects of annoyance induced by neigh-
bour noise. Noise Control Engineering Journal 55(3), 348–356.
H.M.E. Miedema (2004). Relationship between exposure to multiple noise sources
and noise annoyance. Journal of the Acoustical Society of America 116,
949–957.
F.R. Mortensen (1999). Subjective Evaluation of Noise from Neighbours with Focus
on Low Frequencies. Main report. Publication no. 53, Department of Acoustic
Technology, Technical University of Denmark.
H. Niemann, X. Bonnefoy, M. Braubach, K. Hecht, C. Maschke, C. Rodrigues, N.
Röbbel (2006). Noise-induced annoyance and morbidity results from the pan-
European LARES study. Noise & Health 8(31), 63–79.
H.K. Park, J.S. Bradley, B.N. Gover (2008). Evaluating airborne sound insulation
in terms of speech intelligibility. Journal of the Acoustical Society of America
123, 1458–1471.
H.K. Park and J.S. Bradley (2009a). Evaluating standard airborne sound insulation
measures in terms of annoyance, loudness, and audibility ratings. Journal of
the Acoustical Society of America 126, 208–219.
H.K. Park and J.S. Bradley (2009b). Evaluating signal-to-noise ratios, loudness, and
related measures as indicators of airborne sound insulation. Journal of the
Acoustical Society of America 126, 1219–1230.
J.H. Rindel (1999). Acoustic quality and sound insulation between dwellings. Journal
of Building Acoustics 5, 291–301.
J.H. Rindel (2017). A comment on the importance of low frequency airborne sound
insulation between dwellings. Acta Acustica/Acustica 103, 164–168.
J.H. Rindel, A. Løvstad, R. Klæboe (2016). Aiming at satisfactory sound conditions
in dwellings – The use of dose-response curves. Proceedings of BNAM 2016,
Stockholm, Sweden.
J.H. Rindel, A. Løvstad, R. Klæboe (2017). Dose-response curves for satisfactory
sound insulation between dwellings. Proceedings of ICBEN 2017, 18–22 June
2017, Zürich, Switzerland.
Noise effects and subjective evaluation of sound insulation 423
J.H. Rindel, B. Rasmussen (1995). Buildings for the future: The concept of acoustical
comfort and how to achieve satisfactory acoustical conditions with new build-
ings. (18 pages). COMETT-SAVOIR Course, CSTB, Grenoble.
P. Schomer, V. Mestre, S. Fidell, B. Berry, T. Gjestland, M. Vallet, T. Reid (2012). Pole
of community tolerance level (CTL) in predicting the prevalence of the annoy-
ance of road and rail noise. Journal of the Acoustical Society of America 131,
2772–2786.
T.J. Schultz (1978). Synthesis of social surveys on noise annoyance. Journal of the
Acoustical Society of America 64, 377–405.
M. Späh, K. Hagberg, O. Bartlomé, L. Weber, P. Leistner, A. Liebl (2014). Subjective
and objective evaluation of impact noise sources in wooden buildings. Noise
Notes 13, 25–42.
I.H. Turunen-Rise, A. Brekke, L. Hårvik, C. Madshus, R. Klæboe (2003). Vibration
in dwellings from road and rail traffic – Part I: A new Norwegian measurement
standard and classification system. Applied Acoustics 64, 71–87.
A.C.C. Warnock (1998). Floor research at NRC Canada. Proceedings of COST
Action E5 Workshop, Acoustic performance of medium-rise timber buildings,
3–4 December 1998, Dublin, (10 pages).
R. Weeber, H. Merkel, H. Rossbach-Lochmann, K. Gösele, E. Buchta (1986).
Schallschutz in Mehrfamilienhäusern aus der Sicht der Bewohner (Sound
insulation in multi-family houses from the point of view of the resident, in
German). F 2049, IRB Verlag, Stuttgart, Germany.
WHO (2007). Large Analysis and Review of European Housing and Health
Status (LARES). Preliminary overview. WHO Regional Office for Europe,
Copenhagen, Denmark.
S. Wibe (1997). Efterfrågan på tyst boende (The Demand for Silent Dwellings, in
Swedish). Anslagsrapport A4:1997. Byggforskningsrådet, Stockholm, Sweden.
Chapter 15
In the previous chapter, the question was which level of sound insulation
should be required to obtain satisfactory acoustical conditions for dwell-
ings in multi-unit houses. In this final chapter, we shall look at some exam-
ples of experimental buildings with particularly good sound insulation.
The examples are restricted to multi-storey buildings with dwellings. Of
course, the idea of making experimental buildings with very good sound
insulation comes from the ambition that dwellings for the future should
have satisfactory conditions, also for acoustics.
In some cases, the purpose has been to demonstrate in full scale that it
is possible to give priority to sound insulation in the building design and
to achieve practical experience with the suggested technical solution. It is
remarkable that the very high acoustic quality achieved in these projects
does not necessarily lead to higher construction costs, and one of the solu-
tions was actually developed with the purpose of saving construction time.
15.1.1 General
The building was erected 1988–1989 in Kv. Bollen, Lund, Sweden. It has
three stories and 29 apartments. Builder: JM Byggnads och Fastighets
AB. Consultants: Sterner Akustik AB, Bentech Ingenörs AB. The project
was supported by The Swedish Council for Building Research and The
Development Fund of the Swedish Construction Industry.
Figure 15.1 shows a part of the façade of the experimental building.
15.1.2 Purpose of experiment
The project was initiated by Lars Holmgren, Cementa AB. The purpose
was to find out what could be achieved acoustically with a traditional
heavy-weight building construction. Well-known solutions and construc-
tion details were applied, but with an unusual focus on the acoustical needs.
425
426 Sound Insulation in Buildings
15.1.3 Building concept
The main constructions are 240 mm concrete walls and 290 mm concrete
slabs cast on site. The floors were either carpet or wood on cork. This is
type A in Figure 11.1. Flanking transmission is possible in all directions,
but of little importance due to the high weight of the constructions.
The project report mentions, as a drawback to the building concept, that
the time needed for the concrete to harden is longer than usual because of
the thickness of walls and slabs.
Experimental buildings with high sound insulation 427
15.1.4 Acoustical results
An overview of the measurement results is presented in Table 15.1.
The following abbreviations are used in Table 15.1: R′w,50 = R′w + C 50–5000,
L′n,w,50 = L′n,w + C I,50–2500.
Examples of measured airborne sound insulation in the horizontal and
vertical directions are shown in Figure 15.2. In both directions, the trans-
mission loss is very high; at frequencies above 250 Hz, the transmission loss
is better in the vertical direction than in the horizontal direction.
The measured impact sound insulation in the vertical direction is
shown in Figure 15.3. The results are very good and the special project
Table 15.1 Single-number results of measured sound insulation in the ‘Quiet House’
Airborne, Airborne, Impact, vertical
Frequency range (Hz) horizontal (dB) vertical (dB) (dB)
100–3150 R′w = 62 R′w = 65 L′n,w = 43
50–5000 R′w,50 = 61 R′w,50 = 61 L′n,w,50 = 46
Note: Same as in Figures 15.2 and 15.3.
100
90
80
Transmission loss, R΄ (dB)
70
60
50
40
30
20
50 100 200 500 1000 2000 5000
Frequency (Hz)
Minimum requirement Project requirement Horizontal Vertical
Figure 15.2 Measured airborne sound insulation in the horizontal and vertical directions.
The full line is the project requirements in the extended frequency range. The
dashed line is the normal minimum requirement at the time of construction.
428 Sound Insulation in Buildings
80
70
Impact sound pressure level, L΄n (dB )
60
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Minimum requirement Project requirement Measured vertical
Figure 15.3 Measured impact sound insulation in the vertical direction. The full line is
the project requirements in the extended frequency range. The dashed line
is the normal minimum requirement at the time of construction.
requirements are fulfilled (recalling that the reference curve may be passed
with up to 32 dB of unfavourable deviations).
15.2.1 General
The ‘OPEN HOUSE Urban Building System’ has been developed by
Dominia in cooperation with the architect firm Landskronagruppen,
Landskrona, Sweden.
The building concept has been used in several places in Denmark. The
experimental buildings described here were erected 1993 and are located at
Færgeparken III, Frederikssund, Denmark. There are three 3-storey build-
ings; in total, 36 dwellings. Architect: Mangor & Nagel, Frederikssund.
Builder: Arbejdernes Kooperative Byggeforening (a cooperative building
association). Consultant: Dominia, Copenhagen.
Experimental buildings with high sound insulation 429
15.2.2 Purpose of experiment
The purpose was to gather experience with the newly developed building
system, which is a light-weight system with short building time. Improved
sound insulation was not a purpose of the experiment.
15.2.3 Building concept
The ‘OPEN HOUSE Urban Building System’ is a building system based
on factory-produced room modules and a site-produced concrete skeleton.
Each room module consists of a steel framework enclosing one or more
rooms. Walls, floors, and ceilings consist of gypsum boards mounted
on steel ribs. Compared to conventional concrete housing, the weight is
approximately 35 %, the building time at site is 25 % and the building time,
including production of modules, is 70 %. Different stages in the building
process are shown in Figure 15.4.
The construction principle is a column-beam skeleton with light-
weight walls and floors, type D in Figure 11.1. The concrete columns
are cast on site, while the concrete beams are prefabricated (Figures 15.5
and 15.6).
Figure 15.4 Schematic view of different stages in the building process. (From Dominia.
With permission from Dominia, Copenhagen, Denmark.)
430 Sound Insulation in Buildings
Concrete
Ste
el ske
leto
Corner n
profile
Figure 15.5 Detail of a junction. The steel profiles at the edges of the prefabricated
room modules create a form for in situ casting of the concrete columns. The
prefabricated concrete beams are fixed in the steel profiles. (With permis-
sion from Dominia, Copenhagen, Denmark.)
15.2.4 Acoustical results
An overview of the measurement results is presented in Table 15.2. The
measured sound insulation results are 8 dB to 10 dB better than the mini-
mum requirements in Denmark at the time of building. Measurements did
not include frequencies below 100 Hz.
The measured airborne sound insulation is shown in Figure 15.7.
Above 200 Hz, the transmission loss in the horizontal direction is higher
than that in the vertical direction, and the frequency course is typical
for a double wall, decreasing below 200 Hz towards a mass-spring-mass
resonance frequency that may be just below 100 Hz. Although no mea-
surement results are available below 100 Hz, it seems that the airborne
sound insulation at low frequencies may be an issue, both vertically and
horizontally.
300
Acoustic sealing
Gypsum board
30 39 30
Sealing with
242
252
mineral wool 50 50
10 6 10
39
6 mm neoprene
9 mm gypsum 9 mm gypsum
The measured impact sound insulation is shown in Figure 15.8. The fre-
quency course suggests that there may be an issue with the impact sound
below 100 Hz. Otherwise, the results are quite satisfactory.
15.3 ‘ECO-HOUSE’, DENMARK
15.3.1 General
The ECO-House building system was developed in connection with
a competition from 1987 on the question: ‘How can Danish building
industry make the next great jump forward?’ Due to the falling price
on steel, it was considered advantageous to develop an alternative to the
432 Sound Insulation in Buildings
100
90
80
Transmission loss, R΄ (dB)
70
60
50
40
30
20
50 100 200 500 1000 2000 5000
Frequency (Hz)
Horizontal Vertical
15.3.2 Purpose of experiment
The basic idea was to maximize prefabrication in order to minimize the
building time without compromising economy or quality. Acoustically,
there was no ambition to exceed the Danish minimum requirements for
sound insulation at the time of construction.
15.3.3 Building concept
The supporting structure is a framework of steel profiles. This is type D in
Figure 11.1. Floor constructions and walls are prefabricated elements of gyp-
sum board and mineral wool in thin-plate steel profiles. On the building site,
Experimental buildings with high sound insulation 433
80
70
Impact sound pressure level, L΄n (dB)
60
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Measured vertical
the steel framework is erected first. Then, the floor and façade elements and
the roof are mounted. Building materials needed for the internal parts are
brought into the building before the raw house is closed. The rest of the build-
ing process can take place in a dry environment protected from rain and wind.
The total building time is 70 % of that for more traditional building systems.
A construction detail is seen in Figure 15.9. The steel hat profile is a
structural beam that also provides support for the floor elements.
15.3.4 Acoustical results
An overview of the measurement results is presented in Table 15.3. The
average of measured sound insulation results is 5 dB to 10 dB better than the
minimum requirements in Denmark at the time of building. The measured
300
150 150
13 13 248 13 13
2 mm × 13 mm 45 mm 2 mm × 13 mm gypsum
gypsum fixed
mineral
wool
Bottom profile
SKF 70/100 with felt 16 mm wood on felt
3 mm × 13 mm gypsum
20 mm trapez plate
Element
25 162 25
Neoprene
2 mm × 15 mm special gypsum
50 mm × 50 mm angle profile
Figure 15.9 Detail of the junction between floor and partition wall in the ‘ECO-HOUSE’.
15.4.1 General
The building was erected in 1994 in BRF Kajplatsen, Kv Trålen, Stockholm,
Sweden. It has seven stories and 26 apartments. The experimental building
Experimental buildings with high sound insulation 435
80
70
Impact sound pressure level, L΄n (dB)
60
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Measured vertical
15.4.2 Purpose of experiment
Ljunggren (1993, 1995) has shown theoretically that it should be possible
to increase the sound insulation in the vertical direction significantly by
using a column-plate structure with a large slab supported by load-bearing
columns instead of load-bearing walls (type B in Figure 11.1). The idea
was that the flanking transmission through the slab could be turned into
an advantage by distributing the energy of structural vibrations in a plate
with a large area. This implies that the walls should be light-weight double
constructions so that the junction attenuation would be negligible. The
principle is explained by Statistical Energy Analysis (SEA) in Section 7.6.
436 Sound Insulation in Buildings
1994
Figure 15.11 External view of two buildings, one of them being the experimental building
on ‘BRF Kajplatsen’.
The purpose of the experimental building was to prove that the area effect
could work in practice. The building costs are estimated not to exceed
those of a conventional building by more than 1 %.
The theory of this project was developed by Professor Sten Ljunggren
in research projects supported by The Swedish Council for Building
Research and The Development Fund of the Swedish Construction
Industry.
15.4.3 Building concept
Three similar buildings were erected, two of them with traditional load-
bearing walls of 180 mm concrete and 240 mm thick concrete slabs. The
experimental building has 260 mm thick concrete slabs supported by
100 mm × 100 mm steel columns in the façade and 400 mm × 400 mm con-
crete columns hidden in the walls inside the apartments (Figure 15.12). The
double walls used as partition walls between the apartments are shown in
Figure 15.13. All buildings have concrete walls around the stairwell and
elevator in the middle of the building. The experimental building is type B
in Figure 11.1, while the other buildings are type A.
The acoustical principle is illustrated in Figure 15.14. The total area of
the concrete slab is 300 m 2 . The area of the floor in one on the bedrooms
Experimental buildings with high sound insulation 437
Figure 15.12 Plan of the experimental building on ‘BRF Kajplatsen’. The small black
squares are the columns, six from concrete inside the building and 25 from
steel in the façade. (Reproduced from Ljunggren, S., wksb, Zeitschrift für
Wärmeschutz, Kälteschutz, Schallschutz, Brandschutz, Heft 38, 1–4, 1996.
With permission.)
15.4.4 Acoustical results
An overview of the measurement results is presented in Table 15.4.
The measured airborne sound insulation between two rooms in the ver-
tical direction is shown in Figure 15.15, both for the experimental build-
ing and the traditional building. The difference is around 8 dB to 10 dB
in the entire frequency range. The thickness of the slab is not the same
in the experimental building and the traditional building, but this cannot
account for more than 1 dB difference. So, the efficiency of the principle in
the experimental b uilding is clearly demonstrated. In the horizontal direc-
tion, the airborne sound insulation is also very high. At the low frequencies
50 Hz to 80 Hz, the average transmission loss is around 35 dB, which is
quite good for a light-weight double wall.
438 Sound Insulation in Buildings
Independent steel
studs
3 mm × 70 mm
mineral wool
3 mm × 13 mm
gypsum board
Sealing
Figure 15.13 Section of the double walls used between the apartments on ‘BRF
Kajplatsen’. (Reproduced from Ljunggren, S., wksb, Zeitschrift für
Wärmeschutz, Kälteschutz, Schallschutz, Brandschutz, Heft 38, 1–4, 1996.
With permission.)
Figure 15.14 Principle of the area effect. The area of the slab is much larger than the
area separating the source room and receiver room. (Reproduced from
Ljunggren, S., wksb, Zeitschrift für Wärmeschutz, Kälteschutz, Schallschutz,
Brandschutz, Heft 38, 1–4, 1996. With permission.)
80
70
60
Transmission loss, R´ (dB)
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Experimental Traditional Avg. meas.
80
70
Impact sound pressure level, L΄n (dB)
60
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Experimental Traditional
15.5.1 G eneral
Three apartment houses were erected in 1996 on the occasion of a Finnish
Housing Fair in Ylöjärvi. The houses have three stories and 20 apartments
in total. The building concept was developed by Tampere University of
Technology on the initiative of Skanska Oy and Finnforest Oy as part of
the Nordic Wood Project (see next Section 15.6).
15.5.2 Purpose of experiment
The purpose was to find a method to improve the sound insulation in
wooden apartment buildings by reduction of flanking transmission. The
problem to be solved was the development of a resilient structural joint
with which the non-bearing walls could be joined to the bearing beam-to-
column frames.
15.5.3 Building concept
The building concept is a column-beam structure made from laminated
veneer lumber (LVL) wood. This is type D in Figure 11.1. The prefabri-
cated ribbed slab elements are made from 51 mm thick LVL with 299 mm
high ribs having a spacing of 300 mm. The slabs are 1800 mm wide and
can have a possible span of 7 m. Details of the floor and the junctions are
shown in Figure 15.17. More information is found in the work of Keronen
and Kylliäinen (1997).
15.5.4 Acoustical results
The main problem in the research and development project was finding
a solution with sufficient impact sound insulation at low frequencies, i.e.
below 100 Hz. Impact sound pressure levels were measured from 50 Hz
to 3150 Hz, and in addition to the ISO 717-2 reference curve, Bodlund’s
Experimental buildings with high sound insulation 441
51
299
Mineral wool 15 mm – Floor surface
– Mineral wool 30 mm
Acoustic steel Desity 30 kg/m3 (Density 100 kg/m3)
– Ribbed slab 299 mm,
Profiles 25 mm mineral wool between
c/c 400 mm ribs (density 30 kg/m3)
– Steel profiles 25 mm
– Plasterboard 13 mm
225 – Plasterboard 13 mm 150
277
Figure 15.17 Construction details of the floor and the junctions to wall and façade. The
building in Ylöjärvi has two kinds of floor surface: vinyl on 3 mm× 13 mm
plasterboard (30.8 kg/m2) or birch boarding on plastic foam layer on
2 mm× 13 mm plasterboard (18.8 kg/m2). (Reproduced from Keronen, A.,
Kylliäinen, M., Sound Insulation Structures of Beam-to-Column Framed Wooden
Apartment Buildings. Publication 77 Structural Engineering, Tampere
University of Technology, Finland, 1997. With permission.)
80
70
60
Transmission loss, R΄ (dB )
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Horizontal Vertical, birch Vertical, vinyl
Figure 15.18 Measured airborne sound insulation in the horizontal direction and in the
vertical direction both with vinyl and wooden flooring in ‘Ylöjärvi’ apart-
ments. (Adapted from Keronen, A., Kylliäinen, M., Sound Insulation Structures
of Beam-to-Column Framed Wooden Apartment Buildings. Publication 77
Structural Engineering, Tampere University of Technology, Finland, 1997.)
80
70
60
Impact sound pressure level, L΄n, dB
50
40
30
20
10
0
50 100 200 500 1000 2000 5000
Frequency (Hz)
Birch boarding floor Vinyl floor covering
Figure 15.19 Measured impact sound insulation in the vertical direction with two dif-
ferent floorings, vinyl or wood in ‘Ylöjärvi’ apartments. (Adapted from
Keronen, A., Kylliäinen, M., Sound Insulation Structures of Beam-to-Column
Framed Wooden Apartment Buildings. Publication 77 Structural Engineering,
Tampere University of Technology, Finland, 1997.)
Experimental buildings with high sound insulation 443
15.6.1 General
In the years 1993–1997, the so-called Nordic Wood project has lead to a
series of experimental wooden buildings being developed and built in the
Nordic countries (Nordic Wood, 1997; Hveem et al., 2000).
15.6.2 Purpose of experiment
The emphasis was on wood as an attractive building material. However,
the two obvious challenges in wooden multi-storey houses were fire protec-
tion and sound insulation. Thus, the purpose was to initiate research and
give support to experimental buildings that could provide possible solu-
tions to these problems.
15.6.3 Acoustical results
The greatest acoustical challenge in multi-storey wooden houses is the
impact sound at low frequencies. An overview of measurement results
of impact and airborne sound insulation from some of the experimental
buildings is found in Table 15.6. All impact sound insulation measure-
ments include frequencies down to 50 Hz, and in addition to the ISO 717-2
reference curve, Bodlund’s reference curve was applied (see the explana-
tion in Section 5.8.3).
REFERENCES
A wave, 8–9
wave equation, 10, 205
AAD (aggregated adverse deviation), Bias error, 368
146 Bodlund-weighted impact
Absorption area, 82 sound pressure level, 144–145
Acceleration level, 2 Boundary absorption in a plate,
Accelerometer, 45, 353 see Structural absorption
Acoustical comfort, 400 coefficient
Acoustically thick plate, 216, 218 Boundary connections, 259
Airborne sound insulation Box-in-box system, 41
dose–response function, 413 Building materials, 13
measurement, 340–347 Bulk modulus, 4, 9
measurement uncertainty, 388–389
Airborne sound reduction index Ia, 145
Angular frequency, 3 C
Angular wave number, 4
Aperture, 104–106 Cardinale sine, 15
Apparent sound reduction index, see Cavity mode, 377
Transmission loss, apparent Centre frequency, 4
Area effect, 438 exact, 5
Averaging time, 337, 356, 380–381 nominal, 5
Averaging Characteristic dimension of a plate,
arithmetic, 14 167
energetic, 14 Combination of airborne and impact
geometric, 14 transmission, 321
harmonic, 14 Confidence level, 364
Axial modes, 70, 376–378 Core materials, 244
Corner position of microphone,
B 345
Correlation coefficient, 381
Background noise, 337 Coupling factor, 263
Bandwidth Coupling loss factor, 191
effective, 368 Coverage factor, 364
half-power, 32 Crest factor, 340
relative, 5 Criteria for sound insulation,
resonant system, 31 419–420
Bending Critical damping, 23
stiffness, 10 Critical frequency, 12, 112
445
446 Index
E Harmonic vibration, 3, 20
Hearing threshold, 7
Elastic connections, 262, 326–328 Heavyweight wall, 223
Element method, 347
Enclosures, 135 I
Energy balance, in a resonant system,
189 IIC (impact insulation class), 143
Energy balance, in a room, 82 Imaginary unit, 3
Energy density, 51–52, 54, 56 Impact sound
Equal loudness contour, 6 attenuation, 138, 280–297, 302
Equivalent absorption length, 325, 360 complaints, 305
Equivalent number of uncorrelated dose–response function, 415
positions, 383–385 heavyweight floors, 307–309
Excitation, acoustic, 198 index Ii, 145
External transmission loss, 226 insulation, 275
lightweight floors, 306–308
F measurement of, 348–349
Impact sound pressure level, 136, 275
Flanking transmission, 131, 313–332 Impedance
airborne sound, 313–317 characteristic, 107
building types, 314 hammer, 285
impact sound, 318–320 mechanical, 21, 27, 262
Floor covering, 280 point, 21, 276
Index 447
Q S
Q-factor, 37, 308, 310 Sabine’s formula, 83–85, 103
Sandwich plate, 234
R forced transmission, 240
resonant transmission, 238
Radiation efficiency, 163 Schroeder’s limiting frequency, 79,
Radiation efficiency, measurement of, 367, 370
358 SEA (statistical energy analysis),
Radiation loss factor, 210 189–201
Rayleigh’s method, 164 Shear modulus, 10
Reciprocity principle, 193, 227, Shear wave, 9
317–318 Sigmoid functions, 405
Reference curve, airborne sound Sinc, see Cardinal sine
insulation, 141 Sine-sweep, 96, 340, 342
Reference curve, impact sound, Single leaf construction, 106
144 Single-number quantity, measurement
Reflection uncertainty, 389
density, 92 Single-number rating, 140
influence of bandwidth, 61 Sleep disturbance, 396
normal incidence, 49 Slits, 128–131
oblique incidence, 52 SNR (signal-to-noise ratio), 341
random incidence, 54 Socio-acoustic surveys, 408–410
Refraction effect, 184 Sound absorption in air, 91
Requirement for sound insulation, 150, Sound bridges, double walls,
390, 419 254–265
Resonance, 30 Sound bridges, floating floors, 296
Resonance frequency, 30, 42 Sound insulation,
floating floor, 288, 291 airborne, 101
floor covering, 280 homogeneous, isotropic plate,
mass-air-mass, 123, 249 203
suspended ceiling, 297 impact, 136
Resonance peak, floating floor, 308 low frequencies, 221
Resonant system, 189 oblique incidence, 226
Resonant transmission, see Sound subdivided structures, 224
transmission, resonant thick plates, 218–221
Resonant vibration, see Vibration, thin plates, 216–218
resonant Sound power level, 2, 85
Reverberation distance, 85–88 Sound pressure level, 2
Reverberation formula, general, 89 A-weighted, 6
Reverberation time, 26, 83 measurement of, 335–339
calculation, 93 measurement uncertainty,
correction for, 339 374–381
evaluation range, 97 Sound propagation in plates, 8
lower limits, 351 Sound radiation, 160–181
measurement of, 96–98, 350–353 directivity, 181–186
measurement uncertainty, forced vibrations, 166–173
385–387 near field, 278
modal, 72 resonant vibrations, 174–181
RMS (root-mean-square), 2 Sound reduction index, see
Room constant, 87 Transmission loss
Room modes, 70–74 Sound transmission coefficient, see
RR (risk ratio), 398 Transmission coefficient
Index 449