Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

An introduction to statistical energy analysis

Courtney B. Burroughs
Graduate Program in Acoustics, The Pennsylvania State University, State College, Pennsylvania 16804

Raymond W. Fischer
Noise Control Engineering, Incorporated, Billerica, Massachusetts 01821
Fred R. Kern
Atlantic Applied Research Corporation, Burlington, Massachusetts 01803

~Received 28 April 1993; revised 4 September 1996; accepted 20 November 1996!


The basic concepts and underlying assumptions of Statistical Energy Analysis ~SEA! are reviewed.
Using coupled simple oscillators as models for resonant modes, power balance equations are
derived. Equipartition of energy among modes within a subsystem, and strong versus weak coupling
between subsystems are discussed. Input requirements for exercising SEA power balance equations
are defined and methods of obtaining the required input are outlined. Examples of the applications
of SEA are presented and discussed. © 1997 Acoustical Society of America.
@S0001-4966~97!07404-3#
PACS numbers: 43.10.Ln, 43.40.At @DWM#

INTRODUCTION analysis band. By definition, subsystems are structural or


acoustical entities that have modes which are similar in na-
Designers are often required to comply with noise speci- ture and, as will be shown later, have equal modal energies.
fications in a complex system of acoustical and structural Different subsystems typically have different modal ener-
subsystems with several broadband sources of noise and vi- gies. The primary variable of interest in SEA is energy. For
bration. In such systems, noise problems need to be identi- steady-state conditions, a power balance is derived in each
fied early in the design stage so that treatments can be effec- analysis band in which the input power to the subsystem~s! is
tively integrated into the design. This requires the capability either dissipated within the subsystem~s! or coupled to other
to accurately predict vibration and noise propagation subsystems where it is dissipated or radiated to the acoustic
throughout a system which involves coupled subsystems. At farfield. The energy is distributed among the subsystems to
frequencies above the first few frequencies of resonance of establish a balance in the power flow into and between sub-
subsystems, it is difficult, if not impossible, to use classical systems such that the subsystem energies times their cou-
analytic models which involve modal expansions, or numeri- pling, damping, and radiation loss factors account for all of
cal methods which require large numbers of finite elements. the power input to the subsystems. From subsystem energies,
Often it is necessary to predict the average vibration and/or other parameters, such as average vibration and sound pres-
noise levels in various frequency bands. Thus models that sure levels, are obtained. Because equal distribution of en-
make predictions at single frequencies need to be exercised ergy over space within each subsystem is assumed, only av-
repeatedly to cover the range of frequencies in each fre- erage vibration and noise levels in each subsystem can be
quency band of interest. To address the need to predict av- obtained in each analysis band. Neither frequency distribu-
erage vibration or noise levels in frequency bands for com- tions of the levels within each band, nor spatial distributions
plex systems of inter-connected subsystems, Statistical of the levels of the modal energy within a subsystem can be
Energy Analysis ~SEA! methods were developed. Since its obtained from SEA. In most situations with multiple noise
early conception, SEA has been widely used, primarily in the sources and paths, average levels are all that are required.
aerospace industry. The purpose of this article is to outline This makes SEA an ideal tool that provides the required
the basic principles behind SEA, review the development of information with a minimum level of modeling complexity
SEA, present examples of its application, and examine input and detail. Its relatively simply equations and outputs can
parameters as an introduction to SEA. This article can not provide insight in developing an understanding of the impor-
satisfy all of the needs for the development of comprehen- tant paths in designing required noise control treatments.
sive models for the application of SEA to complex systems. The mathematical basics and assumptions for SEA are
The reader seeking more detailed information is referred to outlined in Sec. I. In Sec. II, methods for obtaining the pa-
the book by Lyon and DeJong.1 rameters required to exercise the power balance equations
SEA is statistical in that the frequencies of resonance for used in SEA for predicting the energy distribution within a
each subsystem are assumed to be uniformly distributed in system are reviewed. Examples of the application of SEA are
frequency within each of the frequency bands used in the presented in Sec. III, followed by a summary in Sec. IV.
analysis. Since it is assumed that energy resides only in reso-
nance modes, the total energy in each subsystem is the sum I. BASICS OF SEA
of the energies in the modes. This energy is assumed to be SEA is used to predict the response of a dynamic system
distributed equally among the modes in each subsystem and to external power input. A system will refer to the entire

1779 J. Acoust. Soc. Am. 101 (4), April 1997 0001-4966/97/101(4)/1779/11/$10.00 © 1997 Acoustical Society of America 1779

Redistribution subject to ASA license or copyright; see http://acousticalsociety.org/content/terms. Download to IP: 131.156.59.191 On: Tue, 09 Sep 2014 15:02:03
complement of coupled structures and acoustic spaces under
consideration. The system is then divided into subsystems
which consist of a collection of similar resonant modes
within a structure or acoustic space. For example, a bounded
acoustic space is usually treated as a subsystem containing
acoustic modes. Bending waves in a plate and longitudinal
waves in the same plate may each be treated as separate sets
of modes and therefore as separate subsystems. Once identi-
fied, subsystems are then coupled via coupling loss factors in FIG. 1. Model of coupled simple oscillators.

the development of power balance equations that serve as the


backbone of SEA. masses. Arms connecting the oscillator masses to the top of
The basic concepts of SEA are: the gyroscope slide in a fixed trough and are assumed to be
~1! Power flow between subsystems is proportional to the inflexible along their axes. The bottom of the gyroscope is
differences in the modal energies of the coupled sub- hinged in a ball socket. The displacement of the mass away
systems. This is supported by mathematical modeling, from the base of the oscillator is taken as positive in each
outlined below, that begins with the coupling of two oscillator. A positive velocity for oscillator 1 results in a
simple oscillators. The coupling of two oscillators be- positive force on oscillator 2, and a positive velocity for
comes the building block for coupling of multimode oscillator 2 results in a negative force on oscillator 1. Adding
resonant subsystems where each oscillator serves as a the coupling forces by the spring and gyroscope between the
oscillators to the equations of motion for the two oscillators
modal model.
yields
~2! Power input or transmitted to a subsystem is either dis-
sipated in the subsystem or transmitted to adjacent sub- m 1 j̈ 1 1r 1 j̇ 1 1k 1 j 1 1B j̇ 2 1s 12~ j 2 2 j 1 ! 5F 1 ,
systems via junctions of structures or interfaces between ~1!
structures and acoustic spaces. Thus a complete account m 2 j̈ 2 1r 2 j̇ 2 1k 2 j 2 2B j̇ 1 1s 12~ j 1 2 j 2 ! 5F 2 ,
of all of the power is taken.
where m 1 and m 2 are the masses of the oscillators, r 1 and
~3! Energy resides only in resonance modes, so that the
r 2 are the viscous damping coefficients, k 1 and k 2 are the
more modes a subsystem has in the analysis band, the
oscillator spring constants, B is the gyroscopic constant,
greater the capacity of the subsystem to accept and store
s 12 is the coupling spring constant, and F 1 and F 2 are the
energy. Within each analysis band, the energy in a sub-
magnitudes of the harmonic forces applied to the oscillator
system is uniformly divided among the modes. The net
masses. Grouping terms in Eq. ~1! and defining s 1 5k 1
power coupled between subsystems is proportional to the 2 s 12 and s 2 5 k 2 2 s 12 , gives
difference in their modal energies and only passes from
the subsystem with higher modal energy to those of m 1 j̈ 1 1r 1 j̇ 1 1s 1 j 1 1B j̇ 2 1s 12j 2 5F 1 ,
lower modal energy. When modes have equal energy, it ~2!
is referred to as an equipartition of energy. m 2 j̈ 2 1r 2 j̇ 2 1s 2 j 2 2B j̇ 1 1s 12j 1 5F 2 .

There are several assumptions that are generally made in These are the equations given by Cremer et al. ~p. 478!.2
the development of SEA models. Therefore, we can now follow Cremer et al.
With harmonic time dependence, e i v t , of the applied
~a! The input power spectra is broadband, i.e. there are no forces, Eqs. ~2! reduce to algebraic equations which can be
strong pure tones in the input power spectra. solved for the velocities of the oscillator masses:
~b! Energy is not created in the couplings between sub-
systems. Energy may be dissipated in junctions be- 2i v
n 1 5i vj 1 5 $ m 2 ~ v 2 22i v d 2 2 v 22 ! F 1
tween subsystems, such as in isolation mounts, but m 1m 2D~ v !
generally this effect is added to subsystem damping
loss factors. 1 ~ i v B1s 12! F 2 % ,
~c! The damping loss factor is equal for all modes within a ~3!
2i v
subsystem and analysis band. n 2 5i vj 2 5 $ ~ 2i v B1s 12! F 1 1m 1 ~ v 2
m 1m 2D~ v !
~d! Modes within a subsystem do not interact except to
share an equipartition of energy. The coherent effects 22i v d 1 2 v 21 )F 2 % ,
between modes are ignored so that power sums apply.
where

S
A. Coupled simple oscillator
Two simple oscillators are shown in Fig. 1 coupled via a D ~ v ! 5 v 4 22i v 3 ~ d 1 1 d 2 ! 2 v 2 v 21 1 v 22 14 d 1 d 2
spring and gyroscope. The coupling forces transmitted
through the spring are proportional to the differences in the
displacements, j1 and j2 , of the two masses in the oscillators.
Coupling forces transmitted through the gyroscope are pro-
1
B2
m 1m 2 D
12i v ~ d 1 v 22 1 d 2 v 21 ! 2
s 12
m 1m 2

portional to the velocities, j̇ 1 and j̇ 2 , of the two oscillator 1 v 21 v 22 . ~4!

1780 J. Acoust. Soc. Am., Vol. 101, No. 4, April 1997 Burroughs et al.: Energy analysis 1780

Redistribution subject to ASA license or copyright; see http://acousticalsociety.org/content/terms. Download to IP: 131.156.59.191 On: Tue, 09 Sep 2014 15:02:03
In Eqs. ~3! and ~4!, v i 5 (s i /m i ) 1/2 are the natural frequen- For oscillator 2, the energy is given by Eq. ~10! with the
cies of the uncoupled oscillators, and d i 5 r i /2m i are related indices interchanged. Taking the difference in the energies
to the half-power bandwidths of the oscillators about their gives

F G
natural frequencies. Of interest is the power flow between the
p u F 1u 2 u F 2u 2
oscillators, which can be described as W̄ 1 2W̄ 2 5 2
2D v m 1 d 1 m 2 d 2

F G
P 125 21 Re$ F 12n *
2 %, ~5!
1
2 ~ v 21 2 v 22 ! 2 1 ~ d 1 1 d 2 !~ d 1 v 22 1 d 2 v 21 !
where F 12 is the force applied by oscillator 1 to oscillator 2, 3 ,
given by the last two terms in the second part of Eqs. ~2!, and Q
the star ~*! denotes the complex conjugate. Combined with ~11!
the velocity given in Eqs. ~3!, this yields
where the second term in Eq. ~10! drops out because it is
v 2 s 2121 v 4 B 2 symmetric with respect to the two indices. Dividing Eq. ~11!
P 125 ~ m 2 d 2 u F 1 u 2 2m 1 d 1 u F 2 u 2 ! . ~6! into Eq. ~7! yields
m 21 m 22 u D ~ v ! u 2
This is the power flow at the single frequency, v. For broad- P 125 b ~ W̄ 1 2W̄ 2 ! , ~12!
band excitation, the total power flow given by Eq. ~ 6! must where
be integrated over the frequency bandwidth, Dv, of interest.
If we assume that the natural frequencies of the oscillators 2 ~ d 1 v 22 1 d 2 v 21 ! B 2 1 ~ d 1 1 d 2 ! s 212
b5 .
are far from both limits of the frequency band, then the in- m 1 m 2 ~ v 21 2 v 22 ! 2 12 ~ d 1 1 d 2 !~ d 1 v 22 1 d 2 v 21 !
tegration can be extended to 6` without significant loss in ~13!
accuracy. The integral can be easily evaluated using integral
Equation ~13! is a function of the parameters of the os-
tables @for example, see Ref. 3, p. 218, Eq. 3.112.5#. The
cillators ~d1 , d2 , v1 , and v2! and coupling factors ~B and
result of the integration over frequency is
s 12! and not the energies in the oscillators. The reason we

P 125
p
2F
u F 1u 2 u F 2u 2
2D v d 1 m 1 d 2 m 2 G have gone into so much detail in the development of Eq. ~12!
is because it forms the basis for some of the underlying prin-
ciples of SEA and its derivation illustrates some of the fun-
~ d 1 v 22 1 d 2 v 21 ! B 2 1 ~ d 1 1 d 2 ! s 212 damental assumptions behind SEA. Equation ~12! underpins
3 , ~7!
m 1m 2Q SEA. It shows that power flow between two coupled oscil-
lators, excited by broadband uncorrelated sources, is:
where
~1! Proportional to the differences in the energies in the os-
Q5 ~ v 21 2 v 22 ! 2 14 ~ d 1 1 d 2 !~ d 1 v 22 1 d 2 v 21 ! 1 ~ d 1 1 d 2 ! cillators or, as seen later, the modal energies for multi-

3 S v 21
d2
1
v 22
d1 D B2
m 1m 2
1
~ d 1 1 d 2 ! 2 s 212
d 1d 2m 1m 2
. ~8!
modal subsystems;
~2! The constant relating the power flow to the energy dif-
ferences is a function of the coupling parameters and
In the derivation of Eq. ~7!, it was assumed that the input properties of the oscillators;
forces are independent of frequency, i.e. the excitations are ~3! The direction of power flow is from the oscillator with
broadband with a flat frequency spectra. Next, we would like the higher ~modal! energy to the oscillator with the lower
to relate the power flow between the oscillators, given in Eq. ~modal! energy.
~7!, to the differences in the energies in the two oscillators. Assume that only oscillator 1 is excited by a broadband
The energy, W̄ i , in ith oscillator in Dv bandwidth is given source. Then, the net power flow from oscillator 1 to oscil-
by lator 2 must be dissipated in oscillator 2 to achieve a power

W̄ i 5
mi
2D v
E 2`
1`
n i~ v ! n *
i ~ v !dv. ~9!
balance. Thus
P 125 b ~ W̄ 1 2W̄ 2 ! 5 P 2d . ~14!
In Eq. ~9!, we have made the same assumption on the limits Since
of the bandwidth, relative to the natural frequencies of the
oscillators, made in deriving Eq. ~7! for the power flow. The W̄ i 5 21 m i u n i u 2 , P 2d 5 21 r 2 u n 2 u 2 , ~15!
integral of the terms involving the product F 1 F 2 vanishes Eq. ~14! becomes
when we assume that the input forces are independent. For
oscillator 1, we obtain m 2u n 2u 2 1
5 . ~16!

W̄ 1 5
p
2D v
H F u F 1u
2

m 1d 1
1
2 ~ v 21 2 v 22 ! 1 ~ d 1 1 d 2 !~ d 1 v 22 1 d 2 v 21 !
Q
G m 1u n 1u 2 11r 2 /m 2 b

Since r 2 /m 2 b . 0, Eq. ~16! shows that the energy in the

1
@ s 212~ d 1 1 d 2 ! 1B ~ d 1 v 22 1 d 2 v 21 !#@ m 2 u F 1 u 2 1m 1 u F 2 u 2 #
2m 21 m 22 d 1 d 2 Q
J .
undriven oscillator 2 can not be greater than the energy in the
driven oscillator 1. For small damping in oscillator 2 and
strong coupling between the oscillators, i.e. r 2 /m 2 b ! 1, Eq.
~10! ~16! becomes

1781 J. Acoust. Soc. Am., Vol. 101, No. 4, April 1997 Burroughs et al.: Energy analysis 1781

Redistribution subject to ASA license or copyright; see http://acousticalsociety.org/content/terms. Download to IP: 131.156.59.191 On: Tue, 09 Sep 2014 15:02:03
m 2 u n 2 u 2 5m 1 u n 1 u 2 . ~17! N 1 h 12
h 215 , ~22b!
N2
The energy is equally divided between the two oscillators.
The oscillators are said to be strongly or tightly coupled, and, where v is the center frequency of the analysis band, then
as Eq. ~17! shows, we have equipartition of energy between Eq. ~21! can be written as
the oscillators when they are strongly coupled. Power can be
viewed as passing back and forth between the oscillators to a P 125 v @ h 12W 1 2 h 21W 2 # , ~23!
much greater extent than it is dissipated in the oscillators. where W 1 5N 1 W m 1 and W 2 5N 2 W m 2 are the total energies
For weak coupling and large damping in oscillator 2, in subsystems 1 and 2, respectively, and h12 and h21 are the
r 2 /m 2 b @ 1, and Eq. ~16! becomes coupling loss factors. The first term in Eq. ~23! is the power
m 2 u n 2 u 2 !m 1 u n 1 u 2 . ~18! lost by subsystem 1 to subsystem 2 and the second term in
Eq. ~23! is the power lost by subsystem 2 to subsystem 1.
Little of the power input to oscillator 1 gets into oscillator 2, With the modal densities in subsystems 1 and 2 given by
and that power that does get into oscillator 2 is quickly dis- n 1 5 N 1 /D v and n 2 5 N 2 /D v , respectively, we can write Eq.
sipated, so that there is little energy built up in oscillator 2. ~22b! as
n 1 D v h 12 n 1
h 215 5 h . ~24!
n 2D v n 2 12
B. Coupled multiresonant subsystems
Thus the coupling loss factor for subsystem 2 to subsystem 1
Now consider two coupled subsystems. In a given analy-
is greater than for subsystem 1 to 2 when the modal density
sis band, subsystem 1 has N 1 modes and subsystem 2 has
in subsystem 1 is greater than the modal density in sub-
N 2 modes. The power flow from the m 1 th mode in sub-
system 2. The higher the modal density, the more modes
system 1 to the m 2 th mode in subsystem 2 is similar to that
there are to store energy within a fixed frequency band. Since
for two coupled oscillators, since each of the two modes act
modal densities are often easier to estimate than coupling
like a simple oscillator with inertial ~mass!, elastic ~spring!,
loss factors, Eq. ~24! can be used to reduce the effort re-
and dissipative ~dashpot! elements. Thus
quired in obtaining estimates for the coupling loss factors,
P m 1 m 2 5 b m 1 m 2 ~ W m 1 2W m 2 ! , ~19! since only one of the two coupling loss factors for two
coupled subsystems needs to be calculated.
where P m 1 m 2 is the power flow from the m 1 th mode in sub- The power input to subsystem 1 must either be transmit-
system 1 to the m 2 th mode in subsystem 2, b m 1 m 2 is the ted to subsystem 2 or dissipated in subsystem 1. That is
coupling parameter analogous to Eq. ~13!, and W m 1 (W m 2 ) is P 1input5 P 1d 1 P 12. ~25!
the modal energy in the m 1 th ~m 2 th! mode in subsystem 1
The power dissipated in subsystem 1 is given by
~subsystem 2!. The bars over the variables in Eq. ~19! denote
an average over a frequency band. If we assume that all of P 1d 5 v h 1 W 1 , ~26!
the modes in subsystem 1 are strongly coupled, then the
where h1 is the damping loss factor, which is the ratio of the
energy in subsystem 1 is equally divided among all of the
power dissipated per radian of motion to the total energy.
modes within the analysis band. Strongly coupled modes can
Using Eqs. ~23! and ~26! in Eq. ~25!,
be thought of as modes that are all directly responsive to the
same broadband excitation, such as the acoustic modes in a P 1input5 v $ ~ h 1 1 h 12! W 1 2 h 21W 2 % . ~27!
room responding to a broadband acoustic source in the room.
The energy is equally divided among the modes when the Likewise, for the power input to subsystem 2,
damping loss factor is the same for all of the modes. P 2input5 v $ ~ h 2 1 h 21! W 2 2 h 12W 1 % . ~28!
The coupling of the N 1 modes in subsystem 1 to the
m 2 th mode in subsystem 2 can then be expressed as Equations ~27! and ~28! are the power balance equations for
two subsystems. In Eqs. ~27! and ~28! the reason for defining
P 1m 2 5 ^ b m 1 m 2 & N 1 N 1 ~ W m 1 2W m 2 ! , ~20! the coupling loss factors as in Eq. ~22! can be seen. The
coupling loss factor serves an analogous role to that of the
where ^ & N 1 indicates an average over the N 1 modes in sub- damping loss factor. The damping loss factor is a measure of
system 1. Extending this to include the N 2 modes in sub- the rate at which energy is lost or dissipated within a sub-
system 2, we have system, whereas the coupling loss factor is a measure of the
rate at which energy in one subsystem is transmitted, and
P 125 ^ b m 1 m 2 & N 1 N 2 N 1 N 2 ~ W̄ 1 2W̄ 2 ! , ~21!
thereby lost, to another subsystem through their mutual cou-
where P 12 is the net power flow from subsystem 1 to sub- pling, when the energy in the other system is zero, i.e. it acts
system 2 in a frequency band. Defining as an energy sink.
If power is input only to subsystem 1, then Eq. ~28!
^ b m1m2& N1N2N 2 leads to
h 125 , ~22a!
v W2 h 12
5 . ~29!
and W 1 h 211 h 2

1782 J. Acoust. Soc. Am., Vol. 101, No. 4, April 1997 Burroughs et al.: Energy analysis 1782

Redistribution subject to ASA license or copyright; see http://acousticalsociety.org/content/terms. Download to IP: 131.156.59.191 On: Tue, 09 Sep 2014 15:02:03
This is a useful relation for evaluating treatments for the 1 v 2 F 20 f 2n ~ x 0 !
reduction of energy ~i.e., noise! in subsystem 2. Using Eq. n 25
M (n @~ v 2n 2 v 2 ! 2 1 h 2 v 4n # L n
, ~37!
~24! in Eq. ~29! yields
where we have applied orthogonality @Eq. ~33!#. For uni-
W m2 h 21 formly distributed point sources ~‘‘rain on the roof’’!, we
5 , ~30! integrate over source locations.
W m1 h 211 h 2

which indicates that for h21@h2 , the modal energies in sub-


systems 1 and 2 are nearly equal and the two subsystems
n 25
%
1 1
M S
E(
S n
v 2 F 20 f 2n ~ x 0 ! dx 0
@~ v 2n 2 v 2 ! 2 1 h 2 v 4n # L n
. ~38!

may be combined into one subsystem. Using Eq. ~33!


Equations ~27! and ~28! can be extended to N sub-
systems to obtain the set of N equations given by P. W. v 2 F 20 1
Smith Jr.:4
%
n 25
M2 (n ~ v 2n 2 v 2 ! 2 1 h 2 v 4n
. ~39!

P i,input5 v HS N

D N

h i 1 ( h i j W i 2 ( h ji W j ,
jÞi jÞi
J ~31!
Equation ~39! shows that, for independent localized sources
~or random broadband excitation!, the input energy to all
modes is uniform. Also, Eq. ~39! indicates that the response
for i 5 1,2,...,N. Equations ~31! are the power balance equa- of a subsystem to random broadband excitation is dominant
tions for N coupled subsystems that form the basis for SEA. at the resonance frequencies, v n .
The unknowns are the energies, W i , in the subsystems. The Averaging over a frequency band v22v15Dv, which
inputs are the coupling loss factors, h i j , the damping loss contains the modes n 1 to n 2 , we obtain
factors, h i , and the power inputs, P i,input. The power balance

E
n2
equations are solved for the energies in each frequency F D2 v 1 v2 v2 dv
analysis band. Each analysis band must be wide enough to n D2 v 5
M2 (
n5n 1 Dv v1 ~ v 2n 2 v 2 ! 2 1 h 2 v 4n
, ~40!
contain several modes in each subsystem.
Two key assumptions made in deriving the power bal- where F D2 v is the average of the forces within the frequency
ance equations were equipartition of energy and that reso- band. Assuming that v n . v , i.e., the range of resonance
nant modes act like simple oscillators. In support of the va- frequencies is small compared to the center frequency of the
lidity of these assumptions, we once again follow Cremer frequency band and that the damping loss factor is small so
et al.,2 and consider a finite subsystem ~e.g., a closed acous- that the half-power bandwidths about resonant peaks remain
tic space or a finite beam, plate, or shell! that has resonant well within the frequency band, then Eq. ~40! becomes
modes, f n (x), such that the response can be expressed in a N2
modal ~eigenfunction! expansion F D2 v p
n D2 v 5
M 2 (
D v n5N 1
2hvn
. ~41!

n~ x !5 (n n n f n~ x ! ~32!
Again, applying the assumption that v n . v for all n

where x denotes spatial location. The mode functions are F D2 v p DN


v D2 v ' . ~42!
orthogonal, i.e. M 2
2hv Dv

E s
m f n ~ x ! f m ~ x ! dx5 H Ln ,
0,
n5m,
nÞm,
~33!
For a simple oscillator,
ivF
n5 . ~43!
where L n is the normalization constant for the nth mode and m @ v 20 2 v 2 1i h v 20 #
m is the mass distribution. For a single localized source mod-
eled by Averaging Eq. ~43! over a frequency band containing the
natural frequency, v0 , yields
p ~ x ! 5F 0 d ~ x2x 0 ! , ~34!
F D2 v
p
we have n D2 v 5 . ~44!
m Dv 2hv0
2

i v F 0f n~ x 0 ! Comparing Eqs. ~42! and ~44! shows that, when a multireso-


n n5 , ~35!
@ v 2n ~ 11i h ! 2 v 2 # nant subsystem is excited by random broadband sources, the
modal response is similar to the response of a simple oscil-
where d~•! is the Dirac delta function. The spatial average of
lator and that the energy is equally distributed among the
the response is
modes, such that the subsystem energy in a frequency band

n 25
1
M
E s
m u n ~ x ! u 2 dx, ~36!
is directly proportional to the number of modes in the band.
This is the basis of Eqs. ~20! and ~21! where we simply
multiplied by the number of modes in each subsystem.
where M is the total mass of the subsystem. Using Eq. ~32! The power transmitted into a subsystem by a pressure
in Eq. ~35!, we obtain field, p(x), is given by

1783 J. Acoust. Soc. Am., Vol. 101, No. 4, April 1997 Burroughs et al.: Energy analysis 1783

Redistribution subject to ASA license or copyright; see http://acousticalsociety.org/content/terms. Download to IP: 131.156.59.191 On: Tue, 09 Sep 2014 15:02:03
P5
1
2
Re HE S
p ~ x ! n * ~ x ! dx ,J ~45!

where the star ~*! denotes the conjugate of the complex ve-
locity, n (x), which includes amplitude and phase, and Re~•!
denotes the real part of the complex quantity. Using the same
method used in deriving Eq. ~41! yields
F D v p DN
P Dv5 . ~46!
2 2M D v
This shows that the power transmitted into a subsystem is
proportional to the modal density.
Adding Eqs. ~27! and ~28! for a system with two
coupled subsystems yields
FIG. 2. Illustration of wave-number plot for estimating modal density of a
P 1,input1 P 2,input5 v $ h 1 W 1 1 h 2 W 2 % , ~47! simply-supported rectangular plate of dimensions L x and L y .

which shows that all of the power input to the system is


dissipated. This is a simple power balance and is true regard- tween two subsystems is strong when the ratio of the cou-
less of the strength of the coupling, so that pling to damping loss factors for either subsystem is large.4
h 1 W 1 1 h 2 W 2 5 h 1 W 81 1 h 2 W 82 , ~48!
where the primes denote the energies in the subsystems when II. REQUIRED INPUT FOR POWER BALANCE
the subsystems are strongly coupled. For strong coupling, the EQUATIONS
damping loss factors, h1 and h2 , are much smaller than the The inputs required for the power balance equations are
coupling loss factors, h12 and h21 , so that Eq. ~27! becomes damping and coupling loss factors, subsystem masses, input
P 1,input5 v ~ h 12W 81 2 h 12W 82 ! . ~49! powers, and modal densities. With these inputs, the power
balance equations become algebraic equations with known
In the limit as we increase the coupling loss factors while coefficients which can be solved for the unknown energies in
maintaining a constant input power, Eq. ~49! leads to each of the subsystems. The average vibration and sound
h 12W 81 5 h 21W 82 . ~50! pressures are then computed from predictions of the sub-
system energies. It is assumed that the input powers are
Using Eq. ~22b! known or can be computed based on force or displacement
W 81 W 82 inputs and subsystem admittances. Damping loss factors for
5 , ~51! untreated structures, where the dissipation is in the material
n1 n2
of the structure, must rely on measured data or experience
which is equipartition of energy, which applies to strongly for built-up structures. Mathematical expressions have been
coupled subsystems. developed for coupling loss factors, modal densities, and
From Eq. ~48!, damping loss factors for structures with added damping
treatments. However, space does not permit the discussion of
h2
W 1 5W 18 1 ~ W 28 2W 2 ! . ~52! expressions for the parameters for all types of acoustical
h1 spaces and structures. We will, therefore, restrict ourselves to
Using Eq. ~50! in Eq. ~52!, and the result in Eq. ~28! yields the consideration of rectangular flat plates, straight beams,
and rectangular acoustical spaces.
P 2,input / h 2 2W 82
W 2 2W 82 5 . ~53! A. Modal densities
h 12 / h 1 1 h 21 / h 2 11
Modal densities may be estimated from plots of wave
A similar equation can be derived for the energy in sub-
numbers of resonance in the form of a grid pattern as illus-
system 1, so that,4
trated in Fig. 2 for a simply-supported plate. Since the effect
P i,input / h i 2W i8 of boundary conditions reduces for higher-order modes, the
W i 2W 8i 5 , for i51,2, ~54! simply-supported plate can be a good model for plates for
11 g
purposes of estimating modal densities. The number of
where modes within a frequency band from v1 to v2 is then esti-
h 12 h 21 mated by counting the modes that fall in between the two
g5 1 ~55! circles with radii equal to the wave numbers for freely propa-
h1 h2
gating waves at the upper and lower frequencies of the band.
The transition from weak coupling, where g!1 and Mathematical expressions can also be obtained by dividing
W i 5 P i,input / h i , to strong coupling, where g@1 and the area inside the two circles by the area associated with
W i 5W i8 , can be seen in Eq. ~55! for two coupled sub- each mode. In Fig. 2, this area is p 2 /L x L y 5 p 2 /A, where
systems. Equations ~54! and ~55! show that coupling be- A is the area of the plate.

1784 J. Acoust. Soc. Am., Vol. 101, No. 4, April 1997 Burroughs et al.: Energy analysis 1784

Redistribution subject to ASA license or copyright; see http://acousticalsociety.org/content/terms. Download to IP: 131.156.59.191 On: Tue, 09 Sep 2014 15:02:03
With the lower frequency set to zero, an estimate of the the boundary of the first structure, it is possible to relate the
total number of modes, N, that exist below a given frequency coupling loss factor to the power transmission coefficient
~v! is obtained. The modal density is then ~t12! as follows,
]N~ v ! P trans v h 12W tot v h 12l f
n~ v !5 . ~56! t 125 5 5 , ~58!
]v P inc W tot~ c g /l f ! cg
Also, the mean-value theory5 may be used to estimate where P trans is the power transmitted through the junction,
modal densities from measurements of the drive-point admit- P inc is the power incident on the junction, l f is the mean-free
tance as a function of frequency. The mean value of the path length between incidences on the junction, and c g is the
measured drive-point admittance versus frequency (Y c ) is group velocity, i.e., the velocity of energy propagation. For
related to the modal density (n n ) as beams, c g 52c B where c B is the wave phase speed and
l f 5L, the length of the source beam, so that
pnn
Y c5 , ~57! 2c B
2M n h 125 t . ~59!
v L 12
where M n is the modal mass. For one-dimensional struc- For plates, l f 5 p A/L 12 , where L 12 is the length of the junc-
tures, the modal mass is approximately one-half the total tion, so that
mass for the higher-order modes, and for two-dimensional
structures, one-quarter the total mass. 2L 12
h 125 t , ~60!
p k p A 12
B. Damping loss factors where k p is the wave number for freely propagating bending
Damping loss factors for energy dissipated by the mate- waves in the source plate.
rial of the structure must rely on measurements or tables of For beams or plates, expressions for the power transmis-
representative damping loss factors. Typical ranges of damp- sibility can be obtained by matching the transverse and an-
ing loss factors are given in Figure 12.5 of Ref. 6. Damping gular velocities, and the shear forces and bending moments
treatments, such as free-layer and constrained-layer damping at the junction, or relating the transmissibility to the blocked
treatments, are often added to structures to reduce vibrations. forces generated by the incident wave, the impedances relat-
Estimates of damping loss factors for treated structures are ing the blocked forces to the velocities at the junction, and
given in Section III.6 of Ref. 3. the transmitted waves generated by the velocities at the junc-
tion. In plates, waves that comprise resonance modes are not
all normally incident at junctions nor are all junctions of
C. Coupling loss factors
right angles. With the inclusion of a large number of modes
Coupling between in-plane ~longitudinal and shear! and in an analysis frequency band, the angles of incidence may
bending waves occurs at all junctions of structures except be distributed over all possible incidence angles. In this case,
end-to-end junctions where there is no change in direction. expressions for the transmissibility as a function of incident
Longitudinal waves have low dissipative losses, even in the angle must be derived and averaged over all incident angles,
presence of most damping treatments, and have longer wave- which adds greatly to the complexity of the calculations and
lengths than bending waves. Thus longitudinal waves can results. Unfortunately, the additional complexities introduced
provide an effective means of transmitting energy in struc- by oblique incidence preclude simple expressions for the av-
tures over long distances. The energy in longitudinal waves erage transmissibility required for coupling loss factors for
can be coupled at junctions to bending waves which are plates coupled at right angles. Although the temptation is to
dominant sources of radiated noise. Therefore, both longitu- use normal incident transmissibilities to approximate aver-
dinal and bending waves need to be considered in the devel- aged transmissibilities, coincidences at incident angles,
opment of coupling loss factors for junctions of plates and where wave-number matching occurs, result in high trans-
beams to represent both multiwave type transmission and missibilities. Furthermore, cutoff angles above which trans-
radiation. missibility vanishes could result in the introduction of sig-
There are two approaches for deriving expressions for nificant errors with the use of normal incident
coupling loss factors for structures; the modal and wave ap- transmissibility in coupling loss factors for plate. An ex-
proaches. In the modal approach, the coupling between indi- ample of the transmissibilities for bending-to-bending,
vidual modes is computed and an average taken over the bending-to-longitudinal and bending-to-shear waves for L
modes in each frequency band. In the wave approach, the junctions of two plates of equal thickness2 is given by Fig. 3.
coupling loss factor can be related to the power transmissi- For the thicknesses of the plates in this example, the
bility for semi-infinite structures, which is often easier to bending-to-bending transmissibility peaks for an incidence
estimate than the average of the couplings between modes of angle of 31 deg. The cutoff angle for bending-to-longitudinal
finite structures. The power transmitted from the first to the wave transmissibility is 18 deg and for bending-to-shear
second structure through the junction is then energy lost by wave transmissibility, 28 deg. This example illustrates the
the first structure via the coupling. Since the coupling loss complexities of the coupling loss factors as a function of
factor has been defined as the energy lost per radian of mo- incidence angle and the need to include all angles of inci-
tion relative to the total energy in the structure, and the dence in computing coupling loss factor for junctions of
source of energy loss is transmission through the junction at structures.

1785 J. Acoust. Soc. Am., Vol. 101, No. 4, April 1997 Burroughs et al.: Energy analysis 1785

Redistribution subject to ASA license or copyright; see http://acousticalsociety.org/content/terms. Download to IP: 131.156.59.191 On: Tue, 09 Sep 2014 15:02:03
FIG. 3. Power transmissibilities for bending waves incident at a right angle FIG. 4. Comparison of SEA and finite element model predictions to mea-
L junction of two plates as a function of incidence angle. sured vibration levels for a six-plate system.

Because of the complexities involved in deriving mean- information becomes available and as the model is exercised
ingful expressions for coupling loss factors for structures, and compared to measurements of predictions for similar
coupling loss factors have been the subject of a number of systems.
studies and incorporated into SEA computer codes. For more Larger structures will contain more modes, allowing for
details, see Refs. 7–24. a more accurate analysis at lower frequencies. The results for
The coupling loss factor ( h sa ) for a plate radiating into smaller structures will only be valid when there are several
an acoustic space is related to the radiation efficiency of the modes present, which may be limited to mid- to high-
plate, srad , by frequency ranges.

r 0 c 0 S s rad
h sa 5 , ~61! III. EXAMPLES OF THE APPLICATIONS OF SEA
vMs
For a system of six coupled fat plates, Lu et al.27 con-
where S is the radiating area of the plate and M s the mass of ducted measurements on the broadband vibration levels in
the plate. Expressions for radiation efficiencies may be found 1-kHz bands. The measured vibration levels are compared in
in Ref. 25, Chapter 2 of Ref. 26, and Section 9.6 of Ref. 6. Fig. 4 to predictions made using a finite element model and
an SEA model. The configuration of the plates with the drive
and measurement locations are also shown in Fig. 4. A single
shaker was used to excite the system. At frequencies below 7
D. Assembling an SEA model
kHz, the predictions by both the finite element and SEA
The best approach to the definition of subsystems is to models are in good agreement with the measured levels. At 7
start with subsystems that extend to junctions that cause sig- kHz and above, the finite element predictions show poor
nificant reflections. Bear in mind that different wave types agreement with the measurements, while the SEA predic-
often require subsystems of differing physical size. For ex- tions continue to show good agreement. The limitation on
ample, since longitudinal wavelengths are larger than bend- the number of degrees of freedom in the finite element model
ing wavelengths, longitudinal subsystems may extend over a led to the poor agreement with the measured levels at 7 kHz
number of bending subsystems. If the coupling loss factors and above. The agreement shown in Fig. 4 illustrates the
between two subsystems are such that they are strongly validity of SEA broadband predictions at high frequencies
coupled in both directions, then these subsystems can be where finite element models are too costly to run. In finite
combined into one subsystem without loss of accuracy. element models, each plate must have a number of elements
Smaller models of the most complex part of the structure can at least four times the number of wavelengths of the plate in
sometimes aid in checking for strong coupling and simplifi- each direction while the number of SEA subsystems is given
cation of the overall model. by the number of wave types in the structural element, which
For a first pass, it is useful to hypothesize strong cou- is usually less than four.
pling. The relationship between the assumed damping and A scale model of a Diesel engine foundation and ship
coupling loss factors should be reviewed as the model is bottom structure was modeled with twelve plates by Tratch28
developed. The user not familiar with coupling loss factors in an SEA model. SEA predictions were generated by Tratch
should consider bounding these factors as a starting point for using bending waves only, and using both bending and lon-
the power balance equations, and refining them as additional gitudinal waves in the plates and at the junctions. An ex-

1786 J. Acoust. Soc. Am., Vol. 101, No. 4, April 1997 Burroughs et al.: Energy analysis 1786

Redistribution subject to ASA license or copyright; see http://acousticalsociety.org/content/terms. Download to IP: 131.156.59.191 On: Tue, 09 Sep 2014 15:02:03
FIG. 5. Comparison of SEA predictions for a 12-plate system, with and
without longitudinal waves, to measured vibration levels. FIG. 6. Comparisons of SEA predictions with thin and thick plate theory
coupling loss factors to measurements.

ample of comparisons between the SEA predictions and coupling loss factors for the HAYES were based on analytic
measured levels presented by Tratch28 and Lyon29 is given in models.
Fig. 5, along with a sketch of the SEA 12-plate model. At Results from an SEA model were presented by
frequencies above 2 kHz, the SEA predictions made without Burkewitz32 for the propagation of sonar ping noise in the
including longitudinal waves in the plates show poor agree- forward third of the USS GLOVER ~FF 1098!. The concern
ment with the measured levels. As shown in Fig. 5, including was the prediction of airborne noise in ship spaces. A total of
longitudinal waves in the SEA model improved the predic- 425 structural subsystems and 1513 junctions were included
tions to the point where good agreement of the predictions in the model. The first set of predictions, in which only bend-
with the measured levels was obtained up to 20 kHz. Thus ing waves in the structures were included, produced unsatis-
accurate SEA predictions require the inclusion of longitudi- factory agreement with the measured levels for the noise in
nal waves, particularly when the system has several struc- the ship spaces generated by sonar pings. The inclusion of
tural elements and predictions are made at high frequencies. longitudinal waves improved agreement between the SEA
For additional discussion of this model, see Chapter 15 of predictions and measured levels. Figure 8 shows a compari-
Ref. 1. For thick plates at high frequencies, it also becomes
necessary to incorporate thick plate theory to account for the
higher modal densities of bending waves due to the lower
wave speeds which occur in thick plate theory. This is illus-
trated in Fig. 6, where SEA predictions using thin and thick
plate theory coupling loss factors are compared to
measurements.30
Measurements of the transfer functions from the applied
force in machinery spaces to vibrations levels in the hull
plating of the USNS HAYES, a research ship used to conduct
acoustical measurements, are compared to SEA predictions
in Fig. 7.31 The measurements were taken in narrowbands
and the predictions generated for one-third octave bands. Be-
cause both the measurements and SEA predictions are of
transfer functions, the results, except for band averaging, are
not sensitive to bandwidths, so that direct comparisons can
be made in Fig. 7. For the transfer of forces applied by the
propulsor motor and in the generator room to the vibration of
the hull plating, agreement between the measured and pre-
dicted levels in Fig. 7 is good, with only two exceptions;
from 7 to 10 kHz for the propulsion motor transfer function
and from 10 to 20 kHz for the generator room transfer func-
tion, where the differences are slightly above 10 dB. The
structural losses were larger than predicted indicating that the FIG. 7. Comparisons of SEA predictions to measured force-to-vibration
structural loss factors were larger than predicted. All of the transfer functions on the USNS HAYES.

1787 J. Acoust. Soc. Am., Vol. 101, No. 4, April 1997 Burroughs et al.: Energy analysis 1787

Redistribution subject to ASA license or copyright; see http://acousticalsociety.org/content/terms. Download to IP: 131.156.59.191 On: Tue, 09 Sep 2014 15:02:03
challenge to the SEA user. Coupling loss factors for struc-
tures are related to power transmissibility, which is usually
easier to predict. Coupling between structures often requires
the inclusion of in-plane ~longitudinal and shear! as well as
bending waves which makes the derivation of accurate cou-
pling loss factors beyond the scope of this paper. Compari-
son of predictions from SEA models to measured levels, pre-
sented here, show that SEA is capable of providing accurate
predictions of broadband noise and vibration levels in com-
plex systems.
To be successful in providing accurate predictions, SEA
models must:
~1! Be capable of predicting the input power to all sub-
systems from applied input forces, moments, velocities,
or acoustic pressures. Measurements of input power is
also acceptable but is frequently as difficult as a predic-
tion.
~2! Represent all important mode types which occur within
the modeled subsystem in the frequency range of inter-
est, such as bending ~thin and thick plate and beam!
torsional, transverse shear, and longitudinal for structural
FIG. 8. SEA predictions versus measured cabin noise levels generated by
sonar pings on the USS GLOVER. elements; and one, two and, three dimensional modes for
acoustical spaces, and each subsystem must contain sev-
eral modes down to the lowest frequency band of inter-
son between the predicted and measured levels for twenty
est.
two cases. Agreement within 63 dB occurred when the dots
~3! Have valid coupling loss factors for all mode types
fall inside the dashed lines in Fig. 8. The range in the mea-
which interact at junctions of subsystems, such as struc-
sured levels taken during repeat pings are indicated by the
tural joints, mounts, and radiating surfaces. Damping
horizontal lines through the dots. The agreement shown in
loss factors must also be known or predicted.
Fig. 8 indicates that SEA can be used effectively to predict
~4! Provide the output energies or derived vibration levels
noise levels in complex systems with a large number of sub-
for those modes which are of interest, or will be mea-
systems. However, the SEA model for the GLOVER was de-
sured.
veloped by validating coupling and damping loss factors
with measurements conducted on the GLOVER. It would be Because of the relative simplicity of the SEA equations and
risky to use the model developed for the GLOVER for other power and energy outputs, SEA models can provide great
similar ships without some experimental verification of loss insight to the dynamic system and can provide useful guid-
factors. ance for noise control treatment design and optimization.
Several organizations have prepared SEA computer
IV. SUMMARY codes. Three of these codes have been developed commer-
The basic concepts and assumptions behind Statistical cially. SEAM,33 developed by Cambridge Colloborative Inc.,
Energy Analysis ~SEA! are presented. Starting with the was the first SEA computer code to become commercially
simple oscillator as a model of a resonant mode, coupling available. Recently, AutoSEA,34 developed by Vibro-
between resonant modes in subsystems is shown to be pro- Acoustic Sciences Ltd., has been made commercially avail-
portional to the differences in the energies in the modes. able. AARCSEA,35 which has been used in several naval
Classical modal expansions are used to justify equipartition vehicle and architectural designs, is available on an applica-
of energy among modes in a subsystem and the assumption tion basis to end user customers.
that each mode can be modeled as a simple oscillator. Power
1
balance equations for coupled subsystems are derived, which R. H. Lyon and R. G. DeJong, Theory and Application of Statistical En-
can be solved for the energies in each of the subsystems. The ergy Analysis ~Buttersworths-Heimann, Boston, MA, 1995!.
2
L. Cremer, M. Heckl, and E. E. Ungar, Structure-Borne Sound: Structural
required inputs to the power balance are the input powers to Vibrations and Sound Radiation at Audio Frequencies ~Springer-Verlag,
subsystems, modal densities, masses and damping loss fac- New York, 1988!, 2nd ed.
3
tors for each subsystem, and coupling loss factors between I. S. Gradshteyn and I. M. Ryzhik, Tables of Integral, Series, and Prod-
subsystems. Methods for estimating modal densities are pre- ucts ~Academic, New York, 1965!.
4
P. W. Smith, Jr., ‘‘Statistical models of coupled dynamical systems and
sented. Power inputs are specified by the user of the SEA the transition from weak to strong coupling,’’ J. Acoust. Soc. Am. 65,
power balance equations or computed for simple force or 695–698 ~1979!.
5
displacement inputs. Damping loss factors are usually ob- E. Skudrzyk, ‘‘The mean-value method of predicting the dynamic re-
sponse of complex vibrators,’’ J. Acoust. Soc. Am. 67, 1105–1135
tained from measurements of the materials or typical struc-
~1980!.
tures used in the subsystems. The derivation of accurate ex- 6
L. L. Beranek and I. L. Ver ~Eds.!, Noise and Vibration Control Engi-
pressions for coupling loss factors presents the greatest neering ~Wiley, New York, 1992!.

1788 J. Acoust. Soc. Am., Vol. 101, No. 4, April 1997 Burroughs et al.: Energy analysis 1788

Redistribution subject to ASA license or copyright; see http://acousticalsociety.org/content/terms. Download to IP: 131.156.59.191 On: Tue, 09 Sep 2014 15:02:03
7 22
J. H. Wendel and B. L. Scott, ‘‘Calculation of SEA Coupling Factors J. L. Horner and R. G. White, ‘‘Prediction of vibrational power transmis-
Using the Wave Approach-Results for Plates,’’ Cambridge Collaborative sion through bends and joints in beam-like structures,’’ J. Sound Vib.
Inc. Report 89-10-21610-2 ~1989!. 147~1!, 87–103 ~1991!.
8
V. V. Tyutekin, ‘‘Reflection and refraction of flexure waves at the bound- 23
I-T. Lu, H. L. Bertoni, and H-Y. Chen, ‘‘Coupling of plate waves at
ary of separation formed by two plates,’’ Sov. Phys. Acoust. 8~2!, 180– joints,’’ J. Acoust. Soc. Am. 92, 510–526 ~1992!.
183 ~Oct.–Dec. 1962!. 24
R. C. N. Leung and R. J. Pinnington, ‘‘Wave propagation through right-
9
S. V. Budrin and A. S. Nikiforov, ‘‘Wave transmission through assorted angled joints with compliance: longitudinal incidence wave,’’ J. Sound
plate joints,’’ Sov. Phys. Acoust. 9~4!, 333–336 ~April–June 1964!.
10 Vib. 153~2!, 223–237 ~1992!.
T. Kihlman, ‘‘Sound transmission in building structures of concrete,’’ J. 25
G. Maidanik, ‘‘Response of ribbed panels to reverberant acoustic fields,’’
Sound Vib. 11~4!, 435–445 ~1970!.
11 J. Acoust. Soc. Am. 34, 809–826 ~1962!.
W. Wohle, Th. Beckmann, and H. Schreckenbach, ‘‘Coupling loss factors 26
for statistical energy analysis of sound transmission at rectangular struc- F. Fahy, Sound and Structural Vibration: Radiation, Transmission and
tural slab joints, Part I,’’ J. Sound Vib. 77~3!, 323–334 ~1975!. Response ~Academic, New York, 1985!.
27
12
W. Wohle, Th. Beckmann, and H. Schreckenbach, ‘‘Coupling loss factors L. K. H. Lu, W. J. Hawkins, D. F. Downard, and R. G. DeJong, ‘‘Com-
for statistical energy analysis of sound transmission at rectangular struc- parisons of statistical energy analysis and finite element analysis vibration
tural slab joints, Part II,’’ J. Sound Vib. 77~3!, 335–344 ~1975!. predictions with experimental results,’’ Shock and Vib. Bull. 53~4!, 145–
13
V. A. Veshev and D. P. Kouzov, ‘‘Influence of the medium on the vibra- 153 ~1983!.
28
tions of plates joined at right angles,’’ Sov. Phys. Acoust. 23~3!, 206–211 J. Tratch, Jr., ‘‘Vibration transmission through machinery foundation and
~May–June 1977!. ship bottom structure,’’ MS thesis, MIT, 1985.
14 29
Y. Irie and T. Nakamura, ‘‘Random incidence energy transmission of R. H. Lyon, ‘‘In-plane contribution to structural noise transmission,’’
bending waves at junctions,’’ J. Acoust. Soc. Am. Suppl. 1 64, S26 Noise Control Eng. J. 26„1…, 22–27 ~1986!.
~1978!. 30
A. C. Aubert, J. H. Wendell, and F. R. Kern, ‘‘AARCSEA statistical
15
G. Rosenhouse, H. Ertel, and F. P. Mechel, ‘‘Theoretical and experimental energy analysis program—tutorial manual,’’ AARC Report 126 ~1991!.
investigations of structureborne sound transmission through a ‘‘T’’ joint 31
R. W. Fischer, F. R. Kern, and I. Cross-Whiter, ‘‘Preliminary test report—
in a finite system,’’ J. Acoust. Soc. Am. 70, 492–499 ~1981!. impedance measurements on machinery foundations aboard the HAYES,’’
16
M. J. Sablik, ‘‘Coupling loss factors at a beam L-joint revisited,’’ J. Atlantic Applied Research Corp. Tech. Memo 028 ~1988!.
Acoust. Soc. Am. 72, 1285–1288 ~1982!. 32
B. L. Burkewitz, ‘‘The update and validation of the USS GLOVER ~FF
17
J. F. Doyle and S. Kamle, ‘‘An experimental study of the reflection and 1098! statistical energy analysis model for analysis of ping-noise due to
transmission of flexural waves at an arbitrary T-joint,’’ J. Appl. Mech. 54,
operation of low-frequency, hull-mounted transmitter,’’ BBN Systems and
136–140 ~1987!.
18 Technologies, Inc. Report No. 7469 ~1990!.
R. C. N. Leung and R. J. Pinnington, ‘‘Wave propagation through right- 33
Contact J. Manning, Cambridge Collaborative Inc., 689 Concord Avenue,
angled joints with compliance-flexural incident wave,’’ J. Sound Vib.
Cambridge, MA 02138, Telephone ~617! 876-5777 for more information
31–46 ~1990!.
19
R. S. Langley, ‘‘Elastic wave transmission through plate/beam junctions,’’ on SEAM.
34
J. Sound Vib. 143~2!, 241–253 ~1990!. Contact P. Bremmer, Vibro-Acoustic Science, Inc., 5355 Mira Sorrento
20
M. D. McCollum and J. M. Cuschieri, ‘‘Bending and in-plane wave trans- Place, Suite 100, San Diego, CA 92121, Telephone ~619! 597-7535 for
mission in thick connected plates using statistical energy analysis,’’ J. more information on AutoSEA.
35
Acoust. Soc. Am. 88, 1480–1485 ~1990!. Contact F. Kern, Atlantic Applied Research Corp., 4 A Street, Burlington,
21
J. A. Moore, ‘‘Vibration transmission through frame of beam junctions,’’ MA 01803, Telephone ~617! 273-2400 for more information on AARC-
J. Acoust. Soc. Am. 88, 2766–2776 ~1990!. SEA.

1789 J. Acoust. Soc. Am., Vol. 101, No. 4, April 1997 Burroughs et al.: Energy analysis 1789

Redistribution subject to ASA license or copyright; see http://acousticalsociety.org/content/terms. Download to IP: 131.156.59.191 On: Tue, 09 Sep 2014 15:02:03

You might also like