You are on page 1of 20

Porous Materials for Sound Absorption and

Transmission Control
J. Stuart Bolton

Ray W. Herrick Laboratories, 140 S. Intramural Drive, Purdue University, West Lafayette IN
47907-2031, USA
bolton@purdue.edu

Abstract This paper begins with a discussion of the general types of porous materials,
i.e., rigid, limp, and elastic, and of their general physical properties. The macroscopic
properties (e.g., flow resistivity, porosity, tortuosity, etc.) that control the acoustical
behavior of each type of porous material are then defined and discussed, as are methods
for their measurement. The acoustical characterization of porous media is considered
next, followed by a discussion of modeling porous materials with particular reference to
elastic porous materials such as foams. The special characteristics of elastic porous
materials are illustrated through experimental and computational examples involving
sound absorption and sound transmission. In particular, the importance of apparently
small details of foam layer boundary conditions is emphasized. Finally, foam finite
elements that are capable of predicting the behavior of finite-sized noise control treatments
having realistic shapes are discussed. By using foam finite element models it is possible
to perform optimal design studies: i.e., to design real treatments that yield optimal
acoustical performance at a given treatment volume or weight.
1. INTRODUCTION
Porous noise control materials have two phases: the solid, usually fibrous component referred
to as the frame, and the interstitial fluid contained within the pores formed by the frame.
Owing to their relatively low density, porous materials are not generally useful as barriers,
but they are frequently used as efficient sound absorbing materials, in which role they convert
organized acoustical motion into heat. Sound absorbing materials dissipate acoustical energy
largely by the interaction of their solid and fluid phases. In particular they convert acoustical
energy into heat by: viscous means (associated with oscillatory shearing of the fluid in the
vicinity of the fibers surface); thermal means (irreversible heat flow from the interstitial fluid
to the fibers forming the frame); and by structural means (irreversible losses associated with
flexure of the fibers comprising the frame).

Porous materials such as glass fiber, mineral wool, and open or partially open cell foams have
many applications in automotive noise control in particular: e.g., as headliners, seats, under
carpets, trim lining, dashmat lining, cavity interiors, parcel self lining, panel damping, etc.
And of course they are widely used in appliance noise control, aircraft applications and in
buildings. The objective in this article is not to discuss all possible applications of porous
materials, but rather to introduce fundamental terminology and concepts related to porous
materials, and especially elastic porous materials such as foams, to facilitate later discussion
related to applications.

In this paper, porous materials will first be categorized, and then the macroscopic physical
properties that determine their acoustical performance will be discussed (as will techniques
for measuring those quantities). After a brief discussion of the acoustical quantities that are
used to characterize porous materials, some aspects of wave propagation in porous materials
will be introduced through a discussion of approaches to theoretical modeling of these
materials. A number of examples will be used to highlight the sensitivity of relatively stiff
porous materials to apparently small details of their installation. Finally, recent progress in
finite element modeling of foam will be described, and the use of that technology in the
optimal design of noise control treatments will be illustrated.

2. CATEGORIES OF POROUS MATERIALS
Porous materials may be categorized as being rigid, limp or elastic. The types of porous
materials that are simplest to characterize and model are those in which the expanded solid
phase (i.e., the fibers collectively comprising the frame) may be considered either rigid or
limp. The assumption of a rigid frame is appropriate when the solid phase does not move
significantly compared to the motion of the fluid phase, either because individual fibers are
immobilized by the stiffness of interconnecting filaments, or because they are much denser
than the interstitial fluid [1]. The limp approximation is useful, on the other hand, when the
fibers making up the material are sufficiently unconstrained and light that they can move
freely as a result of viscous or inertial coupling with the interstitial fluid [2,3]. A porous
material may generally be assumed rigid when the in vacuo bulk modulus of elasticity of its
expanded solid phase is much greater than that of the interstitial fluid, usually air in noise
control applications, and if its solid phase is not directly excited by attachment to a vibrating
surface. Conversely, a material may be considered limp when the in vacuo bulk modulus of
elasticity of its expanded solid phase is much less than that of air.

When the frame bulk modulus is significantly greater or smaller than that of the interstitial
fluid, a porous material can support only a single longitudinal wave type [4,5]. As a result,
the material can be treated as an effective fluid characterized by frequency-dependent,
complex parameters (e.g., the density and sound speed) that together account for the effects
of viscosity, pore tortuosity and heat conduction (and frame motion in the case of limp
materials) [5]. The acoustical properties of rigid or limp porous media may thus be described
most easily by defining a complex wave number, k = - j, where is the propagation factor
and is the attenuation factor, and a complex characteristic impedance, z
c
.

On the other hand, published experimental results show that some flexible porous materials
such as polyurethane foam cannot be modeled as being either rigid or limp owing to the
ability of the materials frame to support wave motion independent of the interstitial fluid
[6,7]. The in vacuo frame bulk modulus of an elastic porous material is of the same order
as that of air, and as a result both the frame and fluid take a significant and distinct role in the
wave propagation process.

Elastic porous materials can support three wave types simultaneously: the so-called airborne,
frame and shear waves [8]. The degree to which each of these wave types is excited depends
critically on the boundary conditions at the surfaces of the porous material, particularly the
manner in which those surfaces connect to vibrating solids. For this reason, apparently minor
aspects of surface treatment attachment can play a very important role in determining the
installed acoustical behavior of elastic porous materials such as polyurethane and polyimide
foams [9-11], much more so than is the case for either rigid or limp porous media.

Note that most porous materials are physically anisotropic: i.e., their physical properties are
different in different directions. This is particularly true for foams that exhibit a preferred
rise direction [12], but is also true for many fibrous materials that have a preferred fiber
direction resulting from the process by which they are manufactured [13-15]. Some
theoretical work has been directed towards the study of foam anisotropy [16]; nevertheless, it
is most often assumed that noise control materials are macroscopically homogeneous (except
when explicitly layered) and isotropic.

It is also common experience that most porous materials are spatially inhomogeneous: i.e.,
their macroscopic properties vary randomly throughout the material [12]. It is not
uncommon for the flow resistivities of two closely spaced samples taken from a single piece
of porous material to differ by a factor of two: the same tends to be true of stiffness
properties. Insofar as this effect is acknowledged, the properties of a porous material are
usually estimated by averaging the results for a number of individual samples. Currently
popular porous material theories do not explicitly account for the spatial inhomogeneity of
macroscopic properties: the development of such a theory would result in a useful practical
tool. However, the most important point is that the macroscopic properties of a single, small
sample of porous material are unlikely to represent the average properties of the material.

Porous materials also exhibit nonlinearities resulting from both nonlinear inertial effects
related to pore fluid flow [17-19] and nonlinear frame elasticity [20]. Nonlinear effects, of
course, become progressively more significant as the excitation levels increase. In this
article, the discussion is limited to the linear regime that is normally appropriate when porous
materials are used in noise control applications. As a result, some of the present discussion is
not relevant to situations in which the porous material experiences large static or dynamic
strains (as in seating applications, for example).

3. MACROSCOPIC PHYSICAL PROPERTIES
3.1 Introduction
Most porous material theories are formulated in terms of the macroscopically measurable
physical properties of the frame and the fluid: e.g., the flow resistivity. The advantage of
such an approach is that it enables the influence of the various directly measurable physical
properties of the porous material to be identified; thus a particular set of physical parameters
can be specified that will result in a porous material having a specified level of acoustical
performance. In addition, easy-to-use approximate models derived by neglecting the
physical parameters that are judged to be least significant in a particular situation (e.g., the
bulk elastic properties in the case of a nearly rigid porous material exposed to airborne
incident sound) may be used in preliminary analyses.

It should be noted, of course, that there is a direct link between the microscopic structure of a
porous material (e.g., fiber radius, fiber shape, fiber orientation, fiber material density,
number of fibers per unit material volume, etc.) and its macroscopic properties. To-date,
however, there is little information available regarding the links between a porous materials
microscopic and macroscopic properties (see, however, the discussion by Bies of the relation
between fiber size, bulk density and the flow resistivity of fibrous materials [21]). Although
the link between material microstructure and macroscopic properties represents a promising
avenue for future research, in this article the discussion will largely be limited to
macroscopically measurable parameters.

The most important macroscopic physical properties of a porous material are: flow
resistivity, connected porosity, pore tortuosity (these first three together constituting the
materials fluid-acoustical properties), bulk density, in vacuo bulk modulus, shear modulus
and loss factor (the latter four properties comprising the elastic properties of a porous
material). The acoustical performance of the various types of porous media can be predicted
with a knowledge of all or a sub-set of these properties: i.e., (i) when a material is rigid,
only a knowledge of the fluid-acoustical parameters is required; (ii) when a material is
limp, both the fluid-acoustical parameters and the bulk density must be known; and (iii)
when a material is elastic, a knowledge of both the fluid-acoustical and elastic properties is
required. In the following sub-sections each of the macroscopic properties listed above will
be discussed briefly.
3.2 Porosity
The porosity of a porous material is defined as the ratio of the volume occupied by the fluid
within the porous material (the volume of the voids) to the total volume of the porous
material and it is thus dimensionless. By its nature, the porosity falls within the range zero to
unity.

Since the porosity quantifies the relative volume occupied by the solid and fluid phases
within a unit volume of porous material, it is a key parameter in theories of sound
propagation in porous material [1,5,8]. However, the porosity of typical noise control
materials such as foam and glass fibers is normally very high: i.e., greater than 0.9, and often
greater than 0.98. Since the porosity of most noise control materials is so high, and because
the variation in porosity tends to fall into a narrow range, porosity is generally not very
important when distinguishing amongst typical noise control materials. However, it should
be remembered that much of the energy dissipation within a porous material results from the
relative motion of the solid frame and the interstitial fluid, and that for this process to work
there must be continuous paths through the material: i.e., the materials pores must be
connected. Therefore, materials featuring closed cells, e.g., Styrofoam, do not normally
exhibit useful acoustical properties, and should be modeled as elastic solids rather than as
porous materials.

The most direct way of determining a materials porosity is to measure the volume of air
contained within it. This task may be achieved, for example, by using the apparatus
developed by Champoux et al. [22] that was itself based on a device developed by Beranek
for the measurement of static porosity [23]. The operation of this device is based on the ideal
gas law at a constant temperature (i.e., Boyles law). When the temperature of a rigid
chamber containing a porous sample is held constant, a measurement of the change in air
pressure that accompanies a known change in volume allows the volume of fluid within the
sample to be determined if the change in air pressure accompanying the same volume change
in a rigid, empty chamber of the same total volume is known. Note that it is assumed in this
measurement that the fibers comprising the solid component of the material are themselves
incompressible.

The porosity may also be determined directly from a knowledge of the bulk density of the
material when the density of the material comprising the solid phase is known.
3.3 Flow Resistivity
The specific (unit area) flow resistance of any layer of porous material is defined as the ratio
of the pressure difference across the layer to the steady state air velocity exterior and
perpendicular to the two faces of the layer [23]: it is conventionally given the units MKS
Rayls. The flow resistivity is then the specific flow resistance per unit material thickness and
it has SI units of MKS Rayls/m. The flow resistivities of useful noise control materials vary
widely, but typically fall within the range 1 10
3
MKS Rayls/m to 1 10
6
MKS Rayls/m.

The flow resistivity depends on the porosity of a material as well as its tortuosity, but for high
porosity, low tortuosity fibrous materials, the flow resistivity is approximately inversely
proportional to fiber radius squared at a constant bulk density. That is, a large number of
small diameter fibers results in a higher flow resistivity than does a small number of large
diameter fibers [21]. At a microscopic level, flow resistance results from the formation of a
viscous boundary layer as fluid flows over each fiber, and the amount of shearing in those
boundary layers, and hence the amount of viscous drag per unit mass of fibers exerted by the
flow on the fibrous medium, increases as the mean fiber radius decreases. The flow
resistivity is thus usually taken to be a measure of the viscous coupling between the fluid and
solid phases of the porous material, and so is a measure of the potential for viscous
dissipation of sound.

Note that the steady state flow resistivity as discussed above, gives an accurate estimate of
the dynamic flow resistivity experienced by oscillatory flow within the porous material only
in the low frequency limit. When the flow is oscillatory, as when created by a sound field
within a porous material, the steady state flow resistivity must be corrected to account for
frequency-dependent boundary layer velocity profiles: that correction is an integral part of
modern porous material theories [24].

Flow resistivity measurement has been standardized by the ASTM [25]. Note that the
steady-state flow resistivity of porous materials is a nonlinear function of flow velocity
owing to inertial effects and to turbulence at very high flow rates. Tests should normally be
conducted at flow rates sufficiently low to fall within the linear range, or be extrapolated to
the zero flow rate limit, since the linear flow resistivity is appropriate for acoustical situations
in which particle velocities are typically on the order of several mm/s or less. The pressure
difference across the test sample can be measured by using a manometer, and airflow
velocities are usually inferred from the volume flow rate through the sample.

3.4 Pore Tortuosity
The tortuosity (sometimes referred to as the structure factor [7]) is a measure of the deviation
within the porous material of the pore axis from the direction of wave propagation through
the porous material, and of the pores nonuniformity in cross-sectional dimension. Tortuosity
was introduced into the theory of rigid porous materials by Zwikker and Kosten [7] to
account for fluid acceleration effects associated both with the deviation of pores from straight
lines in the direction of wave propagation, and with pore expansions and contractions. In the
case of limp and elastic porous materials, the tortuosity is also used to quantify the inertial
coupling of the solid and fluid phases of a porous material resulting from the deflection of
fluid flow by the pore walls. The minimum value of pore tortuosity is unity, and fibrous
materials such as glass fibers typically have tortuosities of not much more than unity: e.g.,
1.5 or less. The same is true of fully open cell foams. However, when a foam is partially
reticulated, i.e., when its cells are partially closed by residual membranes, it may possess a
relatively high tortuosity: up to 10, say. The same may be true for relatively dense acoustical
materials: e.g., acoustic plasters and porous ceramics or aluminum.

There is no direct way of measuring a porous materials tortuosity. However, Champoux and
Stinson [26] have developed a procedure based on the measurement of electrical conductivity
to infer the tortuosity of porous materials: their procedure is based on earlier work by Brown
[27]. In this measurement, the voltage difference resulting from the passage of a high
frequency alternating current through a fluid-saturated sample is measured (the saturating
fluid must be electrically conductive). When the conductivity of both the fluid and of the
fluid-filled sample are known by measurement, the effective path length through the material,
and hence the tortuosity, may be calculated (when the sample porosity is known). Note that
this technique has been most widely used to measure the tortuosity of granular media [27]. It
is not clear that this procedure yields accurate results for typical noise control materials
whose structural or chemical properties may be altered by immersion in a conducting fluid.

More recently, it has been suggested that the tortuosity may be estimated from ultrasonic
reflection measurements. This suggestion is based on the observation that the reflection
coefficient from a porous material is directly related to its tortuosity [28] at very high
frequencies. Note, however, that for this measurement to be successful, the probe frequency
must still be such that the wavelength of sound is very large compared to the pore size and
any surface irregularities: otherwise, scattering effects will render the results inaccurate.
3.5 Bulk Density
The bulk density of an expanded porous material is simply defined as the ratio of an
expanded samples mass to its total, expanded volume. The bulk density and the porosity are
simply related by the expression
1
=
t
(1-h), where
1
is the bulk density of the expanded
porous material,
t
is the density of the material that comprises the solid phase of the porous
material, and h is the porosity of the porous material. Note that since the porosity of noise
control materials is often greater than 0.98, the bulk density of a porous material is usually
less than one-fiftieth of the density of the material from which the solid phase is formed.
3.6 Bulk Elastic Properties
The in vacuo bulk Youngs modulus of the expanded solid phase of a porous material is
defined, as usual, as the ratio of axial stress to strain in the absence of lateral constraint. It is
generally made complex to account for losses due to frame flexure and is of the form
E(1+j), where E is the (real) Youngs modulus, or storage modulus, and is the loss factor
or loss tangent [29-31]. The SI unit of the bulk Youngs modulus is Pa. As noted previously,
porous materials can be categorized as limp, elastic or rigid depending on whether the
materials bulk modulus is less than, of the same order as, or greater than the bulk modulus of
air (i.e., 1.4 10
5
Pa), respectively. For elastic porous materials in particular, the bulk
Youngs modulus is an important dynamic mechanical property that significantly affects the
acoustical behavior of the material since it controls (along with the Poissons ratio or shear
modulus) the speed of the longitudinal and shear waves that propagate primarily through the
frame [8]. The bulk Youngs modulus of polyurethane noise control foam of medium
density, say 10 kg/m
3
to 20 kg/m
3
, is typically on the order of that of air, but can vary widely
from one foam type to another depending on the precise chemical composition of the foam
and the details of its manufacture. The stiffness of fibrous materials, on the other hand, tends
to be substantially less than that of air, unless the materials have been purposely stiffened by
the addition of a binding agent, for example.

A dynamic method should be used to measure the bulk Youngs modulus since it has been
observed to be frequency-dependent. Kim, Wijesinghe and Kingsbury [29,30] determined
the complex modulus of polyurethane foams by measuring the complex transmissibility of a
single degree of freedom system in which a disc shaped foam specimen acted as an
essentially massless stiffness element. Okuno [31] proposed a similar experimental setup
that is capable of measuring the complex bulk Youngs modulus of poroelastic materials.
Following Okuno, the bulk Youngs modulus of a foam sample can be estimated from
measurements of the normal mechanical impedance of the sample under the assumption that
the sample layer behaves like a simple spring.

Note that dynamic stiffness measurements should, in principle, be conducted in a vacuum so
that the damping effect of the interstitial fluid is eliminated and the frame in vacuo loss factor
can be accurately identified. In vacuo measurements of porous material elastic properties
have been reported by Ingard et al. [32,33].

The bulk shear modulus of an elastic porous material is also, in general, frequency-dependent
and complex. The magnitudes of the bulk Youngs and the shear moduli are related by the
Poissons ratio, (when it can be assumed that the material is linear, homogeneous and
isotropic). The shear modulus of a foam can be estimated directly by holding two samples
vertically between two fixed outer plates and a central plate that is driven vertically by a
mini-shaker. The shear modulus is then estimated from the measured mechanical impedance.

Finally, note that recent finite element calculations indicate that the Youngs and shear
moduli as estimated from mechanical impedance measurements of the type suggested here
are sometimes not accurate owing to effects related to finite sample size and sample
constraint on the driven surface [34]. The latter observation would suggest that it may be
better to measure the input mechanical impedance of finite-sized blocks of material,
constrained in known ways at their boundaries, and then estimate the frame stiffness
properties by matching the measured results with finite element predictions [35].
4. ACOUSTICAL CHARACTERIZATION OF POROUS MATERIALS
4.1 Introduction
Acoustical porous materials are currently used in many noise control applications in the
automotive, aircraft and appliance industries; they can be used to absorb airborne sound (e.g.,
automotive interior headliners) or to enhance the transmission loss of barrier systems (e.g.,
aircraft fuselage or automotive dashpanel applications). The performance of acoustical
materials can be characterized in a number of ways, and when specifying performance levels
it is important that they be characterized in a manner relevant to their end use.
4.2 Sound Absorption
The performance of a treatment applied to an interior surface to absorb airborne sound is
most often characterized by its normal or random incidence absorption coefficients, which
are measures of the fraction of the incident acoustic power that is absorbed by the treatment
when sound strikes its surface at normal incidence, or with equal likelihood from all angles,
respectively. The normal incidence absorption coefficient of a sample can be measured
directly by using a standing wave tube. The random incidence absorption coefficient can be
measured directly in a reverberant room, or it can be calculated by integrating the
appropriately weighted plane wave absorption coefficient over all possible angles of
incidence [36]. The plane wave absorption coefficient itself is the fraction of the incident
power that is absorbed at a surface when a plane wave strikes it at a particular angle, and it is
defined as = 1 - |R|
2
, where R is the plane wave reflection coefficient. The plane wave
reflection coefficient can, in turn, be determined from a knowledge of the normal specific
impedance of a surface, and the latter quantity can be determined from theory, if the material
properties and its geometry are known, or it can be measured directly. The normal incidence
surface normal impedance can be measured using a standing wave tube, as can the normal
incidence reflection and absorption coefficients. Since the surface normal impedance, the
reflection coefficient and the absorption coefficient can all vary with angle of incidence, it is
sometimes useful to measure them at oblique incidence angles. A number of free field
measurement techniques have been proposed for that purpose [24,37].

All the quantities mentioned in the last paragraph should strictly be referred to as derived
quantities since they are not inherent properties of the porous materials, but rather depend
both on the material and its installation (e.g., layer depth and surface treatment). The
acoustical properties that are inherent to a porous material (at least those that may be
considered to be either limp or rigid, and so may be modeled as an equivalent fluid) are its
characteristic impedance and wave number. If the latter two quantities are known then all
other relevant acoustical quantities of a porous treatment can be derived.

The characteristic impedance of a limp or rigid porous medium is the ratio of pressure to
particle velocity in the direction of wave propagation in a plane wave propagating in an
infinitely extended piece of material. In practice, the characteristic impedance is equal to the
normal incidence surface normal impedance of a porous layer that is sufficiently deep that no
internal reflections return to the surface of the porous layer. Thus the characteristic
impedance of limp or rigid porous materials may be estimated by measuring the surface
normal impedance of a deep layer of porous material placed in a standing wave tube [6].

The wave number of a plane wave propagating freely within a porous material is complex
and is a measure of both the speed of wave propagation (and hence the wavelength within the
material) and the rate at which a propagating wave attenuates. The wave number within a
porous material can be estimated directly by passing a probe tube through a deep sample
exposed to normally incident sound at a single frequency and measuring the wavelength and
rate of attenuation [e.g., 38,39]. In an elastic porous material, the situation is not so
straightforward, however, since three wave types, each having a different speed and
attenuation rate, can propagate simultaneously.

Porous materials, such as glass fibers or foams, can also be categorized as being either locally
reacting or non-locally reacting. Physically, a porous layer can be assumed to be locally
reacting when the speed of wave propagation in the porous layer is much less than the speed
of sound in the ambient medium, with the result that transmitted waves propagate normal to
the material surface regardless of incidence angle. Owing to the latter effect, the surface
normal impedance of a locally reacting porous material is independent of the incidence angle
[36], with the important practical consequence that the reflection coefficient of a locally
reacting porous material may be predicted using a simple one-dimensional model of the
porous material [24,40].

In addition, simple relations between a locally reacting materials plane wave reflection
coefficient and its random incidence absorption coefficient were derived by Morse [41]. For
example, Bies and Hansen [42] have combined relations developed by Morse [41] with semi-
empirical formulae of Delany and Bazley [43] to obtain theoretical predictions of the random
incidence absorption coefficients of unfaced, finite-depth, locally reacting fibrous layers.

Thus, it is very convenient from a computational point of view if an absorbing material may
be considered locally reacting. Note that non-locally reacting materials may be made locally
reacting by segmenting the material into small elements by using rigid partitions: i.e., by
placing the porous material within a honeycomb structure. In this case, the partitions prevent
sound propagation parallel to the surface making the material effectively locally reacting.

On the other hand, when a material is non-locally reacting (i.e., when the sound speed in the
material is not small compared to that of air), its surface normal impedance varies with angle
of incidence. As a result, the random incidence absorption coefficient of such a material
must usually be found by applying numerical integration techniques [24,37]. A two-
dimensional theoretical porous material model of the type described in the next section is
then required to make predictions of the reflection coefficient of non-locally reacting
materials [11]. Obviously, from a theoretical point of view, it would be convenient if all
porous materials could be treated as surfaces of local reaction: however, in most situations
the latter assumption is not strictly valid.
4.3 Sound Transmission
As noted in the Introduction, porous layers are not generally good barriers and are thus not
effective when used by themselves at reducing sound transmission between two spaces.
However, they are often used to line the space between two panels: in that role they serve
both to decouple the panels, that is to suppress the deleterious effects of the mass-air-mass
resonance and inter-panel depth resonances, and to enhance the total transmission loss at high
frequencies through the dissipative action of the porous material [44]. Note that a structural
base panel, a layer of porous material, and a limp barrier layer (made of vinyl or rubber, for
example) forms a double panel system. Dashpanel and dashmat combinations are an
example of this kind of system, as are typical automotive floor treatments.

Whether lined or unlined, the acoustical performance of a single or multi-panel barrier
system can be quantified by measuring or predicting its power transmission coefficient (or its
decibel equivalent, the transmission loss). The power transmission coefficient is defined as
= W
t
/W
i
, where W
t
is the sound power transmitted by a barrier and W
i
is the sound power
incident upon the treatment. The transmission loss is then TL = 10log(1/). The
transmission loss is normally measured by inserting the test barrier in an aperture between
two reverberation rooms [45] or between a reverberation room and an anechoic space [46].
In the latter case, the transmitted sound power can be measured directly by scanning an
intensity probe over the transmission side of the barrier.

The sound transmission loss of barrier treatments lined with porous materials is an important
design factor in vehicles as well as in other applications, and it can be predicted when the
properties of the porous lining are known. For example, Beranek and Work [47] developed
systematic procedures for predicting the normal incidence transmission loss of multi-panel
structures lined with porous materials (and also made measurements using a modified version
of the apparatus described in Reference [48]). An excellent source of information on the
fundamentals of sound transmission through single and double panel systems is the book by
Fahy [49]. A modification of the two-microphone standing wave tube now makes it possible
to measure the normal incidence transmission loss of acoustical materials directly [50].

Recently, Bolton, Shiau and Kang presented predictions and corresponding measurements of
the random incidence transmission of foam-lined double panel structures [10,11]. The latter
measurements are particularly significant since they illustrate the dramatic difference in
transmission loss that can result from apparently minor changes in the way the lining material
is attached to the facing panels. The theory lying behind those predictions will be presented
next and illustrative transmission loss results will be given.
5. MODELING OF WAVE PROPAGATION IN ELASTIC POROUS MEDIA
5.1 Background
In this section, theoretical modeling of elastic porous materials is discussed briefly. Note that
both limp and rigid porous material models may be derived from the model presented here as
limiting cases [2,24]. The main difference between rigid or limp and elastic porous materials
is that the latter type can support several wave types simultaneously while the former type
can support only a single longitudinal wave. Thus elastic porous material theories are
directed towards the prediction of wave numbers for each wave type and the definition of
relations between stresses and motions in the two phases. If such information can be
predicted based on a knowledge of the physical parameters of the material, then accurate
theoretical predictions of the materials acoustical behavior may be made.

Note also that almost all porous material theories are based on the concept that the real
porous material can be replaced by a smeared medium that is homogeneous and isotropic,
but that has the same macroscopic bulk properties as the real porous material. This approach
limits the applicability of such models to frequencies sufficiently low that the wavelength of
sound is very large compared to the typical pore size in the medium. It is also usually
assumed that porous materials are homogeneous in the sense that their properties are
independent of position on a scale large compared to a pore size but small compared to a
wavelength. As was noted above, this assumption is rarely true in practice, but it is not yet
clear how serious a problem it presents. Also note that there have been some attempts to
develop theoretical models based explicitly on particular microstructures. To-date this
approach has been brought to bear on both fibrous media [13-15] and foams [51].

One of the first elastic porous material models was that due to Zwikker and Kosten who
showed (in earlier work summarized in reference [7]) that two types of longitudinal waves
can propagate in a porous material (i.e., the frame-wave and the airborne wave). Beranek
later arrived at the same conclusion using a similar theory: he also examined the rigid and
limp limits. Subsequently, Biot [8] developed a very general model for elastic porous
materials that allows for both longitudinal and shear modes of three-dimensional wave
propagation in elastic porous media. Biots model has, in time, become the standard model
for wave propagation in elastic porous media.

In order to investigate the transmission loss of foam-lined multi-panel structures, Bolton,
Shiau and Kang [10] adapted Biots theory to describe wave propagation in noise control
materials. Since the porous layer in most foam-lined structures has a depth that is much
smaller than its lateral extent, their theory was presented in two-dimensional form. This
theory can be used to predict the behavior of plane, layered systems of infinite lateral extent,
and thus in practice is limited to predicting the transmission loss of barriers that are large in
cross-section compared to the wavelength of the incident sound. Nevertheless, all three wave
types known to propagate in elastic porous materials were taken into account and their theory
is capable of predicting oblique and random incidence behavior of elastic porous materials.
They also developed sets of boundary conditions for several configurations of panel
structures. By comparing their predictions with random incidence transmission loss
measurements for various configurations of panel structures, the model was shown to be
accurate [10]. Allard et al. [24,49] have also predicted the oblique incidence behavior of
elastic porous materials based on Biots theory [8].
5.2 Theoretical Development
The development of a theory for wave propagation in elastic porous media proceeds from
dynamic and stress-strain equations for the two phases of the porous material: the solid frame
and the interstitial fluid. By using the sign convention shown in Fig. 1, one can write
appropriate stress-strain relations and dynamic relations for the solid and fluid phases. Note
that in the dynamic relations there appear inertial fluid-structural coupling terms proportional
to the relative acceleration of the two phases, and viscous coupling terms proportional to the
relative velocity of the two phases. The inertial coupling is directly related to the pore
tortuosity, while the viscous coupling can be related to the flow resistivity. After combining
the two sets of relations and performing a sequence of vector operations, one obtains two
wave equations that govern the propagation of volumetric and rotational strains within the
material [10,11]: i.e.,
0
2
2
1
4
= + +
s s s
e A e A e (1)
0
2 2
= +
t
k
(2)

where e
s
and are the volumetric and rotational strains, respectively, and k
t
is the wave
number of the rotational wave. Note that a fourth order wave equation having the form of
Eq. (1) indicates the existence of two longitudinal wave types; a second order wave equation
having the form of Eq. (2) governs the propagation of a single transverse wave.

When appropriate two-dimensional solutions of Eqs. (1) and (2) are assumed for the solid
volumetric strain and rotational strain, all the field variables such as fluid and solid
displacements and stresses can be expressed in terms of six wave component amplitudes
(representing three wave types propagating upwards and downwards within the foam layer)
and material constants. The unknown wave component amplitudes are then determined by
imposing sets of boundary conditions as appropriate to represent different physical situations.
5.3 Boundary Conditions
As examples, the sets of boundary conditions for an open (i.e., unfaced) surface and a sealed
surface (i.e., the elastic porous material is assumed to be bonded to a facing panel) are shown
below. In the latter case, the panel has been modeled as an Euler-Bernoulli panel: i.e., one
possessing flexural stiffness.

At the open surface of the elastic porous layer (illustrated schematically in Fig. 2(a)) there are
four boundary conditions to be satisfied: two normal stress continuity conditions, a normal
volume velocity continuity condition and one shear stress-free condition. These conditions,
are, respectively [10,11]:
s hP =
(3a)


y
P h = ) 1 (
(3b)


y y y
hU j u h j v + = ) 1 (
(3c)

0 =
xy

(3d)

where h is the porosity, P is the pressure in the exterior acoustic field at the interface, v
y
is the
normal component of the acoustic particle velocity in the exterior medium at the interface, s
is the fluid stress within the porous material,
y
is the solid stress within the porous material,

xy
is the shear stress acting on the solid phase, and u
y
and U
y
are, the solid and fluid
displacements, respectively.

At an open surface of this type there is no constraint placed on the individual velocities of the
solid and fluid phases of the material: only the volume velocity must be continuous at the
interface. As a result, the partial impedances of the fluid and solid phases add in parallel, and
the motion in the external fluid region (i.e., the external sound field) primarily drives the
phase of the porous material that has the lowest impedance, which would normally be the
fluid phase of the porous material. Thus, at an open surface, airborne incident sound couples
most directly to the airborne wave within the material [10]. This is usually advantageous
since the airborne wave is usually much more heavily damped than is the frame wave [11].

When the elastic porous layer is bonded directly to an elastic panel (illustrated schematically
in Fig. 2(b)) there are six boundary conditions to be satisfied: i.e., normal velocity continuity,
normal displacement continuity, tangential displacement, in-plane acceleration, and force
(i.e., the panel equation of motion). These conditions are, respectively [10,11]:

t y
W j v =
(4a)


t y
W u =
(4b)


t y
W U =
(4c)


dx
dW
h
W u
t
p
p x
2
) / ( + =
(4d)

( )
p s x p xy
W m k D
2 2
) / ( = +
(4e)

( )
t s x xy
p
x p
W m Dk
h
jk q P
2 4
2
) / ( ) / ( = + +
(4f)

where h
p
is the panel thickness, D is the panel flexural stiffness per unit width, D
p
is the panel
longitudinal stiffness per unit width, m
s
is the panel mass per unit area, q
p
is the normal force
per unit panel area exerted on the panel by the elastic porous material, W
t
is the transverse
panel displacement, W
p
is the panel in-plane displacement, and the first sign is chosen when
the foam is attached to the panel surface facing in the positive coordinate direction.

In this case, since the panel is directly attached to the solid phase of the porous material, the
velocities of the solid and fluid phases are constrained to be equal to each other (and to the
normal velocity of the panel and the external acoustical field). As a result of this velocity
constraint, the impedances of the fluid and solid phases add in series and, since the
impedance of the solid component is usually higher than that of the fluid component, the
former largely determines the response of the system. In practice this means that it is the
relatively lightly damped frame wave that is primarily excited in this configuration. The
shear wave may also be excited in this case [10].

The open and bonded boundary conditions have been discussed in some detail here to
emphasize that the degree to which each of the wave types that can propagate within an
elastic porous material is actually excited depends on the boundary conditions at the
materials surface. The acoustical performance of a foam layer, for example, may appear to
be completely different depending on whether it is directly attached to a facing panel, or
whether it is separated from it by a small air gap. The difference in performance may be
traced to the dominant role played by the frame wave in one case and the airborne wave in
the other. Thus, when using elastic porous materials in noise control applications it is very
important to understand the effect of boundary conditions on the installed performance of a
treatment so that the optimum arrangement can be chosen in each particular instance.

When the boundary conditions appropriate to a particular panel configuration have been
defined, the various assumed field solutions can then be substituted into them. The result is a
matrix equation for the wave amplitudes in the elastic porous material, facing panels, air gaps
(if any), as well as for the reflected and transmitted wave amplitudes that are usually the
objective of the calculation. Sample results will be discussed in the next section.
6. EXPERIMENTAL AND COMPUTATIONAL EXAMPLES
In this section, theoretical predictions of the transmission loss of foam-lined double panel
structures are compared with experimental results [10,11] with the intention of emphasizing
the importance of the boundary conditions at the foam surface. The foam parameters used in
the predictions were: bulk density of the solid phase
1
= 30 kg/m
3
, in vacuo bulk Youngs
modulus E
m
= 8 10
5
Pa, in vacuo loss factor = 0.265, bulk Poissons ratio = 0.4, flow
resistivity = 25 10
3
MKS Rayls/m, tortuosity ' = 7.8, and porosity h = 0.9. The thickness
of the aluminum panels on the incident and transmitted sides were, respectively, 0.05 in and
0.03 in, and the panels were separated by distances ranging from 21 mm to 41 mm. The
measurements were performed by placing the double panel system in a 1.14 m 1.14 m
aperture between a reverberant source room and an anechoic receiving space. The
transmission loss was measured using the intensity technique described earlier.

Figures 3(a) and (b) show the predicted and measured transmission losses for the BB (foam
bonded directly to both panels) and UU (foam separated from the panels by small air gaps)
configurations. It may be seen that the theoretical predictions follow the measured results
reasonably closely. While the BB configuration gives the higher transmission loss below 400
Hz, the UU configuration transmission loss displays an approximately 18 dB/octave slope
above the double panel resonance near 250 Hz, with the result that the transmission loss of
the UU configuration is substantially higher than that of the BB configuration (by
approximately 15 dB) at frequencies above 400 Hz. The double panel resonance occurs
when the panel masses vibrate against the lining stiffness. Note that the BB double panel
resonance frequency is elevated compared to that of the UU configuration due to the
relatively high stiffness of the directly attached foam lining in the former case.

The relatively poor performance of the BB configuration results from the direct coupling of
the facing panels and the solid phase of the porous material, which in turn results in strong
excitation of the frame wave. By contrast, when the foam is separated from the facing panels
by even the smallest of air spaces, it is the highly damped airborne wave that is primarily
excited in the foam, thus causing the transmission loss in this configuration to be relatively
high. It is important to realize that the masses per unit area of both the UU and BB
configurations shown here were exactly the same (as, for practical purposes, were their
thicknesses). The dramatic difference in their transmission losses originates solely in the
difference in the foam/panel boundary conditions in the two cases.
7. FINITE ELEMENT MODELING OF FOAMS
Although analytical models [8,10,24,52] can account for all three wave-types propagating
within elastic porous materials and are useful and easy to implement, they can in practice
only be used to model laterally infinite planar treatments. Since realistic noise control
elements feature foam pieces that are of finite size, and usually, irregular shape, there is
clearly a need for numerical tools to allow the performance of such treatments to be predicted
accurately. Thus, foam finite elements, based on Biots elastic porous material theory, that
can be coupled with existing acoustical and/or structural finite elements have recently been
developed [53-58]. Here, examples of the use of foam finite elements in noise control design
procedures will be given: the detailed mathematical formulation of foam finite elements and
the procedures required to couple them to adjacent media are described elsewhere [53-58].

The first example involves shape optimization of a foam treatment. In particular, the shape
of a foam wedge terminating a waveguide was optimized to offer the maximum absorption
over a specified frequency range: 500 Hz to 2 kHz. The model considered here is illustrated
in Fig. 4: a foam wedge was placed in a two-dimensional hardwalled waveguide (27 cm long
and 5.4 cm wide) and a rigid piston was considered to generate a plane sound field that was
incident on the front surface of the wedge. When the volume of the wedge and the duct
width were held constant, the geometry of the wedge could be defined by a single design
parameter, the wedge tip angle, . The same foam properties as those used in the previous
section were used in the calculations presented here.

By calculating the normal incidence absorption offered by wedges having various tip angles,
the optimal wedge angle was found to be 36
o
for the foam material considered here. In Fig.
5(a), the absorption offered by the optimized wedge is compared with the absorption of non-
optimal wedges having tip angles of = 132
o
and 180
o
(i.e., a flat foam layer). Generally, it
was found that wedge absorption increased at high frequencies at the expense of a reduction
of absorption at low frequencies as the wedge was made sharper and longer (causing the
constrained edges to become shorter). In Fig. 5(b), the absorption coefficient averaged over
the frequency range 500 Hz to 2 kHz has been plotted for various wedge angles. An
extensive discussion of wedge shape optimization (including inverse and hybrid wedges) can
be found in reference [55]. The main point to be taken from this example, though, is that the
numerical tools are now in place to use the techniques of optimal design to identify the best
possible foam configuration for a particular application.

As another example, foam finite elements were applied to investigate normal incidence sound
transmission characteristics through a foam wedge placed in a hardwalled duct as shown in
Fig. 6(a). As may be seen from Fig. 7(a), it was found that in some frequency bands the
transmission loss of the wedge was significantly higher than that of a plane foam layer of the
same volume (especially when the wedge tip-angle was smaller than 48
o
). Further studies
identified the increase in transmission loss as resulting from the conversion, within the foam,
of the incident plane wave into higher order symmetric waveguide modes that could not
radiate efficiently from the rear surface of the foam wedge. Although this characteristic may
be useful when good high frequency transmission loss is desired, the installation thickness of
a wedge makes this approach of only limited utility in most noise control applications.

However, it was found that a similar sound transmission loss increase could be achieved by
grading the tortuosity of a plane foam layer across the width (not the depth) of the channel.
For the configuration shown in Fig. 6(b), the tortuosity was decreased in piecewise-constant
steps from 7.8 at the center, to a minimum value in the foam adjacent to the hard walls: the
variation was symmetrical with respect to the duct center-line. The dimensions of the foam
layer were 5.4 cm 5.4 cm: i.e., the foam block had the same volume as the wedges. As
may be seen from Fig. 7(b), the high frequency transmission loss increased as the spatial
variation of tortuosity was made larger. Unlike the wedge case, there was not a decrease in
transmission loss at low frequencies in these cases. As was the case for the wedge, the axial
fluid displacement in the central and edge regions was found to be out-of-phase across the
width of the foam layer, thus accounting for the higher transmission. Therefore, it was
concluded that the phase variations across the width of the foam layer resulting from the
spatial variation of tortuosity caused the net axial volume velocity at the rear surface of the
plane foam layer to be minimized in certain frequency ranges, thus enhancing the layers
transmission loss. Extensive investigation of this case and further discussion can be found in
reference [56]. A finite element approach has also been used to guide the design of sound
absorption treatments having inhomogeneous porosity distributions [59].

As illustrated here, foam finite element techniques provide a powerful design tool in noise
control. With the availability of these new numerical tools, the full power of optimal design
procedures can be applied to the design of noise control treatments featuring foam elements.
8. CONCLUSIONS
It has been shown in this article that porous materials may be grouped into three categories,
i.e., rigid, limp and elastic, and that rigid and limp porous materials differ from elastic porous
materials in the number of wave types that can propagate within them. In addition, it has
been shown that the important acoustical properties of a porous material can be predicted
when its macroscopic properties such as porosity, flow resistivity, tortuosity, bulk density,
bulk Youngs shear and modulus of elasticity, are known. Acoustically, porous materials can
be characterized by reflection and absorption coefficients, surface impedance and
transmission coefficients, etc. It has also been shown that wave propagation in elastic porous
materials can be modeled accurately based on the Biots general elastic porous material
theory: in particular, theoretical and experimental results show good agreement in the case of
random incidence sound transmission loss through foam-lined double panel structures.
Finally, it has been demonstrated that foam finite element models can successfully be applied
to the optimal design of noise control treatments.

Although much progress has been made in this area, many issues that relate to porous noise
control materials remain to be resolved. For example, it would be desirable to pursue the
following topics: (i) development of more accurate and efficient techniques for measuring the
tortuosity and the bulk elastic properties of foams (including the loss factor); (ii)
development of theories that can connect a foams microstructure to its macroscopic
properties; (iii) systematic optimization of foam microstructure to yield specified acoustical
properties; (iv) development of SEA-compatible models for efficient prediction of the high
frequency behavior of foam elements; (v) use of porous materials in damping applications;
(vi) development of nonlinear models in the context of the Biot theory; (vii) development of
specialized models to allow increased calculation efficiency under special circumstances,
e.g., for limp materials; (viii) development of models to account for the anisotropy inherent
in most noise control materials; and (ix) development of models that explicitly account for
the inhomogeneity of typical noise control materials.
REFERENCES
[1] L.L. Beranek 1942 Journal of the Acoustical Society of America 13, 248-260. Acoustic Impedance of
Porous Materials.
[2] H.-Y. Lai, S. Katragadda, J.S. Bolton and J.H. Alexander 1997 Paper 972064 Proceedings of the 1997
Noise and Vibration Conference, Society of Automotive Engineers, 1553-1560. Layered Fibrous
Treatments for Sound Absorption and Sound Transmission.
[3] K.U. Ingard 1981 J. Eng. Industry, Trans. ASME 103, 302-313. Locally and Nonlocally Reacting Flexible
Porous Layers; A Comparison of Acoustical Properties.
[4] J.W.S. Rayleigh 1896 The Theory of Sound (Volume II, Article 351). New York: Dover Publication,
second edition, 1945, re-issue.
[5] K. Attenborough 1982 Physics Reports 82, 179-227. Acoustical Characteristics of Porous Materials.
[6] L.L. Beranek 1947 Journal of the Acoustical Society of America 19, 556-568. Acoustical Properties of
Homogeneous, Isotropic Rigid Tiles and Flexible Blankets.
[7] C. Zwikker and C.W. Kosten, 1949 Sound Absorbing Materials. Amsterdam: Elsevier Press.
[8] M.A. Biot 1956 Journal of the Acoustical Society of America 28, 168-191. Theory of Propagation of
Elastic Waves in a Fluid-Saturated Porous Solid. I. Low-Frequency Range. II. Higher Frequency Range.
[9] J.S. Bolton and E.R. Green 1993 Applied Acoustics 39, 23-51. Normal Incidence Sound Transmission
through Double-Panel Systems Lined with Relatively Stiff, Partially Reticulated Polyurethane Foam.
[10] J.S. Bolton, N.-M. Shiau and Y.J. Kang 1996 Journal of Sound and Vibration 191, 317-347. Sound
Transmission through Multi-Panel Structures Lined with Elastic Porous Materials.
[11] N.-M. Shiau 1991 Ph.D. Thesis, School of Mechanical Engineering, Purdue University. Multi-
Dimensional Wave Propagation in Elastic Porous Materials with Application to Sound Absorption,
Transmission and Impedance Measurement.
[12] W. Tsoi 1992 MSME Thesis, School of Mechanical Engineering, Purdue University. Acoustical Modeling
of Polyimide Foams.
[13] K. Attenborough 1971 Journal of Sound and Vibration 16, 419-442. The Influence of Microstructure on
Propagation in Porous Fibrous Absorbents.
[14] J. Sides, K. Attenborough and K.A. Mulholland 1971 Journal of Sound and Vibration 19, 49-64.
Application of a Generalized Acoustic Propagation Theory to Fibrous Absorbents.
[15] Y. Kawasima 1960 Acustica 10, 208-217. Sound Propagation in a Fiber Block as a Composite Medium.
[16] Y.J. Kang 1994 Ph.D. Thesis, School of Mechanical Engineering, Purdue University. Studies of Sound
Absorption by and Transmission through Layers of Elastic Noise Control Foams: Finite Element Modeling
and Effects of Anisotropy.
[17] H.L. Kuntz 1982 NASA CR-167979. High Intensity Sound in Air Saturated Fibrous Bulk Porous Materials.
[18] J.D. McIntosh and R.F. Lambert 1990 Journal of Acoustical Society of America 88, 1939-1949. Nonlinear
Wave Propagation through Rigid Porous Material. I: Nonlinear Parameterization and Numerical Solutions.
[19] R.F. Lambert and J.S. McIntosh 1990 Journal of Acoustical Society of America 88, 1950-1959. Nonlinear
Wave Propagation through Rigid Porous Material. II: Approximate Analytical Solutions.
[20] A. Cunningham, E. Huyghens and J.W. Leenslag 1995 Journal of Cellular Plastics 31, 461-472. MDI
Comfort Cushioning for Automotive Applications.
[21] D.A. Bies 1988 Chapter 10 in Noise and Vibration Control, Ed. L.L. Beranek (Institute of Noise Control
Engineering). Acoustical Properties of Porous Materials.
[22] Y. Champoux, M.R. Stinson and G.A. Daigle 1991 Journal of the Acoustical Society of America 89, 910-
916. Air-Based System for the Measurement of Porosity.
[23] L.L. Beranek 1949 Acoustic Measurements, 854-857. New York: Wiley.
[24] J.F. Allard 1993 Propagation of Sound in Porous Media: Modeling Sound Absorbing Materials. London:
Elsevier Press.
[25] ASTM Standards C552-87. Standard Test Methods for Airflow Resistance of Acoustical Materials.
[26] Y. Champoux and M.R. Stinson 1992 Journal of the Acoustical Society of America 92, 1120-1131. On
Acoustical Models for Sound Propagation in Rigid Frame Porous Materials and the Influence of Shape
Factors.
[27] R.J.S. Brown 1980 Geophysics 45, 1269-1275. Connection between Formation Factor of Electrical
Resistivity and Fluid-Solid Coupling Factor in Biots Equations for Acoustics Waves in Fluid-Filled Porous
Media.
[28] M. Henry, D. Lafarge and B. Castagnede 1994 Proceedings of INTER-NOISE 94, 1935-1938. Ultrasonic
Measurement of Tortuosity on Acoustic Porous Materials.
[29] Y.K. Kim and H.B. Kingsbury 1979 Experimental Mechanics 19, 252-258. Dynamic Characterization of
Poroelastic Materials.
[30] A. Wijesinghe and H.B. Kingsbury 1979 Journal of the Acoustical Society of America 65, 90-95. Complex
Modulus of a Poroelastic Column.
[31] A. Okuno 1986 Ph.D. Thesis, School of Mechanical Engineering, University of Delaware. Dynamic
Response of Structures Containing Poroelastic Materials.
[32] U. Ingard, F. Kirschner, J. Koch and M. Poldino 1989 Proceedings of INTER-NOISE 89, 1057-1062.
Sound Absorption by Porous, Flexible Materials.
[33] U. Ingard, F. Kirschner, J. Koch and M. Poldino 1990 Proceedings of INTER-NOISE 90, 225-230. Further
Studies of Sound Absorption by Porous, Flexible Materials.
[34] Y.J. Kang 1995 Personal Communication.
[35] E. Mariez, S. Sahraoui and J.F. Allard 1996 Proceedings of INTER-NOISE 96, 951-954. Elastic Constants
of Polyurethane Foams Skeleton for Biot Model.
[36] A.D. Pierce 1981 Acoustics-An Introduction to Its Physical Principles and Applications. New York:
McGraw-Hill. 1989 re-published by Acoustical Society of America.
[37] M. Minten, A. Cops and W. Lauriks 1988 Journal of Sound and Vibration 120, 499-510. Absorption
Characteristics of an Acoustic Material at Oblique Incidence Measured with Two-Microphone Technique.
[38] R.F. Lambert 1982 Journal of the Acoustical Society of America 73, 1131-1138. Propagation of Sound in
Highly Porous Open-Cell Foams.
[39] R.F. Lambert 1982 Journal of the Acoustical Society of America 73, 1139-1146. Surface Acoustic
Admittance of Highly Porous Open-Cell Foams.
[40] J.F. Allard, A. Aknine and C. Depollier 1986 Journal of the Acoustical Society of America 79, 1734-1740.
Acoustical Properties of Partially Reticulated Foams with High and Medium Flow Resistance.
[41] P.M. Morse 1948 Vibration and Sound. New York: McGraw-Hill. 1981 re-published by Acoustical
Society of America.
[42] D.A. Bies and C.H. Hansen 1988 Engineering Noise Control: Theory and Practice. Boston: Unwin
Hayman.
[43] M.E. Delany and E.N. Bazley 1969 National Physical Laboratory, Aerodynamics Division Report AC37.
Acoustical Characteristics of Fibrous Absorbent Materials.
[44] L.L. Beranek and S. Labate 1960 Chapter 12 in Noise Reduction, Ed. L.L. Beranek, Second Reprint Edition
(Los Altos: Penisula Press). Acoustical Properties of Porous Materials.
[45] A. Cops and M. Minten 1984 Noise Control Engineering Journal 22, 104-111. Comparative Study
between the Sound Intensity Method and the Conventional Two-Room Method to Calculate the Sound
Transmission Loss of Wall Constructions.
[46] M.J. Crocker, P.K. Raju and B. Forssen 1981 Noise Control Engineering Journal 17, 6-11. Measurement
of Transmission Loss of Panels by Direct Determination of Transmitted Acoustic Intensity.
[47] L.L. Beranek and G.A. Work 1949 Journal of the Acoustical Society of America 21, 419-428. Sound
Transmission through Multiple Structures Containing Flexible Blankets.
[48] R.L. Wallace, Jr., H.F. Dienel and L.L. Beranek 1946 Journal of the Acoustical Society of America 18, 246.
Measurement of the Transmission of Sound through Lightweight Structures.
[49] F.J. Fahy 1987 Sound and Structural Vibration: Radiation, Transmission and Response. New York:
Academic Press, Inc.
[50] B.H. Song and J.S. Bolton 2000 Journal of the Acoustical Society of America 107, 1131-1152. A Transfer
Matrix Approach for Estimating the Characteristic Impedance and Wave Number of Limp and Rigid
Porous Materials.
[51] J.H.B. Zarek 1978 Journal of Sound and Vibration 61, 205-234. Sound Absorption in Flexible Porous
Materials.
[52] W. Lauriks, A. Cops, J.F. Allard, C. Depollier and P. Rebillard 1990 Journal of the Acoustical Society of
America 87, 1200-1206. Modelization at Oblique Incidence of Layered Porous Materials with Impervious
Screens.
[53] Y.J. Kang and J.S. Bolton 1995 Journal of the Acoustical Society of America 98, 635-643. Finite Element
Modeling of Isotropic Elastic Porous Materials Coupled with Acoustical Finite Elements.
[54] Y.J. Kang and J.S. Bolton 1996 Journal of the Acoustical Society of America 99, 2755-2765. A Finite
Element Model for Sound Transmission through Foam-Lined Double Panel Structures.
[55] Y.J. Kang and J.S. Bolton 1996 A.S.M.E. Journal of Vibration and Acoustics 118, 498-504. Optimal
Design of Acoustical Treatments.
[56] Y.J. Kang and J.S. Bolton 1997 Journal of the Acoustical Society of America 102, 3319-3332. Sound
Transmission through Elastic Porous Wedges and Foam Layers Having Spatially Graded Properties.
[57] S. Rigobert, N. Atalla and F.C. Sgard 2003 Journal of the Acoustical Society of America 114, 2607-2617.
Investigation of the Convergence of the Mixed Pressure-Displacement Formulation for Three-Dimensional
Poroelastic Materials Using Hierarchical Elements.
[58] V. Easwaran, W. Lauriks and J.P. Coyette 1996 Journal of the Acoustical Society of America 100, 2989-
3002. Displacement-Based Finite Element Method for Guided Wave Propagation Problems: Application to
Poroelastic Media.
[59] F.C. Sgard, X. Olny, N. Atalla and F. Castel 2005 Applied Acoustics 66, 625-651. On the Use of
Perforations to Improve the Sound Absorption of Porous Materials.








xy

yx

y

s

xy

yx

s

s

s

y

x
solid phase fluid phase
x
y

v
y

yx

s
y

jU
y
ju
y
(a)
P

ju
x
x
y

v
y

w
p

s

y

jU
y

ju
y

jw
y

yx
(b)
P
foam
hard wall
air
c d
L
constrained edges
x
y
rigid piston

u
o
e
jt

a
Figure 3: Measured (solid line) and predicted
(dashed line) transmission loss for foam-
lined double panel systems: (a) Bonded-
Bonded (BB) configuration and (b)
Unbonded-Unbonded (UU) configuration.

Figure 4: Configuration for shape optimization of
foam wedge.
Figure 1: Sign convention for stresses acting on
the solid and fluid phases of foam.
Figure 2: Boundary conditions for: (a) open
surface and (b) bonded Euler-Bernoulli
plate.






0.0 0.
0.2 0.2
0.4 0.4
0.6 0.
0.8 0.
1.0 1.

0 500 1000 1500 2000


= 36
o
(optimal wedge)
Frequency (Hz)
= 132
o
= 180
o
(a)
0.0 0
0.2 0
0.4 0
0.6 0
0.8 0
1.0

16 28 36 41 48 59 74 97 132 180
wedge tip angle ()
(b)
0 0
10 10
20 20
30 30
40 40
10
2
T
r
a
n
s
m
i
s
s
i
o
n

L
o
s
s

(
d
B
)
10
3
180
o
132
o
97
o
74
o
59
o
Frequency (Hz)
48
o
41
o
36
o
28
o
(a)
Figure 6: System configurations for calculation
of sound transmission through (a) a wedge
having uniform properties, and (b) a plane
foam layer (L = 5.4 cm, d = 5.4 cm) having
a spatial variation of tortuosity.
Figure 5: (a) Absorption coefficient vs. frequency and (b) frequency-averaged absorption coefficient vs.
wedge tip angle.
0 0
10 10
20 20
30 30
40 40
10
2
T
r
a
n
s
m
i
s
s
i
o
n

L
o
s
s

(
d
B
)
10
3
case0 (7.8-7.8)
case1 (7.8-7.0)
case2 (7.8-6.0)
case3 (7.8-5.0)
case4 (7.8-4.0)
Frequency (Hz)
case5 (7.8-3.0)
case6 (7.8-2.0)
case7 (7.8-1.0)
(b)
Figure 7: Sound transmission through (a) a
wedge having uniform properties, and
(b) a plane foam layer having a spatial
variation of tortuosity.

foam
lubricated edges
air
d
air
x
y
hard wall
lubricated edges
air
d
air
x
y
(a)
(b)
z
n
=

0
c
z
n
=

0
c
hard wall
foam

u
0
e
jt

u
0
e
jt

You might also like