You are on page 1of 18

A modal-based reduction method for sound absorbing porous

materials in poro-acoustic finite element models


Romain Rumplera) and Jean-François Deu

Structural Mechanics and Coupled Systems Laboratory (LMSSC), Conservatoire National des Arts et M
etiers
(CNAM), Mechanics, Case 2D6R10, 2 Rue Cont e, 75003 Paris, France

€ransson
Peter Go
The Marcus Wallenberg Laboratory for Sound and Vibration Research (MWL), Department of Aeronautical
and Vehicle Engineering, KTH Royal Institute of Technology, School of Engineering Sciences, SE-100 44,
Stockholm, Sweden

(Received 19 July 2011; revised 14 August 2012; accepted 20 August 2012)


Structural-acoustic finite element models including three-dimensional (3D) modeling of porous
media are generally computationally costly. While being the most commonly used predictive tool
in the context of noise reduction applications, efficient solution strategies are required. In this work,
an original modal reduction technique, involving real-valued modes computed from a classical
eigenvalue solver is proposed to reduce the size of the problem associated with the porous media.
In the form presented in this contribution, the method is suited for homogeneous porous layers. It is
validated on a 1D poro-acoustic academic problem and tested for its performance on a 3D applica-
tion, using a subdomain decomposition strategy. The performance of the proposed method is esti-
mated in terms of degrees of freedom downsizing, computational time enhancement, as well as
matrix sparsity of the reduced system. VC 2012 Acoustical Society of America.

[http://dx.doi.org/10.1121/1.4750496]
PACS number(s): 43.55.Ka, 43.50.Gf, 43.20.Jr, 43.55.Wk [NAG] Pages: 3162–3179

I. INTRODUCTION amount of work has been done using numerical element


methods, and particularly the finite element method (FEM),
Use of porous media as a passive solution for reduction
which is considered in this work. Being well suited for prob-
of noise in vehicles has been extensively studied over the
lems in the low frequency range, several formulations have
past two decades. Its interest lies in the strong coupling
been proposed among which a solid and fluid phase displace-
effects between the solid phase, or frame, and the fluid con-
ments formulation,12–14 a mixed solid displacement and fluid
tained in its open cells, which makes it appropriate for
pressure fluctuation formulation,15 which have been com-
acoustic applications. Modeling-wise, the theory of wave
pleted by rigorous energy expressions in Ref. 16. More
propagation in sound absorbing porous materials of Biot and
recently, a solid phase and total displacements formulation
Allard1–4 has emerged as a reference to describe the physics
was established.17 The main drawback pointed out for this
involved. When it comes to design, efficient predictive tools
method is its high computational cost due to the need to
are needed in order to allow industrial applications to be per-
refine the discretization with, e.g., increasing geometric
formed at a reasonable cost, i.e., time and computational
complexity, or frequency range. Modeling poroelastic mate-
resources. Therefore, several methods have been proposed in
rials requires particularly refined meshes, as shorter wave-
order to solve the poroelastic equations of the Biot–Allard
lengths are involved when compared to more classic
theory. Among them, use of frequency-dependent impedan-
conservative acoustic applications. In addition, the generated
ces to represent porous layers5–7 has proved to be extremely
system of linear equations is complex and frequency-
efficient. However, it has for a major drawback its restriction
dependent, which implies an increased amount of resources
to locally reacting interfaces, or the necessity to proceed to a
needed to solve the problem, both in terms of computational
case-dependent identification step which is not an option in
time and memory allocation. This issue was partly improved
the scope of predictive methods. More recently, some works
when the mixed displacement-pressure formulation was
have been initiated on applying the wave-based method8,9 to
introduced, reducing the number of degrees of freedom
modeling of porous foams.10 Its demonstrated computational
(DOF) per node from 6, for the displacement formulations,
efficiency, as well as its applicability over a wide frequency
to 4. Hierarchical elements have also been proposed in that
range makes it a potential powerful tool. Investigations are
sense,18,19 limiting the number of elements needed for dis-
nevertheless still in progress in order to propose extension to
cretization while offering the additional possibility to model
three-dimensional (3D) applications including poroelastic
thinner poroelastic layers than possible with standard linear
materials, and to overcome some of its limitations such as
elements. Modal methods represent another approach for
complex geometry modeling.11 Besides these, a predominant
reduced models which can be used in the scope of widely
spread linear FEM. In Ref. 20, generalized complex modes
a)
Author to whom correspondence should be addressed. Also at: MWL–KTH were used to reduce poroelastic domains described with
Stockholm, Sweden. Electronic mail: rumpler@kth.se the mixed displacement-pressure formulation, which were

3162 J. Acoust. Soc. Am. 132 (5), November 2012 0001-4966/2012/132(5)/3162/18/$30.00 C 2012 Acoustical Society of America
V

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms
shown difficult to apply in an industrial context. The solid and
fluid phase displacements formulation was used in Refs. 21
and 22 to iteratively derive an appropriate set of basis vectors
to reduce the size of the problem to solve, stating the esti-
mated performance of the respective methods in terms of
number of DOF and precision. More recently, Dazel et al.23,24
used their 6 DOF per node formulation17 to propose resolution FIG. 1. Description and notations of the poro-acoustic interaction problem.
methods based on normal modes for the poroelastic equations.
However, while their displacement-based formulation simpli- primary variables for the porous media (Sec. II A 2). The
fies the coupling of porous domains to other domains, domains boundaries are separated into contours of
the associated modal-based resolution methods were either
suitable under the assumption of neglected shear in the porous (1) imposed Dirichlet boundary conditions denoted @1 XF
material,23 or demonstrated on 1D applications, thus not and @1 XP ,
involving shear waves.24 (2) prescribed Neumann boundary conditions denoted @2 XF
In the present work, a modal approach based on stand- and @2 XP ,
ard real-valued eigenvalue solvers is proposed to reduce (3) coupling interface between acoustic fluid and porous
poro-acoustic applications using the solid and fluid phase media (CFP ).
displacements poroelastic formulation.12,13 In its present The FE formulation is presented for a permanent har-
form, it is presented for reduction of 3D homogeneous monic response at angular frequency x.
porous layers. The first section introduces the general poro-
acoustic problem used, as well as the corresponding finite
element (FE) formulation for harmonic excitations. The A. Dynamic equations and constitutive laws
poroelastic equations are rewritten focusing on the possibil- 1. Compressible fluid (p)
ity to associate a standard eigenvalue problem. In the fol-
lowing section, the proposed modal reduction of porous The internal fluid within cavities is assumed compressi-
media is presented, included in a substructuring decomposi- ble, perfect, satisfying the Helmholtz equation derived from
tion strategy. The emphasis being rather to establish the the motion, continuity, and constitutive equations
modal approach feasibility than to investigate the chosen
decomposition strategy and its performance, the porous do- x2
Dp þ p¼0 in XF ; (1)
main only is reduced. The method being general, it can how- c20
ever be included in all forms of classically used modal
synthesis approaches and decomposition strategies,25 whose where c0 is the constant speed of sound in the fluid and p the
choices will affect the efficiency for case-dependent config- pressure fluctuation scalar field. The limit case x ¼ 0 is not
urations. Such an approach, investigating the reduction of given by Eq. (1). Though not considered in this work, the so-
the acoustic domain in poro-elasto-acoustic applications, lution is given by the static solution of the coupled fluid-
can be found in a previous contribution by the authors.26 structure problem.27,28
The third section of this work is dedicated to numerical
applications. A unidimensional example is first investigated
2. Porous media Biot theory (us ; uf )
to establish the potential of the proposed reduction in terms
of accuracy and DOF downsizing. Then, a 3D academic At angular frequency x, the poroelastic medium satis-
application allows one to estimate the performances in terms fies the following elastodynamic linearized equations,
of precision, computational time, size, and sparsity of the derived in the Biot–Allard theory4 (see Table I for notations
reduced problem. While promising, it is showed that there is of the standard porous materials parameters), taking into
room for further improvements to take full advantage of the
reduction strategy when applied to multidimensional TABLE I. List of material parameters.
applications.
Notation Description

qs Density of the material constituting the frame


II. FINITE ELEMENT FORMULATION
ðk; lÞ Lame parameters for the solid frame
A poro-acoustic problem is considered, whose descrip- qf Ambient fluid density
tion and notations are presented in Fig. 1. The formulation, as g Ambient fluid viscosity
presented in this paper, is based on previous research pub- P0 Ambient fluid standard pressure
lished by the authors26 for poro-elasto-acoustic applications, c Heat capacity ratio for the ambient fluid
Pr Prandtl number for the ambient fluid
the main difference being its suitability for modal-based
/ Porosity
reduction of the porous domain. The acoustic fluid and the
a1 Tortuosity
porous medium occupy the domains XF and XP , respectively. r Static flow resistivity
The compressible fluid is described using pressure fluctuation K Viscous characteristic length
(p) as primary variable (Sec. II A 1), while fluid and solid K0 Thermal characteristic length
phases homogenized displacements (us , uf ) are retained as

J. Acoust. Soc. Am., Vol. 132, No. 5, November 2012 Rumpler et al.: Modal-based porous reduction 3163

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms
account inertia and viscous coupling effects between solid which, when introduced in Eq. (5), leads to the following
and fluid phases, expressions of the stress-strain relations using Voigt notation,
~
div rs  i x bðxÞðu s  uf Þ ð1Þ
rs ¼ Dð1Þ
s eðus Þ þ Dsf eðuf Þ
þ x2 ½ðð1  /Þ qs þ qa Þus  qa uf  ¼ 0 in XP ; (2a) ð2Þ
þ ðK~ f  P0 ÞðDð2Þ
s eðus Þ þ Dsf eðuf ÞÞ; (8a)
~
div rf  i x bðxÞðu f  us Þ ð1Þ ð1Þ
rf ¼ Dsf eðus Þ þ Df eðuf Þ
þ x2 ½qa us þ ð/qf þ qa Þuf  ¼ 0 in XP ; (2b)
ð2Þ ð2Þ
þ ðK~ f  P0 ÞðDsf eðus Þ þ Df eðuf ÞÞ; (8b)
where us and uf are, respectively, the solid phase and fluid
phase averaged displacements in the sense of Biot theory, / with
the porosity of the porous material, i.e., the volume fraction !
~
of fluid. bðxÞ (henceforth denoted b, ~ where the tilde refers ð1  /Þ2
Dð1Þ
s ¼ 2lD þ 0
kþ P0 D;
to a complex-valued quantity) and qa are, respectively, the /
complex frequency-dependent viscous drag and the inertia
ð1Þ
coupling parameter, based on the standard notations of mate- Dsf ¼ ð1  /Þ P0 D;
rial parameters introduced in Table I,4 and given by
ð1Þ
 1=2 Df ¼ / P0 D;
4ixa21 gqf
b~ ¼ r/2 1 þ ; (3)
r 2 K2 / 2 ð1  /Þ2
Dð2Þ
s ¼ D;
/
qa ¼ /qf ða1  1Þ: (4)
ð2Þ
Dsf ¼ ð1  /Þ D;
rs and rf are the averaged stress tensors for the solid and
fluid phases, respectively. They satisfy the Lagrangian ð2Þ
Df ¼ / D;
stress-strain relations developed by Biot. Under the assump-
tion of large values for the bulk modulus of the porous frame where
material when compared to the fluid in the pores, which is
2 3
typically the case in the acoustic applications with sound 1 0 0 0 0 0
2 3
absorbing materials, these stress-strain relations can be writ- 60 1 0 0 0 07 1 1 1 0 0 0
6 7
ten as functions of the complex-valued and frequency- 6 7 61 1 1 0 0 07
60 0 1 0 0 07 6 7
dependent equivalent bulk modulus K~ f ðxÞ, 6
6 1 7
7
6
6
7
0 0 7 and D ¼ 6 1 1 1 0 0 07
! D ¼6
0
6
0 0 0
7 60
7:
ð1  /Þ 2
6
2
7 6 0 0 0 0 077
rs ¼ 2~l eðus Þ þ k~ þ K~ f ðxÞ tr½eðus Þ1 6 1 7 6 7
/ 60 0 0 0 07 40 0 0 0 0 05
6 2 5 7
4 0 0 0 0 0 0
þ ð1  /Þ K~ f ðxÞ tr½eðuf Þ1; (5a) 1
0 0 0 0 0
2
rf ¼ ð1  /Þ K~ f ðxÞ tr½eðus Þ1 þ / K~ f ðxÞtr½eðuf Þ1; (9)
(5b)
For the sake of clarity and conciseness of the expressions
where 1 represents a unitary matrix, eðus Þ and eðuf Þ are the derived in the upcoming sections, the Lame parameters for the
strain tensors associated with the averaged displacements solid frame are considered real and frequency-independent.
vector fields us and uf , respectively, defined by Thus, no structural damping is taken into account in the porous
medium behavior. However, the method presented is equally
1
eðuÞ ¼ ðgrad u þ gradT uÞ: (6) valid and straightforward to establish when hysteretic struc-
2 tural damping in the solid frame is included. Thus, for a hyste-
Besides the standard material parameters presented in Table retic proportional structural damping parameter denoted gs ,
I, the dynamic bulk modulus K~ f ðxÞ (henceforth denoted K~ f ) the complex-valued Lame parameters are written as
is given in Ref. 4. In order to propose finite element poroe- ~k ¼ ð1 þ ig Þ k;
s (10a)
lastic equations well suited for an original modal-based
reduction approach (see Sec. III B for further details), its ~ ¼ ð1 þ igs Þ l:
l (10b)
expression is separated into its zero-frequency limit and
complex frequency-dependent behavior, When accounted for in the constitutive equations, Eqs. (8)
can be rewritten as
cP0
K~ f ¼ "  1=2 #1 ð1Þ
8g ixPrK02 qf rs ¼ ð1 þ igs ÞðDð1Þ
s eðus Þ þ Dsf eðuf ÞÞ
c  ðc  1Þ 1 þ 02

ixPrK qf 16g þ ðK~ f  ð1 þ igs ÞP0 ÞðDð2Þ
ð2Þ
s eðus Þ þ Dsf eðuf ÞÞ;
¼ P0 þ ðK~ f  P0 Þ; (7) (11a)

3164 J. Acoust. Soc. Am., Vol. 132, No. 5, November 2012 Rumpler et al.: Modal-based porous reduction

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms
ð1Þ ð1Þ
rf ¼ ð1 þ igs ÞðDsf eðus Þ þ Df eðuf ÞÞ 2. Finite element discretized problem
ð2Þ ð2Þ
þ ðK~ f  ð1 þ igs ÞP0 ÞðDsf eðus Þ þ Df eðuf ÞÞ: The test-function method is used to derive the varia-
(11b) tional formulation of the coupled problem. For this purpose,
the spaces of sufficiently smooth functions Cp , Cus and Cuf
Note that gs does not have to be a constant, and could for are introduced, associated with the field variables p, us ,
instance be a frequency-dependent function. and uf , respectively. Let dp, dus , duf be the frequency-
The original rewriting of the constitutive equations under independent test functions, associated with p, us , uf , respec-
the forms of Eqs. (8) and (11) is put to advantage in the varia- tively, and belonging to their respective admissible spaces
tional formulation as shown in Sec. II B 2. It is the key ena- Cp , Cus ¼ fdus 2 Cus jdus ¼ 0 or dus  nP ¼ 0 on @1 XP g, and
bling for the further development of a new modal approach. Cuf ¼ fduf 2 Cuf jduf  nP ¼ 0 on @1 XP g.
Equations (1), (12), and (14) lead to, 8 dp 2 Cp ,
B. Fluid-structure interaction problem ð ð
x2
In this section, boundary and coupling conditions are gradp  graddp dV  2 p dp dV
XF c0 X F
recalled for the poro-acoustic coupled problem presented in ð
Fig. 1, in order to establish the discretized FE problem. 2
 x qF ð1  /Þ us  n dp dR
CFP
1. Poro-acoustic coupling and boundary conditions ð ð
 x2 qF / uf  n dp dR ¼ x2 qF uFb dp dR: (17)
At external boundary of the acoustic fluid domain, rigid CFP @ 2 XF
cavity conditions are classically imposed by setting a free
pressure field (@1 XF ¼ ;). A harmonic excitation is pre- Equations (2a), (8a), and (13a) lead to, 8 dus 2 Cus ,
scribed via an acoustic source, ð ð
ð1Þ ð1Þ
grad p  n ¼ x2 qF uFb on @2 XF ; (12) tr½Ds eðus Þeðdus ÞdVþ tr½Dsf eðuf Þeðdus ÞdV
XP XP

where uFb is set to zero out of the acoustic source included in þ ðK~ f  P0 Þ tr½Dð2Þ
s eðus Þeðdus ÞdV
@2 XF . XP
Coupling at interface CFP is given by normal stress and ð 
ð2Þ
normal displacement continuity conditions between the þ tr½Dsf eðuf Þeðdus ÞdV
acoustic fluid and both fluid and solid phases of the porous XP
ð ð 
medium,
þ ixb~ us  dus dV uf  dus dV
rs n þ ð1  /Þp n ¼ 0 on CFP ; (13a) XP XP
ð ð 
rf n þ / p n ¼ 0 on CFP ; (13b)  x2 ðð1  /Þqs þqa Þus  dus dV qa uf  dus dV
XP XP
uF  n  ð1  /Þus  n  /uf  n ¼ 0 on CFP : (14) ð
 ð1  /Þ pn  dus dR ¼ 0: (18)
No external force is applied to the outer boundary of the po- CFP
rous medium beside at interface CFP . Therefore, @2 XP ¼ ; in
the considered problem. Finally, at external boundary @1 XP , Equations (2b), (8b), and (13b) lead to, 8 duf 2 Cuf ,
two types of boundary conditions can be prescribed, the po- ð ð
ð1Þ ð1Þ
rous material being considered either as sliding or bonded to tr½Df eðuf Þeðduf ÞdV þ tr½Dsf eðus Þeðduf ÞdV
XP XP
a rigid wall: ð
For a bonded porous layer on a rigid wall, the following ð2Þ
þ ðK~ f  P0 Þ tr½Df eðuf Þeðduf ÞdV
conditions arise—fixed displacement of the solid phase and XP
normal displacement of the fluid phase—and need to be ð 
ð2Þ
satisfied, þ tr½Dsf eðus Þeðduf ÞdV
XP
us ¼ 0 on @1 XP ; (15a) ð ð 
þ ixb~ uf  duf dV  us  duf dV
uf  nP ¼ 0 on @1 XP : (15b) XP XP
ð ð 
For a sliding porous layer on a rigid wall, the following con-  x2 ð/qf þqa Þuf  duf dV  qa us  duf dV
ditions arise—fixed normal displacement of the solid phase XP XP
and fluid phase—and need to be satisfied, ð
/ pn  duf dR ¼ 0: (19)
us  nP ¼ 0 on @1 XP ; (16a) CFP

uf  nP ¼ 0 on @1 XP ; (16b) After discretization of the various terms in Eqs. (17)–(19) by


the FE method and dividing Eq. (17) by qF , the following
where nP is the unit normal vector pointing outward from XP . matrix equation for the coupled problem is obtained:

J. Acoust. Soc. Am., Vol. 132, No. 5, November 2012 Rumpler et al.: Modal-based porous reduction 3165

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms
02 3 2 3 2 3
KF 0 0 0 0 0 0 0 0
B6 ð1Þ 7 6 ð2Þ 7 6 7
B6 ð1  /ÞATFs
@4 Kð1Þ
ss Ksf 7 ~ 6
5 þ ðK f  P 0 Þ4 0 Kð2Þ
ss Ksf 7 ~6
5þixb 4 0 Css Csf 7
5
ð1ÞT ð1Þ ð2ÞT ð2Þ
/ATFf Ksf Kff 0 Ksf Kff 0 CTsf Cff
2 31 2 3 2 2 3
MF ð1  /Þ /AFf P x UFb
6 7C 6 7 6 7
x2 6
4 0 Mss Msf 7 C6 7 6
5A 4 Us 5 ¼ 4 0 5:
7 (20)
0 MTsf Mff Uf 0

   
Bss Bsf Us
This non-symmetric formulation can be symmetrized for a BP ¼ and UP ¼ : (21)
Bsf Bff Uf
resolution in the frequency domain by dividing the acoustic
equation by x2 (x 6¼ 0). The interest of rewriting the porous Similarly, the coupling between the interface acoustic DOF
ð1Þ ð2Þ
medium formulation into four matrices [KP , KP , CP , and (subscript I) and the porous DOF (subscript P) is denoted
MP as introduced for Eq. (21)] is already partly visible. In
fact, it involves constant real-valued matrices which can be AIP ¼ ½ð1  /ÞAIs /AIf : (22)
assembled once, while only the complex and frequency-
dependent factors K~ f and b~ are recomputed at each fre- Consequently, for modal reduction purposes, matrix set of
quency increment. Eq. (20) can be written as
If hysteretic structural damping of the solid frame is ex-
plicitly included, this statement stands, considering that
damping can be introduced as a factor of the assembled mat-
ð1Þ ð2Þ
rices KP and KP , as mentioned in Sec. II A 2, Eq. (11).
In addition, the amount of memory used is the same as
using only two complex-valued and frequency-dependent
matrices, considering that the sparsity is unchanged.13 More
importantly, the main interest, underlined in this approach, is
the possibility to use such a formulation in the context of
modal reduction techniques.
and can be symmetrized by dividing the acoustic equations
(lines 1 and 2) by x2 (x 6¼ 0).

III. MODAL REDUCTION OF THE POROELASTIC


MEDIUM B. Modal reduction
A. Presentation of the proposed solution strategy From the proposed expression of the porous medium FE
The proposed reduction method is applied to the dissipa- problem, real-valued normal modes can be computed associ-
tive porous medium of a poro-acoustic coupled problem, ated with the conservative poroelastic eigenvalue problem
which is the costly part of the problem. For the sake of con-  
ð1Þ
ciseness, the case of a rigid acoustic cavity with a single po- KP  x2 MP / ¼ 0: (24)
rous layer on one wall is considered. Notations used are
presented in Fig. 2. It is supposed that the Dirichlet boundary conditions
The acoustic degrees of freedom (DOF) are separated ð1Þ
imposed to the porous medium result in a nonsingular KP
 and those at interface with
into internal ones (subscript I), matrix, therefore removing zero-frequency modes. A modal
the porous medium (subscript I). These notations allow easy reduction basis UPm is built, selecting m low frequency
extension of the method to problems with multiple interfa- modes. They are normalized with respect to the porous mass
ces.26 The coupled porous medium matrices in Eq. (20) are matrix MP so that
ð1Þ ð2Þ
now considered, involving four matrices KP , KP , CP , and
MP corresponding to the set of unknowns UP such that, for UTPm MP UPm ¼ 1m ; (25a)
ð1Þ ð2Þ
each matrix indexed by P, i.e., BP 2 fKP ; KP ; CP ; MP g,
ð1Þ
UTPm KP UPm ¼ Xm ; (25b)

where 1m is a unitary matrix of dimension m, and Xm a diag-


onal matrix of same size, with the m lowest eigenvalues of
Eq. (24) on its diagonal.
There are two key points that make a reduction method
FIG. 2. Problem description for modal reduction of porous media. computationally efficient, which are its ability to

3166 J. Acoust. Soc. Am., Vol. 132, No. 5, November 2012 Rumpler et al.: Modal-based porous reduction

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms
(1) converge rapidly to the expected solution when adding where 1n , Xn and 1o , Xo , jo , fo are diagonal matrices of re-
modes in the basis, thus allowing a subsequent reduction spective dimensions n and o, while jn and fn are non-
in the number of DOF, as well as a reasonable time allo- diagonal sparse square matrices of dimension n.
cated to the computation of the modes, There are several options for the choice of attachment
(2) preserve or improve the sparsity of the matrices after functions participating to the static correction for improved
projection, and ideally produce diagonal submatrices. convergence. In this work, the single degree of freedom
(DOF) per node associated with the acoustic domain is put
The former aspect will be examined on applications in
to advantage. Attachment functions are computed as the
Sec. IV. Regarding the second aspect, the sparsity of matrices ð1Þ
ð2Þ KP -static responses of the poroelastic domain to unit pres-
KP and CP after projection on the modal basis is fundamen-
ð1Þ sure successively imposed at each interface acoustic DOF,
tal to take advantage of the diagonal form of projected KP
 
and MP . The choices made for the discretization of porous ð1Þ 1I ð1Þ1
media, among which the original separation of the “static”
T
½ AIP KP  ¼ ½ 0  ) WPI ¼ KP ATIP : (27)
WPI
and “dynamic” part of the dynamic bulk modulus [see Eq. ð1Þ
(7)], are of prime importance in order to fulfill this require- Again, assumption is made that KP is not singular. If other-
ment. In fact, as will be shown in Sec. IV B, it results in wise, a shift in frequency using the mass matrix MP can be
ð2Þ set instead, to define pseudo-static attachment functions, but
sparse reduced KP and CP , and even orthogonality of some
modes with respect to these matrices. Therefore, after testing is not considered in this work.
the m retained modes for their orthogonal properties with The corresponding change of basis applied to system (23),
ð2Þ leaving acoustic DOF uncondensed, is then (the hat symbol
respect to KP and CP , they are separated into o “orthogonal”
(UPo ) and n “non-orthogonal” (UPn ) ones, such that refers to an approximate estimation of the reference solution)
  2 3
1n 0 2 3 2 3 P ^ I
T
½UPn UPo  MP ½UPn UPo  ¼ ^ I
P 1I 0 0 0
; (26a) 6P ^I 7
0 1o 4P ^I 5 ¼ 4 0 1I 0 0 56 7; (28)
4a ^ 5
  ^P
U 0 WPI UPn UPo n
T ð1Þ Xn 0 ^o
a
½UPn UPo  KP ½UPn UPo  ¼ ; (26b)
0 Xo
where a^ n and a
^ o are the modal coordinates vectors associ-
 
ð2Þ jn 0 ated with the selected “non-orthogonal” and “orthogonal”
½UPn UPo T KP ½UPn UPo  ¼ ; (26c) modes, respectively.
0 jo
Applying a change of basis (28) to a symmetrized Eq.
  (23) and premultiplying by the transpose of the transforma-
T f 0
½UPn UPo  CP ½UPn UPo  ¼ n ; (26d) tion matrix leads to the following reduced set of equations:
0 fo

ð1Þ ð2Þ BPIo ¼ WTPI BP UPo ¼ BTPoI :


where for BP 2 fKP ; KP ; CP ; MP g,
This reduction is established once as an initial step before
BPII ¼ WTPI BP WPI ; the frequency sweep. It can be further improved using
dynamic condensation of the “orthogonal” modal coordi-
BPIn ¼ WTPI BP UPn ¼ BTPnI ; nates, as presented in the following section.

J. Acoust. Soc. Am., Vol. 132, No. 5, November 2012 Rumpler et al.: Modal-based porous reduction 3167

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms
C. Further condensation of selected modal
coordinates
To take full advantage of the partly diagonal form of the
reduced matrices in Eq. (26a), the modal coordinates corre-
sponding to the “orthogonal” modes can be dynamically con-
densed on the interface DOF at each frequency increment.
^ o in Eq. (29) gives
First, the set of equations associated with a FIG. 3. (Color online) Acoustic cavity mesh and dimensions for 1D
problem.

~ o  x2 1o 1 KðdynÞ
^ o ¼  Xo þ ðK~ f  P0 Þjo þ i x bf
a ^
PoI P I ;
(30) Then, when introducing Eq. (30) in Eq. (29), a further
reduced set of equations can be obtained at the beginning of
where each frequency increment, reducing the number of DOF
without changing the sparsity of the system matrix for the
ðdynÞ ð2Þ ~ PoI  x2 MPoI :
KPoI ¼ ðK~ f  P0 ÞKPoI þ i x bC (31) remaining ones,

where the frequency-dependent dynamic contribution to the a time-harmonic plane wave, with normal incidence on a
ðdynÞ
interface DOF, KPII ; can be expressed as poroelastic layer of infinite lateral extent. The rigid cavity is
25 cm long and covered with a 5 cm-thick poroelastic layer at
ðdynÞ ðdynÞ T ~ o x2 1o 1 KP : ðdynÞ
KPII ¼KPoI ½ Xo þðK~ f P0 Þjo þix bf oI
its end. Appropriate Dirichlet boundary conditions are applied
(33) to the poroelastic domain to achieve unidimensional behavior,
thus allowing solid and fluid displacement in the direction
This contribution involving simple operations—matrix addi- normal to the layer only. In addition, it is considered bonded
tions, multiplications and a diagonal matrix inversion—it is to a rigid wall at its extremity [Eq. (15)]. Material parameters
reasonable to expect a substantially improved efficiency, of the foam are given in Table II. The mesh consists of 20 lin-
provided this additional step is worth doing, i.e. a sufficient ear elements for the acoustic domain, and 10 linear poroelas-
number of “orthogonal” modal coordinates is involved. tic elements in the thickness of the layer, making it a
satisfying mesh for the frequency range considered. The 1D
IV. APPLICATIONS AND RESULTS
TABLE II. Air and porous material parameters.
A. Modal reduction of the porous layer in a 1D
poro-acoustic application Frame Fluid Porous

1. Presentation and reference solution c0 ¼ 343 m/s / ¼ 0:96


k ¼ 905 357 Pa c ¼ 1:4 r ¼ 32 kN s/m4
The proposed modal-based reduction method for l ¼ 264 062 Pa Pr ¼ 0:71 a1 ¼ 1:7
poroelastic sound absorbing materials is first tested on a qs ¼ 30 kg/m3 qf ¼ 1:21 kg/m3 K ¼ 90 lm
unidimensional poro-acoustic academic example. As pre- g ¼ 1:84  105 N s/m2 K0 ¼ 165 lm
sented in Fig. 3, it consists of an acoustic cavity excited with

3168 J. Acoust. Soc. Am., Vol. 132, No. 5, November 2012 Rumpler et al.: Modal-based porous reduction

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms
Mean quadratic pressure in cavity (dB)

Mean quadratic pressure in cavity (dB)


130 130

120 120

110 110

100 100

90 90

Rigid acoustic cavity Rigid acoustic cavity


80 Porous damped cavity 80 Porous damped cavity
500 1000 1500 2000 2500 3000 500 1000 1500 2000 2500 3000
Frequency (Hz.) Frequency (Hz.)

(a) (b)
FIG. 4. (Color online) Mean quadratic pressure reference FRF, 1D problem, with for rigid cavity: (a) poroelastic volume not replaced by acoustic volume; (b)
poroelastic volume replaced by acoustic volume.

FE model thus results in 21 acoustic DOF and 20 poroelastic large volume of the porous layer when compared to the
DOF. Its simplicity is not suited for a performance estimation, dimensions of the cavity.
but is however a relevant starting point to establish the poten- Figures 4 clearly show the strong damping effect of the
tial interest and suitability of the proposed modal basis. sound absorbing foam. Considering first Fig. 4(a), it can be
For this purpose, an estimation of the absorption coeffi- partly interpreted in terms of change of boundary condition
cient of the porous layer is a useful indicator, and is calcu- for the acoustic cavity of constant volume. Thus, at the reso-
lated, for a plane wave at normal incidence, as nance frequencies, both a reduction of the sound pressure
level and a shift toward lower frequencies are observed. This


Z~n ðxÞ  q0 c0
2 is consistent with the classical observations made for damped
a abs ðxÞ ¼ 1 

~
; (34)
Z~n ðxÞ þ q0 c0
systems. Two additional remarks are worth mentioning: (i)
the apparent constant pressure reduction at very low frequen-
where Z~n ðxÞ is the normal acoustic impedance given by cies is mostly induced by the artificial volume increase of the
cavity when the porous layer is added [for large wavelengths,
P the particle velocity of air near boundaries is low, and the po-
Z~n ðxÞ ¼  ; (35) rous layer is then inefficient (see the absorption coefficient in
ix / uf  n þ ð1  /Þus  n Fig. 6)]; (ii) the porous layer exhibits a compressional reso-
nance around 1000 Hz, thus reducing the frequency shift of
with P the incident acoustic pressure, us and uf the vector the higher acoustic resonances compared to the first one.
displacement fields of the solid and fluid phases at the sur- Alternatively, Fig. 4(b) may be interpreted in terms of
face of the material, and n the vector normal to the surface, change of boundary condition with apparent reduction of the
pointing outward from the fluid domain. acoustic cavity volume. In fact, this apparent reduction of
More generally, the mean quadratic pressure in the cav- the acoustic cavity volume implies a global shift of resonan-
ity is given as an output, computed as ces toward higher frequencies. Due to the substantial change
0ð 1 of volume involved, this overcomes the frequency shift to-
P2 dV ward low frequencies generated by the damping boundary
B XF C conditions. However, this configuration better renders the
Lp ¼ 10 logB@ ð C;
A (36)
P0 2 dV very low-frequency reduction of acoustic pressure.
XF The nature of the compressional resonance of the porous
layer around 1000 Hz is detailed in Fig. 5. It corresponds to
and the reference solution of the problem, with and without a quarter-wavelength resonance of the solid-borne wave.
adjunction of the poroelastic layer, is plotted in Fig. 4. For This is confirmed by Fig. 5(b), as all points in the solid parti-
the conservative acoustic problem, the two following config- tion are moving in phase. The tortuosity and viscous cou-
urations can be considered for this 1D case: the poroelastic plings between the two phases imply strong inertial and
is replaced by an equivalent volume of acoustic fluid or not. viscous interactions. Thus, at this resonance frequency, the
The former case is more suited to a physical interpretation of solid and fluid partitions move with a small phase shift and
the damping effect of the porous layer at the resonances. The with a slightly greater amplitude for the fluid phase. The res-
latter is a more realistic representation of the practical conse- onant behavior of the porous frame would be damped if the
quence of adding a porous layer to an existing environment. structural damping of the frame was taken into account,
Of course, for this academic validation, the difference using the previously mentioned constitutive equations, Eq.
between the two configurations is amplified by the rather (11). This is illustrated in Fig. 6, where the absorption

J. Acoust. Soc. Am., Vol. 132, No. 5, November 2012 Rumpler et al.: Modal-based porous reduction 3169

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms
−6
x 10 0
Solid partition
Solid partition Fluid partition
1 Fluid partition −0.2

−0.4

Displacement amplitude (m)


0.8
−0.6

Phase (rad.)
0.6
−0.8

−1
0.4

−1.2
0.2
−1.4

0 −1.6
−0.05 −0.04 −0.03 −0.02 −0.01 0 −0.05 −0.04 −0.03 −0.02 −0.01 0
Position (m) Position (m)

(a) (b)
FIG. 5. (Color online) Response of the solid and fluid porous partitions at 1010 Hz: (a) displacement amplitude; (b) phase.

coefficient at normal incidence is plotted for both cases (the and fluid phases may move in or out of phase according to
structural damping parameter is then gs ¼ 0:1). However, the type of wave and to the frequency.
whether the structural damping is included or not has no One possible shortcoming of the generated coupled
impact on the results presented for the modal approach pro- eigenmodes could be the inability to properly render the cou-
posed, and is therefore not further emphasized. pling effects between the two phases. Both the constitutive
coupling and the inertial coupling are taken into account in
the eigenvalue problem, but the viscous drag is not, even
2. Reduced problem
though viscous and inertial effects are linked by the tortuosity
The modal-based reduction method presented in Section dependency. For instance, the limit case, where the tortuosity
III, included in a substructuring approach for validation pur- (a1 ) is set to 1 (i.e., straight pores), implies no inertial cou-
poses, is applied to reduce the poroelastic part of the 1D pling while viscous coupling exists. This might involve inac-
application. The coupled solid-fluid shapes of the first 6 curacies in the reduced model. Possible corrections could
modes generated by the eigenvalue problem (24) are pre- involve including viscous effects in the eigenvalue problem
sented in Fig. 7. Their corresponding eigenfrequencies are at the cost of a loss of diagonality for the reduced mass ma-
944, 1136, 2856, 3435, 4839, and 5819 Hz. It is interesting trix, or to generate additional vectors to be included in the ba-
to notice that they can be paired, and that for each couple, sis from the computed modes and the viscous terms.29
one mode exhibits a standing wave in phase in the solid and However, the tests ran for a wide range of material properties
fluid partitions, while in antiphase for the other mode. This and for the considered frequency range have shown no such
is in good agreement with the expected behavior of waves restriction so far.
propagating in air-filled poroelastic materials, since the solid The convergence of the reduced model toward the ref-
erence solution is presented in Figs. 8 for 1 to 4 modes
included in the reduction basis. A satisfactory level of preci-
1 sion is achieved in the frequency range of interest, thus
0.9 downsizing the number of poroelastic DOF from 20 to 4.
Regarding sparsity of the reduced matrices, the paired
0.8 ð2Þ
modes observed in Fig. 7 also imply KP - and CP -coupled
Absorption coefficient

0.7 contributions. Thus, although not diagonal, the system ma-


0.6 trix corresponding to the modal coordinates still benefits
from a much improved sparsity when compared to the unre-
0.5
duced problem. While CPU time performance is irrelevant
0.4 for this application, both the reduction in size of the prob-
0.3
lem and the sparsity improvements suggest substantial effi-
ciency of the proposed reduction.
0.2 The convergence rate is further tested on another usual
0.1 Without hysteretic structural damping response output. Thus, Fig. 9 presents the convergence of
With hysteretic structural damping the calculated absorption coefficient with successive pairs of
0
0 500 1000 1500 2000 2500 3000 modes included in the basis. The plot on a linear scale offers
Frequency (Hz)
better insight into the quality of the solution, which is con-
FIG. 6. (Color online) Absorption coefficient at normal incidence, impact of firmed to progressively converge toward the reference solu-
the structural damping (gs ¼ 0:1). tion. It can be noted that the successive pairs contribute in

3170 J. Acoust. Soc. Am., Vol. 132, No. 5, November 2012 Rumpler et al.: Modal-based porous reduction

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms
FIG. 7. (Color online) First six coupled mode shapes of the 1D porous layer (a)–(f), solid (top) and fluid (bottom) phases.

increasing frequency regions of the frequency range, as perfectly match the reference on the upper frequency range
observed in standard modal synthesis methods for conserva- after four pairs included in the basis. In a situation where a
tive problems. Furthermore, on this refined check of the con- refined convergence is sought, this modal truncation effect is
vergence, it appears that the reduced solution does not traditionally compensated by static corrections added to the

130 110

125
Mean quadratic pressure in cavity (dB)

Mean quadratic pressure in cavity (dB)

105
120

115
100
110

105 95

100
90
95

90 85
85
unreduced porous damped pb unreduced porous damped pb
reduced porous damped pb 80 reduced porous damped pb
80
500 1000 1500 2000 2500 3000 500 1000 1500 2000 2500 3000
Frequency (Hz.) Frequency (Hz.)

(a) (b)
110 110
Mean quadratic pressure in cavity (dB)

Mean quadratic pressure in cavity (dB)

105 105

100 100

95 95

90 90

85 85

unreduced porous damped pb unreduced porous damped pb


80 reduced porous damped pb 80 reduced porous damped pb
500 1000 1500 2000 2500 3000 500 1000 1500 2000 2500 3000
Frequency (Hz.) Frequency (Hz.)

(c) (d)
FIG. 8. (Color online) Mean quadratic pressure FRF. Convergence of the reduced model to the reference solution with added modes in the basis: (a) one
mode, (b) two modes, (c) three modes, (d) four modes.

J. Acoust. Soc. Am., Vol. 132, No. 5, November 2012 Rumpler et al.: Modal-based porous reduction 3171

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms
1 1

0.9 0.9

0.8 0.8

Absorption coefficient

Absorption coefficient
0.7 0.7

0.6 0.6

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 Reference solution 0.1 Reference solution


Reduced porous Reduced porous
0 0
500 1000 1500 2000 2500 3000 500 1000 1500 2000 2500 3000
Frequency (Hz.) Frequency (Hz.)

(a) (b)
1 1

0.9 0.9

0.8 0.8
Absorption coefficient

Absorption coefficient
0.7 0.7

0.6 0.6

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 Reference solution 0.1 Reference solution


Reduced porous Reduced porous
0 0
500 1000 1500 2000 2500 3000 500 1000 1500 2000 2500 3000
Frequency (Hz.) Frequency (Hz.)

(c) (d)
FIG. 9. (Color online) Absorption coefficient; Convergence of the reduced model to the reference solution with added modes in the basis: (a) two modes, (b)
four modes, (c) six modes, (d) eight modes.

2 3
modal basis. Such an approach for this 1D case is presented ^
2 3 2 3 P I
in the following section. ^ I
P 1I 0 0 0 0 6P^ 7
6 I 7
4P
^I 5 ¼ 4 0 56 ^ sta 7
1I 0 0 0 6a 7:
^P
U 0 WPI WPsta UPn UPo 4 ^n 5
a
3. Improved precision with use of static response ^o
a
Although it may be argued that satisfactory accuracy is (38)
achieved by the proposed method when applied as is, the
simplicity of the considered 1D application allows one to
demonstrate the beneficial contribution of orthogonalized The improvements in precision are shown in Fig. 10 where
“static” response added to the modal basis for convergence the sound pressure level difference (in dB) to the reference so-
ð1Þ
purposes. For the 1D case, it is computed as the KP -static lution is plotted for 3–6 normal modes included in the modal
response of the poroelastic domain to a unitary plane wave basis (note the different scale on the quadratic pressure differ-
with normal incidence. Application of the superposition ence axis for each plot). The convergence is clearly enhanced
principle implies that it can be directly obtained by summa- both in terms of dispersion (frequency shift) and level error.
ð1Þ
tion of the poroelastic responses to successive interface In addition, the orthonormalized KP -static contribution fol-
acoustic DOF excitation, as detailed for general 3D attach- lows the sparsity properties of the 1D eigenmodes in the sense
ð2Þ
ð1Þ
ment functions [Eq. (27)]. Thus, the KP -static response is that it is KP - and CP -orthogonal if eigenmodes are included
related to the attachment functions: by matching pairs, and coupled to the eigenmodes included
without their matching pair. Notably, for the 1D application,
ð2Þ
WPstai ¼ Rj WPI ij : (37) it is KP - and CP -orthogonal if an even number of eigenmo-
des is selected for the modal basis.
WPsta is by construction orthogonal to the attachment func- Similarly to the previous section, Fig. 11 presents a
tions. A Gram–Schmidt procedure is used to orthogonalize it refined estimate of the convergence using the absorption coef-
with respect to the selected eigenmodes and it is then MP -nor- ficient. As the convergence is much improved compared to
malized. The corresponding applied transformation is then Fig. 9, the plot for four pairs of modes included in the basis is

3172 J. Acoust. Soc. Am., Vol. 132, No. 5, November 2012 Rumpler et al.: Modal-based porous reduction

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms
0.8 0.25
No static contribution No static contribution
With static contribution With static contribution
0.7

Quadratic pressure difference (dB)

Quadratic pressure difference (dB)


0.2
0.6

0.5
0.15

0.4

0.1
0.3

0.2
0.05
0.1

0 0
0 500 1000 1500 2000 2500 3000 0 500 1000 1500 2000 2500 3000
Frequency (Hz.) Frequency (Hz.)

(a) (b)
0.25 0.16
No static contribution No static contribution
With static contribution With static contribution
0.14
Quadratic pressure difference (dB)

Quadratic pressure difference (dB)


0.2
0.12

0.1
0.15

0.08

0.1
0.06

0.04
0.05
0.02

0 0
0 500 1000 1500 2000 2500 3000 0 500 1000 1500 2000 2500 3000
Frequency (Hz.) Frequency (Hz.)

(c) (d)
FIG. 10. (Color online) Error to reference solution (dB difference) with and without orthogonalized static response, including (a) three modes, (b) four modes,
(c) five modes, (d) six modes.

replaced by a plot corresponding to 1 mode and the static cor- B. Performance and limitations associated with the 3D
rection [Fig. 11(a)]. The solution is perfectly matching up to case
half the frequency range of interest with only one pair of
1. Presentation and reference solution
modes included in the basis. The use of two pairs makes it
suitable for the entire frequency range, and perfectly match- The extension to a 3D problem and its performance esti-
ing if three pairs are included. Such performance on a 1D val- mation is tested on a dissipative poro-acoustic example
idation case is in agreement with recent contributions in the initially proposed in Ref. 21, also used in Ref. 26 by the
literature made on this topic,23,24 limited to configurations authors. It consists of a 3D hexahedric acoustic cavity of
where shear waves are not involved, i.e., 1D problems or with dimensions 0:4  0:6  0:75 m3 (see Fig. 12), with rigid
porous materials for which shear can be neglected. For walls, and filled with air. One wall is covered with a 5 cm-
instance, in Ref. 24, it was showed that two real-valued thick porous layer. The low frequency behavior is tested
modes were required in order to properly render the absorp- applying a harmonic volume velocity source [Eq. (12)] at a
tion coefficient frequency dependence over the first porous corner of the cavity opposite the layer.
resonance. Furthermore, the traditionally used truncation rule The cavity is discretized by a 8  12  15 mesh of 8-
of thumb for modal synthesis (keeping modes with eigenfre- node hexahedric elements with pressure as single degree of
quencies lower than 1.5 to 2.5 times the highest frequency of freedom per node. The porous material, identical to the foam
interest) applies on the 1D configurations tested. used in Sec. IV A , is discretized by a 8  12  5 mesh of 8-
One originality of the present approach lies in the fact node hexahedric elements (Fig. 12), with 6 DOF per node
that no assumption is made for the projection of the porous corresponding to the fluid and solid phase displacements.
problem on the calculated coupled poroelastic modes. Thus, Sticking Dirichlet boundary conditions are applied to the po-
it offers the possibility to handle problems were shear effects rous foam face in contact with the covered wall, and sliding
are accounted for. The following section initiates a direct conditions are prescribed on the side faces [Eqs. (15) and
extension of the approach to such problems. It illustrates the (16)]. This leads to an FE model with 1872 acoustic DOF, and
potential as well as the limitations in the present stage of de- 3070 poroelastic DOF. The mean quadratic pressure fre-
velopment, on a 3D configuration, thus offering perspectives quency response in the acoustic cavity is given as an output.
for further improvements. The reference solution is plotted in Fig. 13 up to 1000 Hz,

J. Acoust. Soc. Am., Vol. 132, No. 5, November 2012 Rumpler et al.: Modal-based porous reduction 3173

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms
1 1

0.9 0.9

0.8 0.8

Absorption coefficient

Absorption coefficient
0.7 0.7

0.6 0.6

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 Reference solution 0.1 Reference solution


Reduced porous Reduced porous
0 0
500 1000 1500 2000 2500 3000 500 1000 1500 2000 2500 3000
Frequency (Hz.) Frequency (Hz.)

(a) (b)
1 1

0.9 0.9

0.8 0.8
Absorption coefficient

Absorption coefficient
0.7 0.7

0.6 0.6

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 Reference solution 0.1 Reference solution


Reduced porous Reduced porous
0 0
500 1000 1500 2000 2500 3000 500 1000 1500 2000 2500 3000
Frequency (Hz.) Frequency (Hz.)

(c) (d)
FIG. 11. (Color online) Absorption coefficient; Convergence of the reduced model to the reference solution with added modes in the basis and a low frequency
correction vector: (a) one mode, (b) two modes, (c) four modes, (d) six modes.

where the conservative problem consists of the acoustic geom- tions, leading to good sparsity properties of the reduced mat-
etry presented in Fig. 12 when the porous layer is removed rices. As illustrated in Figs. 15, the 3070 poroelastic DOF
(the largest dimension of the cavity is then 0.7 m). As observed are downsized to 800 modal unknowns, of which 414 corre-
ð2Þ
for the 1D application, the damping effect of the porous layer spond to KP - and CP -orthogonal modes, as introduced in
can be clearly seen, combining a frequency-shift of the Sec. III B. A detailed view of the sparsity associated with the
resonances toward low frequencies (particularly obvious for “non-orthogonal” modal coordinates is given in Fig. 16(b).
the first three acoustic peaks), and a strong attenuation of the The sparsity of the reduced system matrix is mostly affected
acoustic pressure level (some strongly damped resonances do by the use of attachment functions which fully couple inter-
not even appear at all in the dissipative solution). face DOF to modal unknowns. Although out of the scope of
this contribution, reducing the number of attachment func-
tions in the reduction procedure is therefore an important as-
2. Reduced problem
pect to enhance the performance of the method, especially
Similarly to the 1D application, the convergence is first when applications with high interface-to-inner-DOF ratios
checked, increasing the number of modes included in the ba- are considered.30,31 Associating attachment functions to the
sis, for a response in the range ½0–1000 Hz. Although a rela- acoustic DOF rather than to the poroelastic interface DOF,
tively large basis is needed in order to capture the dynamic as established in this contribution, constitutes one step to-
behavior of the porous medium in the considered frequency ward a reduced interface dimension. From a storage perspec-
ð1Þ
range, the solution eventually converges toward the refer- tive, using KP -static responses as attachment functions
ence solution (Fig. 14). Thus, a modal basis including 500 leads to uncoupled interface and modal unknowns in the
ð1Þ
modes—truncation around 775 Hz—provides an approxi- reduction of KP [see Eq. (29)].
mate estimation of the dynamic behavior up to around Further condensation of the modal coordinates corre-
ð2Þ
600 Hz, while with 800 modes—truncation around sponding to the KP - and CP -orthogonal modes, as presented
1250 Hz—a good agreement between the solutions can be in Sec. III C, leads to a contribution of 386 poroelastic DOF
observed over the entire frequency range of interest. in the final coupled problem to be solved. Considering the
Besides the convergence, the interest of the proposed relatively high proportion of orthogonal modes, substantial
method specifically lies in the fact that real-valued modes improvements can be expected applying this frequency-
are computed directly from the coupled poroelastic equa- dependent additional step. The sparsity of the further

3174 J. Acoust. Soc. Am., Vol. 132, No. 5, November 2012 Rumpler et al.: Modal-based porous reduction

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms
190

Mean quadratic pressure in cavity (dB)


180

170

160

150

140

130

120
FIG. 12. (Color online) Acoustic cavity mesh and dimensions for 3D problem.
110
Rigid cavity without porous layer
100 Unreduced problem − porous damped cavity
100 200 300 400 500 600 700 800 900 1000

condensed global system matrix, as well as a detailed view Frequency (Hz.)


of the sparsity associated with the poroelastic part, are pre-
sented in Figs. 16. Presented as an important condition for FIG. 13. (Color online) Mean quadratic pressure reference FRF, 3D problem.
the scalability of the modal approach in Sec. III B, the sub-
matrix sparsity associated with the “non-orthogonal” modal
coordinates [Fig. 16(b)] can be quantified by its matrix den- even offers a matrix density improvement by a factor 2 for
sity, of 0.0187. As a comparison, the original coupled poro- the studied configuration. This is of course hindered by the
acoustic problem exhibits a matrix density of 0.0165 [Fig. attachment functions coupling: the system matrix for
15(a)], while its submatrix associated with the non-reduced the coupled problem with the reduced porous domain and
porous equations has a matrix density of 0.0379. Thus, the condensation of the orthogonal coordinates [Fig. 16(a)] has a
reduction, only accounting for the non-orthogonal modes, matrix density of 0.029.

180 180

170 170
Mean quadratic pressure in cavity (dB)

Mean quadratic pressure in cavity (dB)

160 160

150 150

140 140

130 130

120 120

110 110
Unreduced problem − porous damped cavity Unreduced problem − porous damped cavity
100 Reduced problem − porous damped cavity 100 Reduced problem − porous damped cavity
100 200 300 400 500 600 700 800 900 1000
100 200 300 400 500 600 700 800 900 1000
Frequency (Hz.) Frequency (Hz.)

(a) (b)
180

170
Mean quadratic pressure in cavity (dB)

160

150

140

130

120

110
Unreduced problem − porous damped cavity
100 Reduced problem − porous damped cavity
100 200 300 400 500 600 700 800 900 1000
Frequency (Hz.)

(c)
FIG. 14. (Color online) Mean quadratic pressure FRF. Convergence of the reduced model to the reference solution: (a) 100 modes (127 Hz), (b) 500 modes
(777 Hz), (c) 800 modes (1245 Hz).

J. Acoust. Soc. Am., Vol. 132, No. 5, November 2012 Rumpler et al.: Modal-based porous reduction 3175

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms
FIG. 15. (Color online) Sparsity of system matrix for (a) unreduced and (b) reduced poroelastic domain.

Regarding an estimation of the computational efficiency, the mixed formulation. The frequency-dependent condensa-
Table III completes Figs. 17, presenting a detailed comparison tion of part of the modal coordinates to the interface DOF fur-
of the estimated computational times, between the reference ther improves the efficiency, with a factor 3.3 for the CPU
direct solution using the displacement formulation for the po- time, compared to the reference solution, and almost 50%
rous medium, and the proposed reduced set of poroelastic faster than the solution using the mixed formulation. Such an
equations. A comparison to an efficient implementation of the approach, based on an initial reduction of the set of equations
porous formulation using mixed variables15 is also added to to be solved, proves even more attractive in situations where
the CPU time estimations. With this computationally efficient multiple load cases are required, which is practically the case
formulation, the poroelastic physical DOF are downsized in most industrial applications. In fact, the costly initial step,
from 3070 to 2237 DOF. The associated frequency sweep is which consists in calculating the modal basis and applying the
then 2.25 times faster than the reference solution using the transformation, needs to be done only once for the first load
ðus ; uf Þ formulation. Comparing the proposed approach to case. Such multiple calculations, for the test example of inter-
the reference solutions, a reduction including 800 modes in est, would then exhibit a CPU time enhancement factor of 3.7
the basis leads to an overall enhancement factor of 2.8 for the and 1.6 for the reduced set of equations when compared to the
CPU time, compared to the reference solution, while repre- reference frequency sweep and the solution with the mixed
senting a 25% enhancement of the solution efficiency using formulation respectively. Further condensing the unknowns

(a) (b) 0

50

100

150

200

250

300

350

0 50 100 150 200 250 300 350


nz = 2782

FIG. 16. (Color online) (a) Sparsity of system matrix for reduced poroelastic domain with condensed “orthogonal” modal unknowns and (b) focus on the spar-
sity for the “non-orthogonal” modal unknowns.

3176 J. Acoust. Soc. Am., Vol. 132, No. 5, November 2012 Rumpler et al.: Modal-based porous reduction

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms
TABLE III. Details of computational time estimations. However, several aspects have to be settled in order to make
the presented method fully applicable in an industrial con-
Reference Direct Reduced Reduced
text: a priori truncation criteria, selection of most relevant
(us , uf ) (us , pf ) condensed modes, integration in complex setups (with other poroelastic
domains, structural domain). The following general com-
Modes and projection (s) — — 85 85
Frequency sweep (s) 970 432 261 209
ments can be made with respect to these points.
Total CPU time (s) 970 432 346 294 (1) A priori criteria are needed in order to estimate the trun-
CPU time/increment (s) 1.94 0.864 0.522 0.418 cation of the modal basis according to the frequency
range of interest. On this matter, the rule of thumb con-
sisting in keeping the modes whose eigenfrequencies are
associated with the “orthogonal” poroelastic modes enables a smaller than 1.5 to 2.5 times the highest frequency of in-
frequency sweep 4.6 times as fast as for the reference solu- terest26,32 seems sensible in the considered cases, pro-
tion, and twice as fast as the frequency sweep using the mixed vided a sufficiently large frequency range is considered.
formulation. This needs to be confirmed on a larger range of applica-
However, the provided performance estimations in tions and porous materials.
terms of computational time have to be taken as qualitative (2) In the extension to the 3D application, the convergence
rather than quantitative considering the following aspects. to the reference solution is rather slow, and some
modes included in the basis seem not to bring a signifi-
(1) The unreduced reference solution of the problem is com- cant contribution. Therefore, finding a cost-efficient
puted as given in Eqs. (20) or (23), which involves that way to select the most significant modes to be included
four frequency-independent matrices are assembled once in the truncated basis constitutes a further enhancement
initially, and combined at each frequency increment. It is of the proposed reduction method, currently under
likely that an industrial implementation of the unreduced development.
problem would involve only two frequency-dependent (3) Once the purely modal reduction issues are settled, as ini-
matrices (equivalent mass and stiffness), thus involving tiated in the scope of this contribution, the integration of
assembly procedure at each frequency increment. the approach into more complex situations (geometry,
(2) The efficiency measured is dependent on the program- multiple domains) is an essential point to investigate. The
ming language as well as the quality of the implementa- difficulties are then relocated from volume considerations
tion, especially when frequency-dependent operations (e.g., costly 3D modeling of porous layers) to interface
such as the dynamic condensation have to be performed problems. As previously mentioned in this section, the
at each frequency increment. high number of interface DOF for poroelastic materials
(3) The acoustic DOF being kept unreduced, even though presents the risk of jeopardizing the efficiency gained
reduction is straightforward to establish, the given using a modal approach. Several contributions have been
performance in terms of CPU time does not fully made recently on this topic, in the scope of structural-
render the efficiency associated with the poroelastic acoustics and dynamics (reduction of interface func-
reduction. tions,33,34 modal coupling using decoupled bases and cor-
Nevertheless, besides the above-mentioned points, the rection Ritz vectors,29,35 use of Lagrange multipliers,36,37
proposed modal-based reduction of the poroelastic finite ele- etc.). A follow-up to the present approach is therefore to
ment discretization offers promising possibilities for the test such methodologies in order to further extend the
efficient computation of frequency response functions. range of efficient application of the method.

(a) 1000 (b) 1


Reference direct solution
900 Direct solution − Mixed formulation 0.9
Reference−normalized CPU time (s)

Reduced problem
800 0.8
Reduced and condensed problem
700 0.7
CPU time (s)

600 0.6

500 0.5

400 0.4

300 0.3

200 0.2
Direct solution − Mixed formulation
100 0.1 Reduced problem
Reduced and condensed problem
0 0
0 50 100 150 200 250 300 350 400 450 500 0 50 100 150 200 250 300 350 400 450 500
Number of frequency increments Number of frequency increments

FIG. 17. (Color online) (a) CPU times for FRF computation; (b) reference-normalized CPU times.

J. Acoust. Soc. Am., Vol. 132, No. 5, November 2012 Rumpler et al.: Modal-based porous reduction 3177

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms
8
V. CONCLUSION W. Desmet, “A wave based prediction technique for coupled vibro-
acoustic analysis,” Ph.D. thesis, K.U. Leuven, division PMA, Leuven,
In the present work, an original way of solving the poroe- 1998, 448 pp.
9
lastic equations was presented, as well as the corresponding W. Desmet, B. Van Hal, P. Sas, and D. Vandepitte, “A computationally ef-
ficient prediction technique for the steady-state dynamic analysis of
variational formulation for poro-acoustic applications. This coupled vibro-acoustic systems,” Adv. Eng. Software 33, 527–540 (2002).
approach, enabled by a rearrangement of the poroelastic equa- 10
E. Deckers, N. H€ orlin, D. Vandepitte, and W. Desmet, “A novel wave
tions, was used to define a standard eigenvalue problem for based prediction technique for the efficient dynamic modeling of poro-
the coupled poroelastic system, based on the solid phase and elastic materials,” in Proceedings of Euronoise, Edinburgh, Scotland,
2009, paper 0197.
fluid phase displacements formulation. The generated modes 11
E. Deckers, B. Van Genechten, D. Vandepitte, and W. Desmet, “Efficient
were showed to form a well suited basis for a priori reduction treatment of stress singularities in poroelastic wave based models using
of the finite element problem. The tests on a 1D poro-acoustic special purpose enrichment functions,” Comput. Struct. 89, 1117–1130
application, in a component mode synthesis scheme where (2011).
12
Y. J. Kang and J. S. Bolton, “Finite element modeling of isotropic elastic
only the poroelastic domain was reduced, attested the suitabil- porous materials coupled with acoustical finite elements,” J. Acoust. Soc.
ity of the proposed modal basis. The extension to a 3D appli- Am. 98, 635–643 (1995).
13
cation demonstrated the potential performance in terms of R. Panneton and N. Atalla, “An efficient finite element scheme for solving
the three-dimensional poroelasticity problem in acoustics,” J. Acoust. Soc.
problem size reduction, associated CPU time enhancement, as
Am. 101, 3287–3298 (1997).
well as sparsity of the reduced matrices, which is an important 14
N. Dauchez, S. Sahraoui, and N. Atalla, “Convergence of poroelastic finite
asset for use in very large applications. Out of the 3942 DOF elements based on Biot displacement formulation,” J. Acoust. Soc. Am.
involved, the 3070 belonging to the poroelastic domain were 109, 33–40 (2001).
15
N. Atalla, M. A. Hamdi, and R. Panneton, “Enhanced weak integral for-
reduced to 800 modal coordinates. Among those, 414 corre- mulation for the mixed (u,p) poroelastic equations,” J. Acoust. Soc. Am.
sponding to linearly independent equations, were dynamically 109, 3065–3068 (2001).
16
condensed leading to a poroelastic domain downsized from O. Dazel, F. Sgard, F. X. Becot, and N. Atalla, “Expressions of dissipated
3070 to 386 DOF. An estimation of the CPU time enhance- powers and stored energies in poroelastic media modeled by {u, U} and
{u, P} formulations,” J. Acoust. Soc. Am. 123, 2054–2063 (2008).
ment leads to a factor 3:1 to 4:4, depending on whether the 17
O. Dazel, B. Brouard, C. Depollier, and S. Griffiths, “An alternative Biot’s
time elapsed to build the reduction basis is taken into account displacement formulation for porous materials,” J. Acoust. Soc. Am. 121,
or not. A priori criterion for modal basis truncation, as well as 3509–3516 (2007).
18
filtering of relevant modal components to be included in the N. E. H€ orlin, M. Nordstr€om, and P. G€ oransson, “A 3-D hierarchical FE
formulation of Biot’s equations for elasto-acoustic modeling of porous
basis, are two complements to this work in order to make it media,” J. Sound Vib. 245, 633–652 (2001).
fully applicable for larger and more industrial-like applica- 19
S. Rigobert, N. Atalla, and F. C. Sgard, “Investigation of the convergence
tions. Furthermore, focusing on enhanced attachment func- of the mixed displacement-pressure formulation for three-dimensional
tions, especially in the case where multilayer poroelastic poroelastic materials using hierarchical elements,” J. Acoust. Soc. Am.
114, 2607–2617 (2003).
materials are considered, presents an extension to this contri- 20
O. Dazel, F. Sgard, and C. H. Lamarque, “Application of generalized com-
bution, in order not to lose the benefits of subdomain reduc- plex modes to the calculation of the forced response of three-dimensional
tion because of large interfaces. poroelastic materials,” J. Sound Vib. 268, 555–580 (2003).
21
P. Davidsson and G. Sandberg, “A reduction method for structure-acoustic
and poroelastic-acoustic problems using interface-dependent lanczos
ACKNOWLEDGMENTS vectors,” Comput. Methods Appl. Mech. Eng. 195, 1933–1945 (2006).
22
C. Batifol, M. N. Ichchou, and M. A. Galland, “Hybrid modal reduction
The authors gratefully acknowledge the Marie Curie for poroelastic materials,” C. R. Mec. 336, 757–765 (2008).
RTN project: A Computer Aided Engineering Approach for 23
O. Dazel, B. Brouard, N. Dauchez, and A. Geslain, “Enhanced Biot’s fi-
Smart Structures Design (MC-RTN-2006-035559), and the nite element displacement formulation for porous materials and original
resolution methods based on normal modes,” Acta Acust. Acust. 95, 527–
ITN Marie Curie project GA-214909 “MID-FREQUENCY– 538 (2009).
CAE Methodologies for Mid-Frequency Analysis in Vibra- 24
O. Dazel, B. Brouard, N. Dauchez, A. Geslain, and C. H. Lamarque, “A
tion and Acoustics.” free interface CMS technique to the resolution of coupled problem involv-
ing porous materials, application to a monodimensional problem,” Acta
Acust. Acust. 96, 247–257 (2010).
1 25
M. A. Biot, “Theory of propagation of elastic waves in a fluid-saturated R. R. Craig and C. Chang, “A review of substructure coupling methods for
porous solid. I. Low-frequency range,” J. Acoust. Soc. Am. 28, 168–178 dynamic analysis,” Adv. Eng. Sci. NASA 2, 393–408 (1976).
26
(1956). R. Rumpler, A. Legay, and J. De€ u, “Performance of a restrained-interface
2
M. A. Biot, “Theory of propagation of elastic waves in a fluid-saturated substructuring FE model for reduction of structural-acoustic problems
porous solid. II. Higher frequency range,” J. Acoust. Soc. Am. 28, 179– with poroelastic damping,” Comput. Struct. 89, 2233–2248 (2011).
27
191 (1956). H. J. Morand and R. Ohayon, Fluid Structure Interaction (Wiley, Chiches-
3
M. A. Biot, “Generalized theory of acoustic propagation in porous dissipa- ter, 1995), 220 pp.
28
tive media,” J. Acoust. Soc. Am. 34, 1254–1264 (1962). R. Ohayon, “Reduced models for fluid-structure interaction problems,”
4
J. Allard and N. Atalla, Propagation of Sound in Porous Media: Modelling Int. J. Numer. Methods Eng. 60, 139–152 (2004).
29
Sound Absorbing Materials (Wiley, New York, 2009), 372 pp. Q. H. Tran, M. Ouisse, and N. Bouhaddi, “A robust component mode syn-
5
A. Bermudez and R. Rodrıguez, “Modelling and numerical solution of thesis method for stochastic damped vibroacoustics,” Mech. Syst. Signal
elastoacoustic vibrations with interface damping,” Int. J. Numer. Methods Process. 24, 164–181 (2010).
30
Eng. 46, 1763–1779 (1999). D. M. Tran, “Component mode synthesis methods using partial interface
6
A. Bermudez, L. Hervella-Nieto, and R. Rodrıguez, “Finite element com- modes: Application to tuned and mistuned structures with cyclic
putation of the vibrations of a plate-fluid system with interface damping,” symmetry,” Comput. Struct. 87, 1141–1153 (2009).
31
Comput. Methods Appl. Mech. Eng. 190, 3021–3038 (2001). M. Junge, D. Brunner, J. Becker, and L. Gaul, “Interface-reduction for the
7
J. De€u, W. Larbi, and R. Ohayon, “Vibration and transient response of Craig-Bampton and Rubin method applied to FE-BE coupling with a large
structural-acoustic interior coupled systems with dissipative interface,” fluid-structure interface,” Int. J. Numer. Methods Eng. 77, 1731–1752
Comput. Methods Appl. Mech. Eng. 197, 4894–4905 (2008). (2009).

3178 J. Acoust. Soc. Am., Vol. 132, No. 5, November 2012 Rumpler et al.: Modal-based porous reduction

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms
32 35
M. A. Tournour, N. Atalla, O. Chiello, and F. Sgard, “Validation, perform- E. Balmès, “Optimal ritz vectors for component mode synthesis using the
ance, convergence and application of free interface component mode syn- singular value decomposition,” AIAA J. 34, 1256–1260 (1996).
36
thesis,” Comput. Struct. 79, 1861–1876 (2001). D. Rixen, C. Farhat, and M. Geradin, “A two-step, two-field hybrid
33
E. Balmès, “Use of generalized interface degrees of freedom in component method for the static and dynamic analysis of substructure problems with
mode synthesis,” in Proceedings of IMAC, 1996, pp. 204–210. conforming and non-conforming interfaces,” Comput. Methods Appl.
34
J. Herrmann, M. Maess, and L. Gaul, “Substructuring including interface Mech. Eng. 154, 229–264 (1998).
37
reduction for the efficient vibro-acoustic simulation of fluid-filled piping D. Rixen, “A dual Craig-Bampton method for dynamic substructuring,” J.
systems,” Mech. Syst. Signal Process. 24, 153–163 (2010). Comput. Appl. Math. 168, 383–391 (2004).

J. Acoust. Soc. Am., Vol. 132, No. 5, November 2012 Rumpler et al.: Modal-based porous reduction 3179

Downloaded 09 Nov 2012 to 134.212.178.10. Redistribution subject to ASA license or copyright; see http://asadl.org/terms

You might also like