Vous êtes sur la page 1sur 30

Towards Lattice-Boltzmann Prediction of Turbofan

Engine Noise
D. Casalino

A. F. P. Ribeiro

E. Fares

S. Nolting

Exa GmbH, Curiestrasse 4, Stuttgart, 70563, Germany


A. Mann

F. Perot

Exa Corporation, 150 North Hill Drive, Brisbane, CA, 94005, USA
Y. Li

P.-T. Lew

C. Sun

P. Gopalakrishnan

R. Zhang

H. Chen

Exa Corporation, 55 Network Dr, Burlington, MA,01803, USA


K. Habibi

Department of Mechanical Engineering, McGill University, Montreal, QC H3A 0C3, Canada


The goal of the present paper is to report verication and validation studies carried
out by Exa Corporation in the framework of turbofan engine noise prediction through the
hybrid Lattice-Boltzmann/Ffowcs-Williams & Hawkings approach (LB)-(FW-H). The un-
derlying noise generation and propagation mechanisms related to the jet ow eld and the
fan are addressed separately by considering a series of elementary numerical experiments.
As far as fan and jet noise generation is concerned, validation studies are performed by
comparing the LB solutions with literature experimental data, whereas, for the fan noise
transmission through and radiation from the engine intake and bypass ducts, LB solutions
are compared with nite element solutions of convected wave equations. In particular, for
the fan noise propagation, specic verication analyses are carried out by considering tonal
spinning duct modes in the presence of a liner, which is modelled as an equivalent acoustic
porous medium. Finally, a capability overview is presented for a comprehensive turbofan
engine noise prediction, by performing LB simulation for a generic but realistic turbofan
engine conguration.

Technical Director, Aeroacoustics, Aerospace, AIAA member.

Team leader, Aerospace.

Technical Director, Aerospace, Senior AIAA member.

Vice President, Aerospace, AIAA member.

Senior Aeroacoustics Engineer, Aeroacoustics, AIAA member.

Senior Technical Director, Aeroacoustics, AIAA member.

Consulting Scientist, Physics, AIAA member.

Senior Physics Validation Engineer, Physics, AIAA member.

Consulting Scientist, Physics.

Principle Scientist, Physics.

Senior Director, Physics.

Chief Scientic Ocer, Physics.

Research Assistant, AIAA student member.


1 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

20th AIAA/CEAS Aeroacoustics Conference
16-20 June 2014, Atlanta, GA
AIAA 2014-3101

AIAA Aviation
Nomenclature
C Particle collision operator
D
j
Nozzle diameter
J Impedance tting cost function
K Modelled turbulent kinetic energy
M Mach number
R Acoustic porous medium viscous resistance
R
f
Fan plane radius
Re Reynolds number
T Fluid temperature
T
s
Solution sampling simulation time
T
t
Transient (settling) simulation time
X
w
Witze jet potential core length ratio
Z Acoustic porous medium impedance
c Speed of sound
d Acoustic porous medium thickness
f Particle probability distribution function or
frequency
k Acoustic wavenumber (/c

)
m, n Azimuthal and radial duct mode order
p Pressure
u Flow velocity
u

, v

, w

Fluctuating ow velocity components


u, v, w Flow velocity components
v Particle velocity
x Space coordinates
Subscripts
0 Tonal frequency
Free-stream/ambient conditions
i Discrete particle velocity index
j Jet centerline exit conditions
Symbols
t Time step
x Voxel size
Turbulent dissipation rate
Acoustic wavelength
Acoustic angular frequency
Acoustic porous medium porosity
Density
Relaxation time
Fourier transform
Complex conjugate
I. Introduction
T
he costs and the risks of development of new air vehicles can be signicantly reduced by reducing the
performance uncertainties associated with each aircraft system/component during the dierent design
stages, from the conceptual design phase, through the preliminary design phase, until the detailed design
phase. Following a system engineer approach, the performance uncertainties associated with one single
system have an impact on the requirements specied for all the other systems. Therefore, the sooner
the uncertainties are reduced under admissible tolerances, the faster the design process converges towards
the target conguration. Moreover, keeping uncertainties within narrow ranges reduces the risk of costly
adjusting procedures to achieve some requirements at aircraft level during the detailed design phase or, even
worse, during the aircraft certication process. This scenario is valid for any performance requirement, thus
including aircraft community noise. For instance, an accurate knowledge during the conceptual design phase
of the lift penalty due to the presence of a high-lift low-noise device would allow to accurately estimate
the required increase of approach velocity with respect to a nominal value for an untreated wing, and thus
the expected increase of landing gear noise. Then, once the expected airframe noise has been estimated
for a newly determined nominal approach velocity, the required engine thrust and the corresponding noise
contribution can be evaluated and added to the airframe noise components to evaluate the overall aircraft
noise to be checked against the target certication levels.
Computational Fluid Dynamics (CFD) and Computational AeroAcoustics (CAA) numerical experiments
can play a crucial role in reducing the uncertainties associated with the aerodynamic and aeroacoustic per-
formance of each system during the various design phases, and in particular during the conceptual design
phase, when a very broad design space needs to be explored before dening reliable and optimal require-
ments for each system. The main advantage of CFD/CAA experiments with respect to Wind-Tunnel (WT)
experiments is that full aircraft details and ight conditions (Mach and Reynolds numbers) can be taken
into account relatively easier, thus helping to keep the scattering of the uncertainties under admissible limits.
Such a capability is nowadays exploited for the evaluation of aircraft aerodynamic performance in dierent
ight conditions, including take-o and landing under ground eects, but not yet exploited for the evaluation
of the aeroacoustic performances. The reason of this gap is twofold: on one side, conventional CAA tools
have not reached yet the adequate maturity level in terms of reliability compared to WT experiments, mainly
due to the physical complexities associated with the sound generation mechanisms in turbulent ows; on the
2 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

other side, the computational resources required by space-/time-resolved turbulent ow simulations are still
prohibitively large for most cases of practical relevance.
Recently, a new CFD/CAA technique based on a Lattice-Boltzmann ow Model (LBM) has been em-
ployed to tackle airframe noise problems, both at a system level
1, 2, 3, 4, 5
and at full aircraft level.
6, 7
Such
a technique has achieved the required readiness level to become a viable alternative to WT experiments,
both in terms of costs for a target accuracy level of about 1 dB, and turnaround time for down-selecting
low-noise designs and concepts. With the goal of extending the application of PowerFLOW to aero-engine
aeroacoustics, towards a comprehensive CFD/CAA prediction of aircraft community noise, including engine
installation eects, the present eort addresses the prediction of the main turbofan engine noise components.
Increasing the readiness level of PowerFLOW in this new application eld through benchmark studies and
the development of simulation best practices is a necessary condition for achieving the ultimate goal of a
virtual aircraft noise certication process to be executed at dierent phases of the aircraft design.
In the present study, the LBM is used to compute the generation and propagation of acoustic uctuations
from the fan/Outlet Guide Vane (OGV) system and the jet of a turbofan engine. The turbulent ow
uctuations are resolved up to a certain scale using a Lattice-Boltzmann Very Large Eddy Simulation
(LBM-VLES) approach. One original aspect is the use a new non-isothermal version of the LBM, which
allows extending the Mach number range of the standard LBM scheme up to about 0.95. Such a capability is
validated by computing the ow eld and the associated noise of a single hot subsonic jet. Another original
aspect of the present study is the use of an Acoustic Porous Medium (APM) included in the ow simulation
domain to mimic the eects due of a honeycomb liner installed on the walls of the engine intake. The
physical properties of the APM (thickness, resistivity and porosity) are tuned against a target impedance-
frequency law using a genetic algorithm minimization technique. The liner simulation capability is veried
by computing the propagation of spinning duct modes through an aero-engine intake and comparing the far-
eld noise directivity to the one computed using a Finite Element Method (FEM), which solves a convected
wave equation in the frequency domain.
The paper is organized in the following way. The numerical method is presented in section II, including
the newly developed high-speed LB formulation and the acoustic porous medium. Section III presents
elementary verication and validation studies for the dierent source components of an aero-engine. A
demonstration of a comprehensive turbofan noise prediction is provided in section IV. The main outcomes
of the present eort are nally summarized in the conclusion section V. The derivation of the impedance
boundary condition used for the FEM computations of section III is reported in the Appendix.
II. Numerical approach
The LBM-based CFD/CAA solver PowerFLOW developed and distributed by Exa Corporation is used
to compute unsteady ow physics and the resulting generation and propagation of acoustics waves. A FW-H
approach is then used to extrapolate the near eld solution sampled on a permeable surface to the far-eld.
A standard D3Q19 lattice scheme (3 dimensions, 19 velocity states per mesh node) is employed, but the
Mach number range is extended beyond the typical value of 0.4 by relaxing the iso-thermal assumption
of the standard LB model. The eects due to the presence of a liner are modelled using an APM
8
directly
implemented in the LB scheme. Details about the standard LB formulation, its high-speed extension and
the APM are reported in the following subsections.
II.A. Standard LB model
The standard LBM implemented in PowerFLOW has the following form:
f
i
(x +v
i
t, t + t) f
i
(x, t) = C
i
(x, t) (1)
where f is the particle density function, which represents the probability for particles to travel with speed v
from the position x at time t in the discrete direction i. The collision term C is modeled with the well-known
Bhatnagar-Gross-Krook (BGK) approximation
9, 10
as follows:
C
i
(x, t) =
1

[f
i
(x, t) f
eq
i
(x, t)] , (2)
where is the relaxation time, which is related to the uid viscosity, and f
eq
i
is the equilibrium distribu-
tion, which is approximated by a third order expansion with constant temperature.
11
Once the distribution
3 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

function is computed, the density and linear momentum are simply determined through discrete integra-
tion: (x, t) =

i
f
i
(x, t) and u(x, t) =

i
f
i
(x, t) v
i
. All the other quantities are determined through
thermodynamic relationships for an ideal gas.
With the Chapman-Enskog expansion
12
expansion it is possible to retrieve the compressible Navier-Stokes
equations from LBM.
10, 13
However, in contrast to the Navier-Stokes equations, the current method is based
on linear formulation which relies on simple computational operations, allowing for ecient, accurate, and
highly scalable implementations. An explicit time advancement scheme is employed, which also facilitates
massively parallel simulations and the use of a very small time step, which is valuable when attempting to
simulate high frequency noise.
PowerFLOW solves the D3Q19 formulation of the Lattice-Boltzmann equation for direct numerical sim-
ulations. For high Reynolds ows, turbulence modeling is introduced
14
by solving a variant of the RNG
k model
15, 16
on the unresolved scales,
17
selected via a swirl model,
18
a method referred to as LBM Very
Large Eddy Simulation (LBM-VLES). An extended wall model including pressure gradient eects is used in
the near-wall region.
19
The LBM scheme is solved on a grid composed of cubic volumetric elements (voxels). A Variable Reso-
lution (VR) by a factor of two is allowed between adjacent regions. Consistently, the time step is varied by
a factor of two between two adjacent resolution regions. Solid surfaces are automatically facetized within
each voxel intersecting the wall geometry using planar surface elements
20, 21
(surfels).
Although PowerFLOW has intrinsic CAA capabilities and can compute the noise propagation directly
from the unsteady ow simulations, this is generally computationally expensive and limited to the near eld.
Therefore, in order to compute the far eld noise, an integral extrapolation based on the FW-H acoustic
analogy
22
is used, both using solid and permeable
23
surface formulations. A forward-time solution
24
of the
FW-H equation based on Farassats formulation 1A
25
is employed. The solver is embedded in the post-
processing tool PowerACOUSTICS, which is also used to perform statistical and spectral analysis of any
unsteady solutions generated by PowerFLOW (volume elds, surface elds, and probe signals).
The numerical methods herein described were extensively validated for a wide variety of applications
ranging from academic cases using direct numerical simulation
26
to industrial ow problems in the elds of
aerodynamics,
27
thermal management,
28
and aeroacoustics.
2, 1, 4, 6, 7
II.B. High-speed non-isothermal LB model
For the solution in the high subsonic Mach number range, e.g, ows with local Mach number greater than
0.5, a standard D3Q19 LBM is applied. The BGK collision operator in Eq.2 was replaced by a regularized
collision operator which can signicantly increase both numerical stability and accuracy when local ow
Mach number is high. An interaction force was introduced in Eq.1,
29
which can modify the equation of
state thus the speed of sound so that high Mach number ows can be simulated by a low order LB scheme.
Moreover, in order to take into account the ow heating due to compression work and viscous dissipation,
a hybrid approach was applied for the thermodynamics of energy eld, by solving the entropy equation
through a Lax-Wendro nite dierence scheme on the Cartesian LB mesh.
II.C. Acoustic porous medium model
In the LB method, external forces can be included in the uid dynamics by altering the local-instantaneous
particle distributions during the collision step. This technique can be used to model, for example, buoyancy
eects due to gravity. The solver used in the present study implements a porous medium model by applying
an external force driven by the ow resistivity and as function of the local ow velocity.
30
This model can
be used to predict pressure losses that aect the time-averaged ow eld solution and, at the same time, the
instantaneous acoustic uctuations.
The APM model in PowerFLOW, a recently patented technology by Exa Corporation,
31
is characterized
by three main parameters: the viscous resistance R, the porosity and the APM thickness d. The char-
acteristic surface impedance of the APM, say Z
APM
(f, R, , d), can be fully determined with these three
variables via an analytical model, where f is the frequency in Hertz; Z
APM
is also parameterized through
quantities that are directly related to the LB formulation, but these details are beyond the present scope.
The characteristic surface impedance Z(f) of an acoustic liner can be obtained via experiments, or via nu-
merical simulations of the complete geometry, in a normal or grazing impedance tube setup.
32
The three
4 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

parameters are then educed by minimization of the cost function J(R, , d):
J(R, , d) = max
f
(|Z(f) Z
APM
(f, R, , d)|) (3)
The cost function is minimized using a genetic algorithm for (R, , d) constrained on specic variation
ranges. The optimal results (R
0
,
0
, d
0
) are then validated with the simulation of the resulting APM in a
normal impedance tube and the computation of its numerical characteristic impedance. The target acoustic
resistance and reactance curves, their analytical tting of the LBM educed impedance curves are then plotted
for visual verication, as shown in subsection III.A.
III. Verication and validation
This section deals with the unitary verication and validation studies carried out to build up the
required knowledge in terms of ow physics and simulation setup towards a comprehensive turbofan engine
aeroacoustic analysis. Fan and jet noise sources are addressed, whereas the internal turbomachinery ow
features (compressor, combustion chamber and turbine) are not in the present scope.
Subsection III.A deals with the propagation of a spinning duct mode through an acoustically treated
intake. LBM results are compared with a FEM solution of a convected 2nd order wave equation in the
frequency domain.
33, 34
The main goal of this study is to verify the LBM propagation capabilities in the
presence of an impedance wall modelled via an APM.
Subsection III.B deals with the propagation of a spinning duct mode through a bypass duct. LBM results
are compared with a FEM solution of a convected 3rd order wave equation in the frequency domain.
35
The
main goal of this study is to perform a cross validation of sound propagation in the presence of a shear layer.
Subsection III.C deals with the problem of tonal and broadband noise generation from rotating fan
blades. The Advanced Noise Control Fan (ANCF) conguration
36
is used as benchmark validation case.
The complete 3-D in-duct rotor/stator model and the rotating rotor is simulated for the nominal value of
Mach rotor tip of 0.33 used in the experiments. Far-eld measurements conducted at the NASA Glenn
research center are used to validate the simulation results.
Finally, subsection III.D focuses on the prediction of single stream jet ow noise. The NASA Glenn
research center hot jet experiments
37
carried out with the SMC000 nozzle geometry are used to validate the
LBM solution for high speed cold and hot jets.
Throughout, the size of the simulation problems is reported in terms of equivalent number of voxels and
surfels as the solution in the whole simulation domain would be updated every time step. In other words,
one voxel or surfel belonging to the second nest mesh resolution level, for instance, would count as a half
in the overall count of Fine Equivalent Voxels (FEV) and Fine Equivalent Surfels (FES). Simulation time is
reported in terms of overall computational hours (CPUh) on a cluster of Intel Xeon X5570 2.93GHz CPUs
connected by a Mellanox FDR Inniband 56Gb/s network.
All simulations presented in this section were performed using the recently released PowerFLOW version
5.0c, with the exception of the jet ow simulations presented in subsections III.B and III.D, which were
performed using a beta release of PowerFLOW 5.1, providing the high-speed non-isothermal functionality.
III.A. Intake fan noise radiation
In this subsection we show results for the transmission of a spinning mode through an idealized aero-engine
intake. The Helmholtz number, based on the fan plane outer radius, is kR
f
=5.257. The acoustic mode (6, 1)
is considered, with a magnitude corresponding to an overall inlet acoustic power of 130 dB. The acoustic
mode is introduced in the LBM simulation through a time-varying pressure/velocity boundary condition on
the fan plane. LBM results without and with mean ow convective eects and without and with an acoustic
liner are compared against a frequency-domain FEM solution of a convected wave equation for the acoustic
potential
34, 38, 39
put forward by Pierce.
40
Simulations are carried out for a three-dimensional rigid and acoustically treated (soft) nacelle. A zero-
splice liner is considered, located on the outer wall of the intake, as shown in Fig. 1. Both for the LBM and
the FEM computations, the near-eld solution is extrapolated to the far eld through a permeable FW-H
approach. In order to reduce the impact of the FW-H extrapolation to the far-eld noise prediction, the
same integration surface is used for the two simulations, as depicted in Fig. 1.
5 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

Figure 1. Nacelle intake simulation setup.
The LBM simulations are performed using a mesh resolution that guarantees 14 points per acoustic
wavelength (without convective eects) in the near eld volume inside the nacelle up to the FW-H integration
surface. Views of the mesh for the rigid and soft cases are shown in Fig. 2. The volume occupied by the
APM is also meshed, as this is treated as a uid material with dierent properties. Quantities that are useful
to estimate the computational eort for the present case are reported in Table 1.
Figure 2. Nacelle intake simulation mesh. Rigid case (left), soft case (right).
VRs x FEV (10
6
) FES (10
6
) T
t
T
s
CPUh
9
0
/56 4.999 2.222 377/f
0
10/f
0
430
Table 1. Nacelle intake simulation properties. VRs denotes the number of variable mesh resolution levels.
As already mentioned, the properties of the APM are dened by using an analytical expression that relates
the surface liner impedance to resistivity, porosity and thickness of the APM. Then, a genetic algorithm is
used to minimize the dierence between a given (target) impedance and the one corresponding to a prescribed
set of APM parameters. Finally, an LBM simulation is performed for the optimal APM in order to educe
the eective value of the impedance to be compared with the target one and thus verify the accuracy of the
best t procedure. As an example, in this study we have tted the impedance of the well-known Ceramic
Tubular liner (CT57) measured by NASA.
41
Fig. 3 shows the comparison between the measured impedance,
its analytical tting, and the educed impedance using an APM with the optimal parameters and a simple
numerical setup similar to a Kundts tube. The comparison between the analytical tting and the educed
values is quite good, meaning that the use of the analytical impedance model is appropriate for a search
involving thousands of designs. More interestingly, the agreement between the educed impedance and the
target experimental impedance is satisfactory, as resulting from a best t over a wide frequency range. The
properties of the equivalent APM made dimensionless by the acoustic wave properties are: R
0
=11.16/f
0
,
0
=0.5742 and d
0
=0.132
0
.
The equivalent surface liner impedance corresponding to the reference frequency of the present study is
6 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

Figure 3. CT73 liner impedance tting with APM. Resistance (left), reactance (right). Experiments (red), APM
analytical tting (blue), LBM educed impedance (black).
extracted from the best t curve plotted in Fig. 3 and used to perform FEM computations in axi-symmetric
modality. The mesh used for the FEM simulation is the one required to run a case at a Helmholtz number
about 12 times higher than the present one, thus guaranteeing a mesh independent reference solution. The
time-averaged ow solution used for the FEM computations is computed by setting slip wall conditions in
the same setup used to perform the LBM acoustic simulations. The free-stream Mach number for the cases
with ow is M

=0.2.
Fig. 4 shows the comparison between the near-eld acoustic pressure at zero cycle computed using LBM
and the reference FEM solution for cases in the absence of convective eects (M

= 0). The agreement


between the two solutions is quite satisfactory. Discrepancies are mainly due to a lack of accuracy in the
space/time synthesis of the acoustic mode prescribed on the fan plane in the LBM simulation. Interestingly,
the accuracy of the LBM solution is preserved up to the FW-H surface, meaning that the employed resolution
is adequate for this type of verication study. It is worthwhile to mention that two dierent visualization
software were used to generate the images for the two solutions, and this can be also responsible for some
discrepancies.
Figure 4. Acoustic pressure. Comparison between LBM solution (top) and FEM solution (bottom). Rigid case (left),
soft case (right).
More quantitative comparisons between LBM and FEM results, both without and with convective eects,
7 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

are shown in Fig. 5, where the noise directivity for both hard and soft wall cases are plotted. Noise was
computed using a FW-H extrapolations to microphones located along an arc of radius equal to 15.6
0
,
centered in the fan plane midpoint and covering an angular sector from 15 deg (forward) to 165 deg (aft). In
30
40
50
60
70
80
90
100
110
120
15 30 45 60 75 90 105 120 135 150 165
S
P
L

[
d
B
]
Observation angle [deg]
LBM - no liner
FEM - no liner
LBM - liner
FEM - liner
30
40
50
60
70
80
90
100
110
120
15 30 45 60 75 90 105 120 135 150 165
S
P
L

[
d
B
]
Observation angle [deg]
LBM - no liner
FEM - no liner
LBM - liner
FEM/GIC-beta=1 - liner
FEM/GIC-beta=0 - liner
Figure 5. Nacelle intake noise directivity. Comparison between LBM and FEM results without (left) and with (right)
convective eects.
the presence of mean ow, two sets of FEM results have been reported, corresponding to the Generalized
Impedance Condition (GIC) of Eq. 29 with blending factor =1 and =0, respectively. As discussed in
more detail in the Appendix, the rst condition corresponds to neglecting any eect due to the presence of
a boundary layer in the FEM computation, whereas the second one corresponds to the well-known Myers
impedance boundary condition,
42
which takes into account the eects due to a boundary layer on the acoustic
propagation past an impedance wall under the assumption of zero-thickness boundary layer.
In the absence of mean ow, discrepancies occur only in the shallow radiation angles where noise is
expected to vanish and where the LBM solution is contaminated by background noise, which is inherent the
LB unsteady solution and prevents to reach the round-o accuracy. A dynamic range of more than 40 dB
is however predicted, and this is quite high for a non-linear CFD solution. Small dierences can be also
observed in the forward interference lobes that might be due to a lack of resolution in the LBM simulation.
Overall the agreement between FEM and LBM solution is good.
In the presence of mean ow, the agreement between LBM and FEM solution for the hard wall case is
again quite satisfactory. In the presence of a lined wall, the LBM solution is in better agreement with the
FEM solution obtained by using a standard impedance condition u
n
= p/Z corresponding to a GIC =1
(Eq. 28). Without entering in the debate of the most appropriate impedance boundary condition for a FEM
solution based on a linearized potential ow model, we can argue that, for the addressed acoustic frequency,
using the Myers boundary condition in the FEM model yields to larger discrepancies between LBM and FEM
results. This trend is conrmed by the velocity and pressure uctuations extracted along a line parallel to
the outer nacelle wall, at a distance of 4.47 10
2
R
f
from the wall, which is in the asymptotic boundary
layer region. The comparisons between the LBM and the two FEM solutions are shown in Fig. 6.
110
115
120
125
130
135
140
0 0.5 1 1.5 2 2.5
S
P
L

[
d
B
]
x/Rf [-]
LBM - liner
FEM/GIC-beta=1 - liner
FEM/GIC-beta=0 - liner
-2
-1.8
-1.6
-1.4
-1.2
-1
-0.8
-0.6
-0.4
0 0.5 1 1.5 2 2.5
lo
g
1
0
(
u

)
x/Rf [-]
LBM - liner
FEM/GIC-beta=1 - liner
FEM/GIC-beta=0 - liner
-4
-3.5
-3
-2.5
-2
-1.5
-1
-0.5
0 0.5 1 1.5 2 2.5
lo
g
1
0
(
v

)
x/Rf [-]
LBM - liner
FEM/GIC-beta=1 - liner
FEM/GIC-beta=0 - liner
Figure 6. Pressure and velocity uctuation magnitudes along the line y=4.47 10
2
R
f
parallel to the intake outer wall.
Comparison between LBM and FEM results in the presence of mean ow.
8 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

III.B. Exhaust fan noise radiation
This subsection is focused on the prediction of noise transmission through a turbofan bypass duct and the
propagation through the shear layer. The LBM solution is compared with frequency domain FEM solution
of a 2nd order wave equation for the acoustic potential,
40
the same wave model used in subsection III.A, and
to the solution of a 3rd order wave equation
35
derived from the Goldstein-Lilley
43
acoustic analogy for the
logarithmic pressure perturbation. It is important to mention that, both FEM solutions are not theoretically
suited to address the present model problem: the acoustic potential equation does not allow to take into
account the Kutta condition at the bypass nozzle edge, whereas the Goldstein-Lilley equation is strictly valid
only for a unidirectional transversely sheared mean ow, which is not the case of the present problem. The
comparison between LBM and FEM results is therefore only made for the sake of completeness, keeping in
mind that the LBM solution is the one without theoretical limitations.
The three-dimensional turbofan bypass geometry considered for both the LBM and FEM computations
is shown in Fig. 7(a). The LBM mesh is shown in Fig. 7(b). The mesh resolution in the jet shear layer is
suciently ne to resolve the time-averaged shear layer ow, but not enough to promote the occurrence of
large scale hydrodynamic instabilities that would result in high jet noise contributions, thus making dicult
to isolate the tonal noise radiation from the bypass exhaust. The acoustic mode (3, 1) is considered, with
a magnitude corresponding to an overall inlet acoustic power of 140 dB. The acoustic mode is introduced
in the LBM simulation through a time-varying pressure/velocity boundary condition on the fan plane.
The choice of this mode is due to the fact that, as discussed in Ref.,
44
this is the most energetic mode
for the fundamental Blade Passing Frequency (BPF) for the ANCF fan rotor-stator conguration used
in subsection III.C. Moreover, since the turbofan conguration simulated in section IV was derived by
inserting the ANCF fan system into a realistic nacelle whose exhaust prole is the same as the one used in
this subsection, a certain consistency among all cases can be preserved by considering the same dominant
spinning mode as for the ANCF conguration at the 1
st
BPF. The Helmholtz number, based on the fan
plane outer radius, is kR
f
=10.37.
(a) Geometry sketch (b) LBM mesh
Figure 7. Turbofan exhaust conguration.
The mean ow solution used for the FEM simulations is obtained by time-averaging the LBM solution
computed using the non-isothermal solver. LBM simulations are performed by prescribing a Mach number
of 0.3 both on the fan plane inlet boundary and on the core jet inlet boundary. The ow temperature on
the fan plane is equal to the free-eld temperature, whereas a temperature ratio of 2.5 is prescribed on the
core jet inlet boundary. A quiescent free-eld is assumed. Fig. 8 shows the contour levels of the velocity
magnitude made dimensionless by the ambient speed of sound (acoustic Mach number) and the temperature
made dimensionless by the ambient temperature (temperature ratio). Quantities useful to estimate the
computational eort for the present case are reported in Table 2.
Fig. 9 show contour levels of the real part of the Fourier component at the fan plane excitation frequency.
A qualitative comparison is made with the two FEM solutions. Both the FEM solution of the Lilley-
Goldstein equation and the LBM solution exhibit hydrodynamic perturbations in the shear layer. However,
9 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

Figure 8. Turbofan exhaust LBM time-averaged ow solution: acoustic Mach number (left) and temperature ratio
(right).
VRs x FEV (10
6
) FES (10
6
) T
t
T
s
CPUh
9
0
/80 13.3 3.2 528/f
0
10/f
0
3563
Table 2. Nacelle exhaust simulation properties.
as expected for a linear wave model, the correctness of the perturbations taking place in the FEM solution
is quite questionable. The uctuations levels in the shear layer for the LBM solution are signicantly
higher and cause a more pronounced interference pattern with the waves radiated from the bypass duct.
Interestingly, the FEM solution of the Pierce equation exhibits a weak discontinuity of the wave fronts along
a direction almost normal to the axial direction, in correspondence of the bypass duct nozzle. Such a feature
is a consequence of the fact that a potential wave model is not able to satisfy the proper condition for the
acoustic pressure at the edge in the presence of a discontinuous mean ow about the edge.
Figure 9. Real part of the pressure uctuation: FEM solution of Pierce equation (left), FEM solution of Lilley-Goldstein
equation (middle) and LBM solution (right).
A more quantitative comparison between LBM and FEM solutions is made in Fig. 10, where the noise
Sound Pressure Level (SPL) directivity plots are shown. For all solutions, the SPL was computed by using
pressure uctuations directly extracted from the near-eld solution, without any integral extrapolation. For
the LBM solution, a very narrow band integration around the tonal frequency was performed to lter out the
jet noise contribution. As depicted in Fig. 7(a), the radiation arc is centered in the midpoint of the bypass
duct outlet section and has a radius of 2.06 R
f
. Interestingly, both for the LBM and the Lilley-Goldstein
FEM solutions, the sound pressure levels up to about 40 deg are dominated by the jet perturbations. Away
from the jet inuence, the LBM and the Lilley-Goldstein FEM radiation pattern are in fairly good agreement,
with the largest discrepancies of about 4 dB taking place in the upward radiation arc. Conversely, the Pierce
FEM solution is in fairly good agreement with the other solutions only in the angular sector between 45 deg
and 75 deg; furthermore, the upward radiation levels are underestimated by about 20 dB, and this is related
to the incorrect treatment of the edge condition in the presence of a discontinuous mean ow about the edge.
This pitfall of an acoustic potential wave model is well known in the literature and inferred to the violation
of the Kutta condition at the nozzle edge.
45
Having stated the physical reliability of the fan noise radiation from a turbofan bypass duct in the presence
of a complex non-isothermal sheared mean ow, it is worthwhile to highlight some interesting feature of the
LBM solution. Although the mesh resolution both in the primary and secondary jets is not ne enough to
accurately predict the jet noise contribution, which is indeed in the scope of subsection III.D, it is interesting
to investigate the eects of the tonal excitation on the shear-layer hydrodynamic instabilities. Excitation
levels of 140 dB are strong enough to promote non-linear eects that can be investigated by computing the
10 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

0
20
40
60
80
100
120
140
160
0 20 40 60 80 100 120 140
S
P
L

[
d
B
]
Observation angle [deg]
LBM/High-Speed Formulation
FEM/Pierce Equation
FEM/Lilley-Goldstein Equation
Figure 10. Exhaust fan noise directivity. Comparison between FEM and LBM solutions.
cross-bicoherence between a reference pressure signal (#1) extracted by a probe located in the laminar core of
the secondary jet, and a pressure signal (#2) extracted at three angular locations along the noise directivity
arc depicted in Fig. 7(a), say 45 deg (downstream), 90 deg and 135 deg. The following formula was used to
compute the cross-bicoherence:

12
(f
i
, f
j
) =
N
w
|

p
1
(f
i
) p
2
(f
j
) p

2
(f
i
+ f
j
)|
2

p
1
(f
i
) p

1
(f
i
)

p
2
(f
j
) p

2
(f
j
)

p
2
(f
i
+ f
j
) p

2
(f
i
+ f
j
)
, (4)
where summations are carried out over N
w
=68 windows with an overlapping coecient of 0.5 and by using a
spectral bandwidth of f
0
/10. Results are plotted in Fig. 11 as function of the harmonic counts. Interestingly,
signicantly high non-linear coherence levels can be observed for the microphone at 45 deg, which is more
strongly aected by the near-eld noise induced by the shear-layer hydrodynamic wave packets. Non-linear
coherence can be also observed in the very low frequency range for the microphone at 90 deg.
Signal #2 at 45 deg
0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
f
j

0.5
1
1.5
2
2.5
3
3.5
4
4.5
5
f
i
0
0.02
0.04
0.06
0.08
0.1
Signal #2 at 90 deg
0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
f
j

0.5
1
1.5
2
2.5
3
3.5
4
4.5
5
f
i
0
0.02
0.04
0.06
0.08
0.1
Signal #2 at 135 deg
0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
f
j

0.5
1
1.5
2
2.5
3
3.5
4
4.5
5
f
i
0
0.02
0.04
0.06
0.08
0.1
Figure 11. Cross-bicoherence 12 as a function of the rst fi/f0 and second harmonic count fj/f0.
III.C. Fan noise generation and radiation
In this subsection we report some results of a validation study recently carried out by simulating the NASA
Glenn ANCF
36
fan conguration with PowerFLOW 5.0c. The simulation setup was derived from an existing
one used in the past to perform pioneering fan noise simulations and analyses.
44, 46
Despite the low Mach
tip operating condition ( 0.33), which is not representative of a real turbofan engine, and the uncertainties
associated with some geometry detail in the experimental setup and to the installation of the so-called Inlet
Control Device (ICD) around the nacelle intake, the ANCF constitutes a very useful benchmark experiment
for the validation of CFD/CAA tools.
11 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

As shown in Fig. 12, the full rotor/stator ANCF geometry, i.e. fan blades, guide vanes, hub and duct are
included in the simulation. The main properties of the ANCF conguration are: a rotor diameter of 1.219 m,
16 rotor blades, 13 guide vanes, 28 deg rotor pitch angle and stator located one stator chord downstream the
rotor. The rotor blades are fully embedded in a rotating mesh volume, as indicated in Fig. 12.
Figure 12. Sketch of the ANCF computational setup.
Three main dierences between the simulated geometry and the experimental conguration should be
pointed out: no tip clearance is simulated, the outlet shaft system is not included in the simulation model
and the eects due to the presence of the ICD on both the ow eld and the far-eld noise are not modelled.
The nominal operating condition is simulated, corresponding to 1800 rpm. The overall simulation size is
summarized in Table. 3.
VRs x[ mm] FEV (10
6
) FES (10
6
) T
t
[ s] T
s
[ s] CPUh (10
3
)
11 0.65 114.3 33.4 0.333 0.5 25
Table 3. ANCF simulation properties.
As depicted in Fig. 12, two permeable surfaces, which encompass the nacelle intake and exhaust, are
used to perform far-eld FW-H computations. These surface are fully enclosed in a mesh resolution level
VR8 (three coarsening levels with respect to the nest VR11 one) that guarantees a space/time resolution
of 20 points per acoustic wavelength and 35 samples per period, respectively, at the 7th BPF harmonic
(3.36 kHz). Far-eld noise was computed at 15 microphones located on a forward arc, which covers the
angular sector [0 deg : 90 deg] (Mic. # 1 located at 0 deg), and at 15 microphones on an aft arc covering the
sector [90 deg : 160 deg]; the forward arc has a radius of 3.66 m and is centered in the midpoint of the intake
section, whereas the aft arc has a radius of 3.66 m and is centered in the midpoint of the exhaust section.
The time-averaged pressure and Mach number eld solutions are shown in Fig. 13. The highest values
of Mach number are achieved in proximity of the rotor blades. The ow compression induced by the fan
system is clearly evident in the pressure eld. The ow eld about the nacelle lip is very regular and does
not reveal any ow separation.
Fig. 14 shows an instantaneous view of the time derivative pressure eld, which is quite eective in
visualizing the instantaneous acoustic waves. Very interestingly, a dominant wave front takes place in the
acoustic eld radiated from the nacelle intake, and the radiation direction seems to be related to the main
tonal content (fundamental BPF). Conversely, the noise radiated from the nacelle exhaust exhibits at least
two dominant wave fronts, one associated with the fundamental BPF, the other one with 2
nd
BPF. This
is a clear diraction eect from the sharp edge of the exhaust nozzle. Another interesting feature of the
instantaneous acoustic pressure eld is the complex interference pattern taking place between the waves
radiated from the intake and exhaust openings.
Noise spectra at two angular positions are plotted in Fig. 15. Both for the measured and computed
acoustic pressure signals, Fourier transforms were computed with a bandwidth equal to 30 Hz; spectra are
plotted in terms of BPF harmonic count. Some discrepancies can be observed between numerical and
experimental results, both in the tonal and in the broadband spectral components. More precisely, the
broadband components exhibit a systematic underestimation ( 5 dB), which could be due to two reasons:
a lack of mesh resolution in the rotor/stator region, resulting in more coherent turbulent structures in the
rotor wake, and the absence, in the simulation, of the rotor tip clearance, where interactions take place
12 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

Figure 13. ANCF simulation: time-averaged pressure eld (left) and Mach number (right). Front view on the bottom
extracted from the rotating mesh, cutting through the rotor.
Figure 14. ANCF simulation: snapshot of pressure time derivative in one azimuthal plane.
13 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

between the nacelle boundary layer turbulent structures and the rotor blades. It is important to mention
that the ANCF is characterized by a predominant tonal character, with tones arising by about 30 40 dB
from the broadband levels, and that it is dicult to cover such a dynamic range in CFD simulations.
20
40
60
80
100
0 1 2 3 4
P
S
D

[
d
B
/
H
z
]
BPF harmonic count [-]
Measurements
LBM simulation
20
40
60
80
100
0 1 2 3 4
P
S
D

[
d
B
/
H
z
]
BPF harmonic count [-]
Measurements
LBM simulation
Figure 15. ANCF simulation: narrow-band noise spectra at two angular locations: Mic. #11 (left) and Mic. #23
(right). Comparison between measured and computed results.
Fig. 16 shows a comparison between measured and predicted noise directivity patterns for the rst two
BPF harmonics and for the Overall SPL (OASPL) covering a broad frequency range. It is important to
mention that, for the sake of computational eciency, the noise signals along the forward and aft arcs
have been computed by performing separate FW-H integrations on the two permeable surfaces depicted in
Fig. 12, thus neglecting the interference eects between the intake and exhaust radiation that are expected
to have a non negligible eect only around 90 deg. The numerical predictions exhibit a systematic OASPL
overestimation that is due to an overestimation of the tonal peaks. This is conrmed by the BPF SPL
directivity plots of Fig. 16. The overestimation of the tonal components can be also related to a lack
of resolution in the rotor/startor region, which results in a too deterministic rotor-wake/stator-vanes
interaction. It is interesting to observe that the discrepancies between measurements and predictions are
higher in the forward direction ( 5 dB) compared to the aft direction ( 3 dB). This can be related to the
presence of the ICD in the experiments, which results in a noise transmission loss from the nacelle intake to
the microphones. Indeed, the measured forward noise levels are 2 3 dB lower than the measured aft levels.
The reader is remanded to Ref.
44
for other comparisons between measurements and predictions.
55
60
65
70
75
80
85
90
95
100
105
0 20 40 60 80 100 120 140 160 180
S
P
L

[
d
B
]
Observation angle [deg]
SPL - 1st BPF (480 Hz)
Measurements - forward arc
LBM simulation - forward arc
Measurements - aft arc
LBM simulation - aft arc
55
60
65
70
75
80
85
90
95
100
105
0 20 40 60 80 100 120 140 160 180
S
P
L

[
d
B
]
Observation angle [deg]
SPL - 2nd BPF (960 Hz)
Measurements - forward arc
LBM simulation - forward arc
Measurements - aft arc
LBM simulation - aft arc
55
60
65
70
75
80
85
90
95
100
105
0 20 40 60 80 100 120 140 160 180
S
P
L

[
d
B
]
Observation angle [deg]
OASPL in band [450 Hz : 1470 Hz]
Measurements - forward arc
LBM simulation - forward arc
Measurements - aft arc
LBM simulation - aft arc
Figure 16. ANCF simulation: SPL noise directivity along forward and aft microphone arcs in dierent frequency bands.
Comparison between measured and computed results.
III.D. Jet noise generation and radiation
The goal of this subsection is to validate the jet noise prediction capabilities of the high-speed non-isothermal
LBM formulation implemented in a beta release version of PowerFLOW 5.1. The NASA Glenn research
center hot jet experiment
37
performed by using the SMC000 nozzle was modelled for two ow conditions:
setpoint 07, characterized by an acoustic Mach number M
a
= U
j
/c

= 0.902 and a temperature ratio


T
r
= T
j
/T

= 0.842 (cold jet), and setpoint 46, for which M


a
= 0.901 and T
r
= 2.702 (hot jet). The
corresponding nominal centerline exit Mach numbers are 0.983 and 0.548, respectively.
14 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

Fig. 17 illustrates the computational setup. A total pressure inlet boundary condition is imposed at
the nozzle inlet, with a prescribed value of the temperature computed from the experimental value at the
nozzle exit centerline by using isentropic relationships. The external boundary of the nozzle is extended up
to the far-eld upstream boundary through a conic surface. The nozzle is embedded in a computational
anechoic environment consisting of three layers of acoustic sponge regions. The FW-H permeable surface
is fully embedded in a VR10 mesh resolution level (two coarsening levels with respect to the nest VR12
one), which has a grid cut-o frequency estimated around 20 kHz (20 voxels per wavelength). The cup of the
FW-H surface is included in the sampled ow le, but it is excluded from the FW-H integration in order to
avoid the generation of ctitious noise uctuations due to the strong vortical uctuations passing through
the surface. A view of the computational mesh is illustrated in Fig. 17, showing the dierent VR levels
around the nozzle and in the jet plume. A VR11 oset region has been generated from all the internal walls
of the nozzle, whereas a VR12 (nest one) oset region has been generated from a small section of the nozzle
close to the exit. The overall simulation size is summarized in Table. 4. Other details about the jet noise
validation setup are reported in a companion paper.
47
Figure 17. SMC000 simulation: computational setup (left) and mesh (right).
VRs x[ mm] FEV (10
6
) FES (10
6
) T
t
[ s] T
s
[ s] CPUh (10
3
)
13 0.198 94.3 2.8 0.150 0.191 51
Table 4. SMC000 jet noise simulation properties.
Figs. 18 and 19 show time-averaged eld results for setpoint 07 and 46, respectively. The streamwise
velocity standard deviation was computed by making an isotropic turbulence assumption for the small
unresolved turbulence scale, thus adding
_
2K/3 to the resolved uctuation levels. All quantities, including
the velocity standard deviation exhibit a regular and symmetric pattern, and this is a qualitative indication
of simulation accuracy and statistical convergence.
Figure 18. SMC000 simulation: setpoint 07.
37
Mach number (left), temperature ratio (middle) and dimensionless
streamwise velocity standard deviation (right).
A quantitative analysis of the LBM solution accuracy is performed by comparing the time-averaged
streamwise velocity and its standard deviation levels levels along the jet centerline. Results made dimen-
15 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

Figure 19. SMC000 simulation: setpoint 46.
37
Mach number (left), temperature ratio (middle) and dimensionless
x-velocity standard deviation (right).
sionless by the exit centerline velocity U
j
are plotted in Figs. 19 and 20 for setpoint 07 and 46, respectively.
For both setpoints, the extension of the predicted laminar core length X
w
and the turbulent mixing are in
very good agreement with value estimated by using the formulas put forward by Witze
48
and used in Ref.,
49
i.e.:
X
w
= 4.375 D
j
(
j
/

)
0.28
1 0.16 M
j
, and
U
U
j
= 1 e
1,43/(1x/Xw)
. (5)
It is important to mention that no tuning of the turbulent levels prescribed at the inlet nozzle boundary
condition was performed to get the proper value of the laminar core length, which is a genuine result
of the simulations; moreover, no random forcing at the nozzle inlet was employed to seed the turbulence
uctuations in the nozzle boundary layer, which develops naturally along the nozzle walls, thus providing the
proper boundary layer integral quantities at the nozzle exit. Therefore, the accuracy of the time-averaged
centerline velocity is a good indication of the proper bahavior of the turbulence model in the shear layer, as
well as in the wall region inside the nozzle. The streamwise velocity standard deviation velocity along the jet
centerline is compared with the curve plotted in Fig.(36) of Ref.
49
The agreement is very good for setpoint
46, whereas a signicant overestimation of the uctuation levels in the laminar core region is predicted
for setpoint 07. It should be mentioned that setpoint 07 is characterized by almost sonic conditions, thus
constituting a challenging test case for the high-speed D3Q19-based LBM ow solver. At the present stage
of the research, it is not clear whether the overestimation of the uctuation levels for setpoint 07 is due to the
computational setup or to the fact that we approached the usage limits of the high-speed LB formulation.
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
0 1 2 3 4 5
u

/

U
j

[
-
]
x / D [-]
Witze formula
LBM simulation
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0.16
0 1 2 3 4 5
u

/

V
j

[
-
]
x / Xw [-]
Measurements
LBM simulation
Figure 20. SMC000 simulation: setpoint 07.
37
Dimensionless centerline time-averaged (left) and standard deviation
x-velocity (right).
Finally, noise results are reported in Figs. 22 and 23 for setpoint 07 and 46, respectively. The computed
third-octave SPL spectra at 90 deg and 135 deg (downstream) are compared with measurements by Brown
& Bridges.
37
The estimated grid cuto frequency for the employed FW-H surface is about 20 kHz, and this
seems to be conrmed by the rapid drop-o of the noise levels around that frequency for both cases. The
low-frequency part of the predicted noise spectra exhibit a certain lack of statistical convergence. The noise
directivity for setpoint 07 is predicted within 1 dB accuracy up to 135 dB. For higher value (downstream)
16 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
1.1
0 1 2 3 4 5
u

/

U
j

[
-
]
x / D [-]
Witze formula
LBM simulation
0
0.02
0.04
0.06
0.08
0.1
0.12
0.14
0.16
0 1 2 3 4 5
u

/

V
j

[
-
]
x / Xw [-]
Measurements
LBM simulation
Figure 21. SMC000 simulation: setpoint 46.
37
Dimensionless centerline time-averaged (left) and standard deviation
x-velocity (right).
noise levels are signicantly underestimated, with the maximum predicted levels about 5 dB lover than in
the experiments. This is due to the fact that we omitted the cup of the permeable surface from the FW-H
integration, thus neglecting part of the noise contribution generated by the largest wave packets that are
very eective at shallow radiation angles, in particular in almost sonic conditions. The same trend can be
observed for setpoint 46 in Fig. 23, but the underestimation of the maximum noise levels is about 2 dB.
Further work is required in order to develop a procedure that allows to extend the FW-H integration to the
surface cup.
55
60
65
70
75
80
85
90
95
100 1000 10000 100000
S
P
L

[
d
B
]
Frequency [Hz]
Measurements - 135 deg
LBM simulation - 135 deg
Measurements - 90 deg
LBM simulation - 90 deg
92
94
96
98
100
102
104
106
40 60 80 100 120 140 160 180
S
P
L

[
d
B
]
Observation angle [deg]
Measurements
LBM simulation
Figure 22. SMC000 simulation: setpoint 07.
37
One-third octave band SPL (left) and OASPL directivity (right).
IV. Prediction of turbofan engine noise
The previous section focused on the fundamental validation of the major turbofan components responsible
for noise generation. In this section the simulation results of a complete turbofan demonstrator geometry
with bypass and a rotor-stator fan/OGV conguration are discussed. These results are to be understood
as a capability demonstration to address a comprehensive turbofan noise problem, taking simultaneously in
one single simulation all the noise generation mechanisms into account that were validated separately in the
previous sections. The simulation was performed by using a beta release of PowerFLOW 5.1, which provides
the high-speed non-isothermal functionality.
17 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

50
55
60
65
70
75
80
85
90
95
100 1000 10000 100000
S
P
L

[
d
B
]
Frequency [Hz]
Measurements - 135 deg
LBM simulation - 135 deg
Measurements - 90 deg
LBM simulation - 90 deg
88
90
92
94
96
98
100
102
104
40 60 80 100 120 140 160 180
S
P
L

[
d
B
]
Observation angle [deg]
Measurements
LBM simulation
Figure 23. SMC000 simulation: setpoint 46.
37
One-third octave band SPL (left) and OASPL directivity (right).
IV.A. Geometry and ow conditions
The geometry shown in Fig. 24 was designed by considering a generic shape for the nacelle with a realistic full
scale outer diameter of about 3.5 m and a length of about 5.2 m. The core nozzle and the bypass diameters
were chosen to be about 1.5 m and 2.9 m, respectively. The fan and stator geometries are taken from the
ANCF simulations described previously in subsection III.C with a geometric scale factor of 2.3 to match the
overall geometrical size of the chosen turbofan.
Figure 24. Geometry of the complete turbofan demonstrator.
The simulations were conducted at a free-stream Mach number M

=0.1 and standard sea level atmo-


spheric free-stream conditions for temperature, pressure, and molecular viscosity. The core jet was chosen
to be at M
j
=0.424 and a temperature ratio of 2 compared to the ambient value. The fan RPM was chosen
at around 1400 resulting in a nominal fan tip Mach number of about 0.6 and a BPF of 376.6 Hz.
IV.B. Computational setup
The geometry was simulated in a free air condition with inow/outow BC located at 200 nacelle diameters
away. In addition, three sponge layers were included to damp out acoustic uctuations and reections
from the far eld boundary conditions. The rotating fan blades and hub were enclosed in a sliding truly
rotating mesh region, as described in subsection III.C. This demonstrator simulation was conducted at a
comparatively coarse resolution in relation to the previous sections, but great care was taken in locally
resolving the fan blades, interaction regions between rotor and stator, as well as the two stream jet shear
layers. The automatically generated volume mesh is depicted in Fig. 25 and clearly highlights the regions
of local renement and the extended far eld ne resolution used to directly resolve the acoustic wave
18 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

propagation with sucient resolution up to the porous surface, where the ow data are collected for the
subsequent FW-H computations.
Figure 25. Volume mesh, near eld (left) and far eld (right) with every second line shown
The overall simulation size is summarized in Table. 5. The simulated physical time corresponds to 2 s
with 0.2 s at the start used for the settling time for the ow after which the simulation data is used for
statistical analysis, far eld propagation and further postprocessing. The history of the fan torque and the
pressure at a near eld direct probe 25 nozzle diameters away in the downstream direction documented in
Fig. 26 indicates that the choice of settling time is adequate for the statistical analysis. With the choice of
simulation time and resolution the expected numerically resolved frequency range is from f
min
20 Hz to
f
max
1 kHz taking into account 20 spectral averages for the lowest frequency and 15 voxels to resolve
the propagation to the porous FW-H surface of the wave length at the high frequency.
Figure 26. Fan torque time history (left) and exemplary near eld uid probe pressure history (right)
VRs x[ mm] FEV (10
6
) FES (10
6
) T
t
[ s] T
s
[ s] CPUh (10
3
)
14 2.98 168 10.5 0.2 1.8 54
Table 5. Turbofan simulation properties.
IV.C. Results and discussion
A qualitative view of the ow structures are depicted in Fig.27. The snapshot of the ow clearly highlights
the small vortical structure emanating from the fan blade tips and their interactions with the stator as
well as the large and small structures generated in the highly unsteady region of the two stream jet shear
19 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

layers. The large structures in the jets are quickly breaking up into smaller structures and indicate that the
chosen resolution of the simulation was sucient to capture the unsteady structures of the jet ow. This can
also be seen in the temperature eld shown in Fig.28 of the spreading of the hot core jet. Furthermore the
dilatation eld achieved by dierentiating the pressure eld over two subsequent unsteady time frames shown
in Fig.28 visually illustrates the sound propagation as well as the directivity of the acoustics in addition to
the hydrodynamic pressure uctuations in the jet shear layers. The pressure eld can be also ltered around
the BPF as illustrated in Fig. 29 to qualitatively visualize the shape of the pressure eld at that specic
frequency and its dominant modal content. An azimuthal mode order equal to 3 is clearly visible.
Figure 27. Snapshot of 2 iso-surfaces colored by velocity magnitude.
Figure 28. Snapshot of temperature eld (left) and dilatation eld (right).
A far eld propagation using FW-H was carried out based on the data collected on the porous surface
illustrated in Fig. 30. The propagation is performed to several far eld microphones located at 120 m distance,
but only three located at 45 deg, 90 deg and 135 deg are discussed here as representative of forward, y-over
and backward directions. The FW-H surface is also subdivided into an intake region surrounding only the
intake part of the turbofan and an exhaust region surrounding the complete exhaust and the large jet region
of the turbofan engine.
The total noise at the three microphone locations is documented in Fig. 31. As expected the downstream
probe at 135 deg has an increased level of broadband noise due to its proximity to the jet ow, whereas the
20 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

Figure 29. Narrow bandpass ltered pressure around the 1
st
BPF harmonic across the fan inlet face through the nacelle
(left) and on the nacelle surface (right).
Figure 30. Permeable FW-H surface around the turbofan and far-eld microphone locations.
21 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

tonal noise contributions from the rst 5 BPF harmonics are present at the expected frequencies in both
forward and backward microphone locations. The propagation of the tonal modes through the bypass ow
at the exhaust part of the ow can also be seen in the Fig. 30.
Figure 31. Noise PSD at three far-eld microphone locations.
It is interesting to investigate the breakdown of the contributions to the noise based only on the intake
and the exhaust FW-H surfaces as juxtaposed in Fig. 32. The results follow the expectation that the intake
part is greatly contributing to the tonal parts of the noise, whereas the exhaust region is generally much
higher in broadband contributions except for the backward microphone at 135 deg, where the contribution
from the tonal part propagating through the nacelle is additionally observed. This comparison can be
done for each microphone separately as depicted in Fig. 33, again highlighting the same behavior for each
microphone. This type of analysis is very helpful to generally distinguish contributions and hence potential
noise sources based on the geometrical region and type of noise generation mechanism in the far eld. From
an engineering perspective, the presented tools are capable of not only predicting the noise generation of a full
turbofan engine, but also guide potential designers in better understanding of the dierent noise generation
mechanisms and subsystem contributions.
Figure 32. Contribution break-down of the noise spectra at three microphones when using only the intake FW-H region
(left) and the exhaust FW-H region (right).
22 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

Figure 33. Comparison of intake and exhaust FW-H surface contributions to far eld microphones at 45 deg (left),
90 deg (middle) and 135 deg (right).
V. Conclusion
A comprehensive verication and validation study of PowerFLOW CFD/CAA solver capabilities in the
eld of turbofan noise prediction was carried out. Emphasis was given to the overall view of the problem,
and not on the specic details of the dierent elementary case studies. The usage envelope of the LBM in
this new application eld was outlined, highlighting both potentialities and areas of further improvement.
The main outcomes of this eort are summarized hereafter.
- Fan noise transmission through and radiation from the nacelle intake and bypass duct, also in the presence
of an acoustic liner, can be accurately predicted. For the intake radiation problem, other techniques based
on the solution of simple wave models in the frequency-domain are certainly more convenient. However,
whenever geometrical three-dimensionality and nacelle installation eects are in the scope of the analysis,
integrated engine/airframe LBM simulations are one of the few exploitable computational technologies. For
the bypass duct radiation problem, linear wave models have limited usage potentialities and the solution
of the full non-linear ow governing equations or, more continently, LBM simulations, are denitively more
reliable.
- The eect of noise absorption due to an acoustic treatment in the presence of grazing ow can be modelled
by means of a specically tuned APM. A primary validation of this new capability in PowerFLOW was
performed. However, further studies are required to assess the capability of reproducing the absorption
properties of a wider range of acoustic treatments, including honeycomb liners, as well as to understand the
relationship between boundary layer properties and acoustic properties of the porous material. Space-/time-
resolved ow/noise measurements are required to accomplish this goal.
- Tonal and broadband fan noise generation can be predicted by LBM. However, the outlined uncertainty
range of 3-to-5 dB compared to experimental results for the addressed NASA Glenn ANCF conguration,
is not fully satisfactory. Further work is required to reduce this uncertainty range, by improving the com-
putational setup on the simulation side, and by quantifying the uncertainties related to the test rig and the
required corrections on the experimental side.
- Jet noise can be predicted by the non-isothermal version of the LBM up to almost sonic conditions, by
taking into account the primary jet heating eects. Further work is required to optimize the computational
setup and reduce the simulation eorts for a prescribed accuracy level. Areas for further improvement were
found in the low-frequency range and in the downstream radiation; longer physical simulation times will be
tested in the future, as well as the inclusion of the permeable surface cup in the FW-H integration.
A qualitative demonstration of PowerFLOW integrated fan/jet noise prediction capabilities was nally
achieved and constitutes a primary milestone towards future lattice-Boltzmann prediction of turbofan engine
noise.
Acknowledgments
The authors are grateful to James Bridges and to Daniel L. Sutli at NASA Glenn Research Center for
providing the SMC000 nozzle geometry and the detailed ANCF experimental data needed for the validation.
23 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

References
1
Casalino, D., Ribeiro, A. F. P., and Fares, E., Facing Rim Cavities Fluctuation Modes, Journal of Sound and Vibration,
Vol. 333, No. 13, 2014, pp. 28122830.
2
Casalino, D., Ribeiro, A. F. P., Fares, E., and Nolting, S., Lattice-Boltzmann Aeroacoustic Analysis of the LAGOON
Landing Gear Conguration, AIAA Journal , Vol. 52, No. 6, 2014, pp. 12321248.
3
Murayama, M., Yokokawa, Y., Imamura, T., Yamamoto, K., Ura, H., and Hirai, T., Numerical Investigation on Change
of Airframe Noise by Flap Side-edge Shape, AIAA Paper 2013-2067, May 2013.
4
Br`es, G. A., Freed, D. M., Wessels, M., Noelting, S., and Perot, F., Flow and Noise Predictions for the Tandem Cylinder
Aeroacoustic Benchmark, Physics of Fluids, Vol. 24, No. 3, March 2012, pp. 036101.
5
Simoes, L. G. C., Souza, D. S., and Medeiros, M. A. F., On the Small Eect of Boundary Layer Thicknesses on Slat
Noise, AIAA Paper 2011-2906, June 2011.
6
Casalino, D., Noelting, S., Fares, E., Van de Ven, T., Perot, F., and Br`es, G. A., Towards Numerical Aircraft Noise
Certication: Analysis of a Full-Scale Landing Gear in Fly-Over Conguration, AIAA Paper 2012-2235, June 2012.
7
Khorrami, M. R., Fares, E., and Casalino, D., Towards Full Aircraft Airframe Noise Prediction: Lattice Boltzmann
Simulations, AIAA Paper 2014-xxxx, June 2014.
8
Perot, F., Freed, D. M., and Mann, A., Acoustic Absorption of Porous Materials Using LBM, AIAA Paper 2013-2070,
May 2013.
9
P.L.Bhatnagar, Gross, E., and Krook, M., A Model for Collision Processes in Gases. I. Small Amplitude Processes in
Charged and Neutral One-Component Systems, Physical Review, Vol. 94, No. 3, 1954, pp. 511525.
10
Chen, H., Chen, S., and Matthaeus, W., Recovery of the Navier-Stokes Equations Using a Lattice-Gas Boltzmann
Method, Physical Review A, Vol. 45, No. 8, 1992, pp. 53395342.
11
Chen, H. and Teixeira, C. M., H-Theorem and Origins of Instability in Thermal Lattice Boltzmann Models, Computer
Physics Communications, Vol. 129, 2000, pp. 2131.
12
Chapman, S. and Cowling, T., The Mathematical Theory of Non-Uniform Gases, Cambridge University Press, 1990.
13
Shan, X., Yuan, X.-F., and Chen, H., Kinetic Theory Representation of Hydrodynamics: a Way Beyond the Navier-
Stokes Equation, Journal of Fluid Mechanics, Vol. 550, 2006, pp. 413441.
14
Chen, H., Kandasamy, S., Orszag, S. A., Succi, S., and Yakhot, V., Extended Boltzmann Kinetic Equation for Turbulent
Flows, Science, Vol. 301, No. 5633, 2003, pp. 633636.
15
Yakhot, V. and Orszag, S. A., Renormalization Group Analysis of Turbulence. I. Basic Theory, Journal of Scientic
Computing, Vol. 1, No. 1, 1986, pp. 351.
16
Yakhot, V., Orszag, S. A., Thangam, S., Gatski, T. B., and Speziale, C. G., Development of Turbulence Models for
Shear Flows by a Double Expansion Technique, Physics of Fluids A, Vol. 4, No. 7, 1992, pp. 15101520.
17
Teixeira, C. M., Incorporating Turbulence Models into the Lattice-Boltzmann Method, International Journal of Modern
Physics C, Vol. 9, 1998, pp. 11591175.
18
Alexander, C. G., Chen, H., Kandasamy, S., Shock, R., and Govindappa, S. R., Simulations of Engineering Thermal
Turbulent Flows Using a Lattice Boltzmann Based Algorithm, ASME-PUBLICATIONS-PVP, Vol. 424, 2001, pp. 115126.
19
Anagnost, A., Alajbegovic, A., Chen, H., Hill, H., Teixeira, C., and Molvig, K., Digital Physics Analysis of the Morel
Body in Ground Proximity, SAE Paper 970139, 1997.
20
Chen, H., Teixeira, C. M., and Molvig, K., Digital physics approach to computational uid dynamics: some basic
theoretical features, International Journal of Modern Physics C, Vol. 8, No. 4, 1997, pp. 675684.
21
Chen, H., Teixeira, C., and Molvig, K., Realization of Fluid Boundary Conditions via Discrete Boltzmann Dynamics,
International Journal of Modern Physics C, Vol. 9, No. 8, 1998, pp. 12811292.
22
Br`es, G. A., Perot, F., and Freed, D. M., A Ffowcs Williams-Hawkings Solver for Lattice-Boltzmann Based Computa-
tional Aeroacoustics, AIAA Paper 2010-3711, June 2010.
23
Di Francescantonio, P., A New Boundary Integral Formulation for the Prediction of Sound Radiation, Journal of Sound
and Vibration, Vol. 202, No. 4, 1997, pp. 491509.
24
Casalino, D., An Advanced Time Approach for Acoustic Analogy Predictions, Journal of Sound and Vibration, Vol. 261,
No. 4, 2003, pp. 583612.
25
Farassat, F. and Succi, G. P., The Prediction of Helicopter Discrete Frequency Noise, Vertica, Vol. 7, No. 4, 1983,
pp. 309320.
26
Li, Y., Shock, R., Zhang, R., and Chen, H., Numerical Study of Flow Past an Impulsively Started Cylinder by Lattice
Boltzmann Method, Journal of Fluid Mechanics, Vol. 519, 2004, pp. 273300.
27
Fares, E., Unsteady Flow Simulation of the Ahmed Reference Body using a Lattice Boltzmann Approach, Journal of
Computers and Fluids, Vol. 35, 2006, pp. 940950.
28
Fares, E., jelic, S., Kuthada, T., and Schroeck, D., Lattice Boltzmann Thermal Flow Simulation and Measurements of
a modied SAE Model with Heated Plug, Proceedings of FEDSM2006-98467, 2006.
29
Nie, X., Shan, X., and Chen, H., A Lattice-Boltzmann/Finite-Dierence Hybrid Simulation of Transonic Flow, AIAA
Paper 2009-0139, 2009.
30
Freed, D. M., Lattice-Bolzmann Method for Macroscopic Porous Media Modeling, International Journal of Modern
Physics, Vol. 9, 1998, pp. 14911505.
31
Sun, C., Perot, F. L., Zhang, R., Chen, H., Freed, D. M., and Staroselsky, I., Computer Simulation of Physical Processes,
May 9 2013, US Patent App. 13/292,844.
32
Mann, A., Perot, F., Kim, M.-S., and Casalino, D., Characterization of Acoustic Liners Absorption using a Lattice-
Boltzmann Method, AIAA Paper 2013-2271, May 2013.
24 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

33
Casalino, D., Benchmarking of Dierent Wave Models for Sound Propagation in Non-Uniform Flows, Proceedings of
the IUTAM Symposium on Computational Aero-Acoustics for Aircraft Noise Prediction, 29-31 March 2010, University of
Southampton, UK, Institute of Sound and Vibration Research, University of Southampton, Elsevier, Procedia Engineering,
March 2010, pp. 163172.
34
Casalino, D. and Barbarino, M., Stochastic Method for Airfoil Self-Noise Computation in Frequency Domain, AIAA
Journal , Vol. 49, No. 11, 2011, pp. 24532469.
35
Casalino, D., Finite Element Solutions of a Wave Equation for Sound Propagation in Sheared Flows, AIAA Journal ,
Vol. 50, No. 1, 2012, pp. 3745.
36
Sutli, D. L., Rotating Rake Turbofan Duct Mode Measurement System, NASA TM-2005-213828, 2005.
37
Brown, C. and Bridges, J., Small Hot Jet Acoustic Rig Validation, NASA/TM-2006-214234, 2006.
38
Casalino, D., Barbarino, M., and Visingardi, A., Simulation of Helicopter Community Noise in Complex Urban Geom-
etry, AIAA Journal , Vol. 48, No. 8, 2011, pp. 16141624.
39
Casalino, D. and Barbarino, M., Optimization of a Single-Slotted Lined Flap for Wing Trailing-Edge Noise Reduction,
AIAA Journal of Aircraft, Vol. 49, No. 4, 2012, pp. 10511063.
40
Pierce, A. D., Wave Equation for Sound in Fluids with Unsteady Inhomogeneous Flow, Journal of the Acoustical
Society of America, Vol. 87, No. 6, 1990, pp. 22922299.
41
Parrott, T. L., Watson, W. R., and Jones, M. G., Experimental Validation of a Two-Dimensional Shear-Flow Model for
Determining Acoustic Impedance, NASA Technical Paper 2679, 1987.
42
Myers, M. K., On the Acoustic Boundary Condition in the Presence of Flow, Journal of Sound and Vibration, Vol. 71,
No. 3, 1980, pp. 429434.
43
Goldstein, M. E., An Exact Form of Lilleys Equation with a Velocity Quadrupole/Temperature Dipole Source Term,
Journal of Fluid Mechanics, Vol. 443, 2001, pp. 231236.
44
Mann, A., Perot, F., Kim, M.-S., Casalino, D., and Fares, E., Advanced Noise Control Fan Direct Aeroacoustics
Predictions using a Lattice-Boltzmann Method, AIAA Paper 2012-2287, June 2012.
45
Gabard, G. and Astley, R. J., Theoretical Model for Sound Radiation from Annular Jet Pipes: Far- and Near-Field
Solutions, Journal of Fluid Mechanics, Vol. 549, 2006, pp. 315341.
46
Perot, F., Mann, A., Kim, M.-S., , and Casalino, E. F. D., Investigation of inow condition eects on the ANCF
aeroacoustics radiation using LBM, InterNoise, New York, 2012.
47
Lew, P.-T., Shock, R., and Casalino, D., An Extended Lattice Boltzmann Methodology for High Subsonic Jet Noise
Prediction, AIAA Paper 2014-XXXX, 2014.
48
Witze, P. O., Centerline Velocity Decay of Compressible Free Jets, AIAA Journal , Vol. 12, No. 4, 1974, pp. 417418.
49
Bridges, J. and Wernet, M., Establishing Consensus Turbulence Statistics for Hot Subsonic Jets, AIAA Paper 2010-
3751, 2010.
50
Taylor, K., Acoustic Generation by Vibrating Bodies in Homentropic Potential Flow at Low Mach Number, Journal
of Sound and Vibration, Vol. 65, No. 1, 1979, pp. 125136.
51
Ingard, U., Inuence of Fluid Motion Past a Plane Boundary on Sound Reection, Absorption, and Transmission,
Journal of the Acoustical Society of America, Vol. 31, No. 7, 1959, pp. 10351036.
52
Rienstra, S. W., Sound Transmission in Slowly Varying Circular and Annular Lined Ducts with Flow, Journal of Fluid
Mechanics, Vol. 380, 1999, pp. 279296.
53
Rienstra, S. W. and Eversman, W., A Numerical Comparison Between the Multiple-Scales and Finite-Element Solution
for Sound Propagation in Lined Flow Ducts, Journal of Fluid Mechanics, Vol. 437, 2001, pp. 367384.
54
Astley, R. J. and Gamallo, P., Partition of Unity Finite Element Method for Short Wave Acoustic Propagation on
Nonuniform Potential Flows, AIAA Paper 2004-2893, May 2004.
55
Casalino, D., Roger, M., and Jacob, M., Prediction of Sound Propagation in Ducted Potential Flows Using Greens
Function Discretization, AIAA Journal , Vol. 42, No. 4, 2004, pp. 736744.
56
Brambley, E. J. and Peake, N., Surface-Waves, Stability, and Scattering for a Lined Duct with Flow, AIAA Paper
2006-2688, May 2006.
57
Casalino, D. and Genito, M., Achievements in the Numerical Modelling of Fan Noise Radiation from Aero-engines,
Aerospace Science and Technology, Vol. 12, 2008, pp. 105113.
58
Tam, C. K. W. and Auriault, L., Time-Domain Impedance Boundary Conditions for Computational Aeroacoustics,
AIAA Journal , Vol. 34, No. 5, 1996, pp. 917923.
59
Chevaugeon, N., Remacle, J.-F., and Gallez, X., Discontinuous Galerkin Implementation Of The Extended Helmholtz
Resonator Model In Time Domain, AIAA Paper 2006-2569, May 2006.
60
Brambley, E. J. and Peake, N., Classication of Aeroacoustically Relevant Surface Modes in Cylindrical Lined Ducts,
Wave Motion, Vol. 43, 2006, pp. 301310.
61
Brambley, E. J., Fundamental Problems with the Model of Uniform Flow over Acoustic Linings, Journal of Sound and
Vibration, Vol. 322, No. 4-5, 2009, pp. 10261037.
62
Brambley, E. J., A Well-posed Modied Myers Boundary Condition, AIAA Paper 2010-3942, June 2010.
63
Rienstra, S. W. and Darau, M., Boundary-Layer Thickness Eects of the Hydrodynamic Instability Along an Impedance
Wall, Journal of Fluid Mechanics, Vol. 671, 2011, pp. 559573.
64
Auregan, Y., Starobinski, R., and Pagneux, V., Inuence of Grazing Flow and Dissipation Eects on the Acoustic
Boundary Conditions at a Lined Wall, Journal of the Acoustical Society of America, Vol. 109, No. 1, 2001, pp. 5964.
65
Brambley, E. J., Viscous Boundary Layer Eects on the Myers Impedance Boundary Condition, AIAA Paper 2009-
3241, May 2009.
66
Renou, Y. and Auregan, Y., Failure of the IngardMyers Boundary Condition for a Lined Duct: An Experimental
Investigation, Journal of the Acoustical Society of America, Vol. 130, No. 1, 2011, pp. 5260.
25 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

67
Morse, P. M. and Ingard, K. U., Theoretical Acoustics, McGraw-Hill, 1968.
68
Suzuki, S., Maruyama, S., and Ido, H., Boundary Element Analysis of Cavity Noise Problems with Complicated Bound-
ary Conditions, Journal of Sound and Vibration, Vol. 130, No. 1, 1989, pp. 7991.
69
Marburg, S. and Nolte, B., Computational Acoustics of Noise Propagation in Fluids, chap. A Unied approach to FEM
and BEM in acoustics, Springer, 2007.
70
Dowell, E. H., Chao, C.-F., and Bliss, D. D., Absorption Material Mounted on a Moving WallFluid/Wall Boundary
Condition, Journal of the Acoustical Society of America, Vol. 70, No. 1, 1981, pp. 244245.
Appendix: generalized impedance boundary condition
In this appendix the problem of an impedance boundary condition for an acoustic velocity potential
wave model in the presence of grazing ow is reviewed with the intent of highlighting the main assumptions
behind the two boundary conditions employed for the FEM computations of sound transmission through an
acoustically treated nacelle intake discussed in subsection III.A. For the sake of generality, a vibrating surface
is considered in the following analysis, although only non vibrating surfaces were considered in present paper.
Historical note
During the Seventies, some discussion and controversy arose in the literature about the proper acoustic
boundary condition to be applied on a generic vibrating surface in the presence of a grazing ow. The
rst formulation for an untreated vibrating surface was rst suggested by Taylor,
50
and it was successively
derived in a formal way by Myers.
42
In the same paper, Myers also extended the formulation to the case of
an impedance steady surface. Myers Impedance Condition (MIC) can be seen as an extension of the well
known Ingards condition
51
that is valid for a at plate in a uniform stream. The MIC has been widely
accepted and used in a large number of numerical and analytical works.
52, 53, 54, 55, 56, 33, 57, 34
In the last years
some problems arose, in particular with time-domain numerical solvers using the MIC, that evidenced how,
in the presence of impedance, not only the numerical solution diverges due to a numerical instability, but also
the boundary condition itself is ill-posed and intrinsically unstable.
58, 59, 60, 61
These results, however, seem to
be in contrast with the experiments, since evidence of this instability has been reported only rarely. Several
eorts have been therefore undertaken in order to eliminate or mitigate the instability, and the attention has
been recently focused in the direction of adding more physics to the phenomenon, thus deriving modied
boundary conditions that better reproduce the sound propagation in a very thin layer close to the surface.
62, 63
These attempts have succeeded in showing that in some case the boundary condition can become well-posed
and stable.
In 2001 Auregan et al
64
and more recently Brambley
65
have argued that the viscous and turbulent eects
near the wall can alter the MIC. In particular, they have demonstrated that the continuity of the normal
particle displacement across the vortex sheet, which is used to model the eect of the boundary layer in an
otherwise inviscid medium, holds only when the thickness of the acoustic boundary layer is much smaller
than the thickness of the stationary boundary layer, i.e., typically in the high frequency range. At very low
frequencies, continuity of the acoustic velocity normal across the vortex sheet must be applied instead. Very
recently, Renou & Auregan
66
have conducted some measurements in a duct with the aim of proving the
failure of the MIC and validate a modied model in which a factor
v
, varying from 0 to 1 and reproducing
the eect of transfer of momentum into the lined wall induced by molecular and turbulent viscosities, allows
to blend the MIC (
v
=0) and a condition reproducing the continuity of the acoustic velocity normal to the
wall (
v
=1).
In the present Appendix the mathematical formalism introduced by Myers is used to derive a Generalized
Impedance Condition (GIC) for a vibrating surface. In a rst step the concept of acoustic impedance for
a vibrating surface is discussed, starting from the consolidated denition adopted in the vibro-acoustic
literature that expresses the surface impedance as the ratio between the pressure and the relative velocity
between uid and surface. In a second step, two forms of the GIC are derived, depending on the hypothesis
made to model the velocity of the vortex sheet. Finally, following Auregan et al ,
64
a blending formula is
proposed in order to incorporate the two formulations into a unique GIC.
Impedance denition for a vibrating surface
The impedance denition that we are considering in the present work is the so called normal surface
impedance or boundary impedance that, in the hypothesis of locally reacting material, can fully charac-
26 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

terize the absorbing properties of a liner placed over a solid surface. It is worthwhile to mention that this
impedance is just a macroscopic quantity that can be used to obtain a macroscopic description of the complex
uid/structure interactions taking place at microscopic level in the sound absorbing material. The situation
is illustrated in Fig. 34, where a porous liner is placed over a solid plate in the presence of a grazing ow
and an incident acoustic eld. It is evident that, at the microscopic level, the acoustic eld at the interface
between the liner and the uid greatly change locally, but once that one averages the pressure and the
velocity over a suciently large region some averaged quantities can be obtained that describe the eect of
the liner at a macroscopic level. In the following we will focus on these macroscopic uid quantities and we
will try to identify a proper denition of the surface impedance in terms of these macroscopic quantities for
the case of a liner placed on a vibrating surface.
Flow
Sound wave
U(y)
p
av
u
n av
p
u
n
Figure 34. Macroscopic versus microscopic ow behaviour at the interface between uid and a sound absorbing material
placed on a solid plate.
Looking at the literature we nd substantially two denitions of the boundary impedance for an acous-
tically treated surface. The rst one that we denote as Z
a
can be found in the acoustic and aeroacoustic
literature,
51, 67, 42
and denes the impedance as the ration between the acoustic pressure p and the acoustic
normal velocity u
n
:
Z
a
= p/ u
n
, (6)
where u
n
is the averaged acoustic velocity at the liner interface projected in the surface normal direction.
Throughout, uctuating quantities (uid pressure and velocity, surface velocity) are represented by their
Fourier transform counterpart denoted with the symbol. The second denition, denotes as Z
v
and typically
used in the vibro-acoustic domain,
68, 69
states that the impedance is dened as the ratio between the acoustic
pressure p and the relative normal velocity u
r
of the uid with respect to the surface:
Z
v
= p/ u
r
. (7)
The surface velocity projected in the surface normal direction is denoted as

V
n
, as sketched in Fig. 35. Notice
that in both the above impedance denitions the minus sign is due to the fact that we have used the typical
convention for Z according to which the velocity is positive when entering the liner, while the unit normal
vector points from the surface into the uid. Obviously Z
a
and Z
v
are coincident when the surface is at rest
V
n
u
n
Vibrating support surface
Vibrating
liner surface
n
Figure 35. Illustration of a liner placed on a vibrating surface; Vn is the velocity of the liner surface, while un is the
acoustic velocity at the uid/surface interface.
(

V
n
=0), but when the surface is vibrating the two expression are substantially dierent, and so we have to
try to identify the most appropriate denition to be used. In order to do that, it is useful to focus on the
expected properties for the impedance. It is clear that one important property would be that the impedance
27 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

of a given surface does not change with the surface motion. Let us look at Z
a
and Z
v
for an untreated
surface for which we expect to have Z . For an untreated vibrating surface we have u
r
=0 and so we
get Z
v
, but for Z
a
instead we get Z
a
and in addition Z
a
is a function of the surface velocity.
The denition in terms of the relative uid/structural velocity seems therefore preferable for our scopes.
This is further reinforced by considering other aspects, such as the physical mechanism of damping through
the liner orices that is driven by the relative velocity, and also the classical coupling of dynamical systems
where the relative velocity is used.
70
It can be therefore concluded that the more convenient denition for
the surface impedance is based on the relative velocity between uid and surface and we will therefore adopt
this denition in the following.
Expression of the relative velocity normal to a vibrating surface
In this section the expression of the relative velocity u
r
between uid and a vibrating surface is derived. The
expressions is valid for a generic material surface, both solid or uid (e.g. a vortex sheet).
Let us consider a generic curved surface dened by the equation f(x, t) =0. The unit normal vector to
the surface is given by:
n = f/|f|, (8)
while the surface velocity V is such that df/ dt =0, i.e.:
f
t
+V f =
f
t
+|f| V n = 0, (9)
from which the normal surface velocity V
n
can be derived, i.e.:
V
n
=
1
|f|
f
t
. (10)
Following Myers,
42
the surface f =0 can be modeled as a time-averaged shape (x)=0 with superimposed
the unsteady perturbationg(x, t), i.e. f =g, where the function g is of the same order as and 1.
Hence the moving surface is dened by the equation (x) = g(x, t). Considering the Fourier component
with angular frequency (+i t convention) of the surface motion and using Eq. 10, we obtain:

V
n
=
i g
||
+ O
_

2
_
. (11)
It is also useful to derive an expression for the unsteady unit normal vector, which reads:
n(x, t) =
g
||
+ O
_

2
_
=

||

g e
i t
||
+ O
_

2
_
; (12)
hence, making use of Eq. 11 and introducing the unit normal vector to the time-averaged surface n
0
, we can
write:
n(x, t) = n
0
(x)

V
n0
(x)
i
e
i t
+ O
_

2
_
on = g. (13)
Let us now suppose that the unsteady velocity eld can be decomposed into a time-averaged eld U(x)
and a superimposed acoustic perturbation u(x, t), such that |u| |U|. Then, let us introduce the relative
uid/surface velocity projected on the normal direction and dened on the deformed surface = g, which
reads:
u
r
(x, t) = [u(x, t) +U(x) V(x, t)] n(x, t)
=
_
u(x) e
i t
+U(x)

V(x) e
i t
_

_
n
0
(x)

V
n0
(x)
i
e
i t
+ O
_

2
_
_
(14)
= u(x) n
0
(x) e
i t
+U(x) n
0
(x) U(x)

V
n0
(x)
i
e
i t


V(x) n
0
(x) e
i t
+ O
_

2
_
.
By applying the perturbation analysis adopted by Myers,
42
the relative velocity u
r
on the moving surface
can be computed through a Taylor expansion applied to the time-averaged surface, i.e.:
u
r
(x, t) |
=g
= u
r
(x, t) |
=0
+ g
u
r
(x, t)

|
=0
+ O
_

2
_
. (15)
28 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

Hence, making use of Eq. 14, we can write:
u
r
(x, t) |
=0
= u(x) n
0
(x) e
i t
+U
0
(x) n
0
(x) U
0
(x)

V
n0
(x)
i
e
i t


V(x) n
0
(x) e
i t
+O
_

2
_
, (16)
and:
g
u
r

|
=0
=
g
||
u
r
n
|
=0
=
g
||
(U
0
n
0
) n
0
+ O
_

2
_
(17)
=

V
n0
i
(U
0
n
0
) n
0
e
i t
+ O
_

2
_
,
where use of Eq. 11 has been made, only the zero-th order term in Eq. 16 has been retained (u
r
(x, t) |
=0
=
U
0
(x) n
0
(x) + O()), and the hypothesis of smoothly curved surface has been introduced, i.e.:
u
r
= [U
0
(x) n
0
(x)] + O() [U
0
(x)] n
0
(x) + O() . (18)
Finally, substituting Eqs. 16 and 17 into Eq. 15, and making use of the slip condition U
0
n
0
=0, we obtain
the expression of the harmonic component of the relative velocity on the deformed surface:
u
r
= u
n0
U
0

V
n0
i
+

V
n0
i
(U
0
n
0
) n
0


V
n0
+ O
_

2
_
(19)
Impedance boundary condition for a vibrating surface
In this subsection, a GIC for a vibrating surface is derived, with reference to the conceptual scheme depicted
in Fig. 36. The eect of the boundary layer is modeled through a vortex sheet that supports an idealized
discontinuous time-averaged ow that below the vortex sheet is forced to be at rest (U=0). The liner and
vortex-sheet normal velocities are indicated as

V
n
and

V
I
n
, respectively.
l
i
n
e
r

s
u
r
f
a
c
e
v
o
r
t
e
x
-
s
h
e
e
t V
I
n
V
n
+
-
U=0
U 0
a(x)=0
a
I
(x)=0
n
I
0
n
0
Figure 36. Conceptual scheme for a moving surface in the presence of a discontinuous time-averaged velocity eld U
and a uctuating vortex sheet.
The derivation proceeds with the denition of the relative normal velocity in the dierent layers of the
domain indicated in Fig. 36. Immediately above the liner surface, Eq. 19 yields:
u
w
r
= u
w
n0


V
n0
. (20)
Hence, making use of the surface impedance denition, as given in Eq. 7, we obtain:
p
w
Z
= u

n0
+

V
n0
. (21)
Immediately below and above the vortex sheet, Eq. 19 yields:
u

r
= u

n0


V
I
n0
(22)
u
+
r
= u
+
n0


V
I
n0
U
0

V
I
n0
i
+

V
I
n0
i
_
U
0
n
I
0
_
n
I
0
, (23)
where U
0
is the time-averaged velocity immediately above the vortex sheet. Since the relative velocity u
r
is
continuous across the vortex sheet, by subtraction of Eqs. 22 and 23 we obtain:
u
+
n0
= u

n0
+U
0

V
I
n0
i

V
I
n0
i
_
U
0
n
I
0
_
n
I
0
. (24)
29 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

Finally, making use of the hypothesis of continuity of pressure across the vortex sheet ( p
w
= p

= p
+
), and
by setting, for simplicity of notation, p
+
p and u
+
n0
u
n0
, the following condition relating the inviscid
acoustic normal velocity, the acoustic wall pressure, the liner normal velocity and the vortex-sheet normal
velocity can be derived:
u
n0
=
p
Z
+

V
n0
+U
0

V
I
n0
i

V
I
n0
i
_
U
0
n
I
0
_
n
I
0
. (25)
This equation models the eect on the acoustic propagation due to an impedance vibrating surface in the
presence of ow and a vortex sheet.
Eq. 25 cannot be used unless the normal velocity of the vortex sheet is not dened. Two dierent scenarios
can be dened. The rst one is based on the following hypotheses:
the unsteady ow displacement across the vortex sheet due to the viscous interaction between the
acoustic uctuation and the liner surface does not aect the vortex-sheet velocity or, equivalently, the
acoustic boundary layer is much thinner than the steady boundary layer, i.e.
a

_
/f ;
the vibration of the liner surface does not aect the vortex-sheet velocity or, equivalently, the normal
liner displacement is much smaller than the steady boundary layer, i.e. V
n
/f .
As a consequence, in the rst scenario, the vortex sheet is driven by the impinging acoustic wave, i.e.

V
I
n
= u

n0
, hence from Eq. 21 we have

V
I
n
= p/Z +

V
n0
, and the particle displacement is continuous across
the vortex sheet.
The second scenario results from the failure of one of the two hypotheses above: if the acoustic boundary
layer is larger than the steady boundary layer, then the acoustic velocity u
n
is continuous across the vortex
sheet or, equivalently, the vortex sheet has a vanishing eect; if the normal liner displacement is larger than
the steady boundary layer, then the inviscid velocity uctuation u
+
n
is directly aected by the liner velocity.
In both cases, the eect is mathematically equivalent to assume that the vortex sheet is rigidly driven by
the liner surface, i.e.

V
I
n
=

V
n
.
The boundary conditions associated with the depicted scenarios are:
u
n0
=
p
Z
+

V
n0
U
0

_
p/Z

V
n0
_
i
+
p/Z

V
n0
i
(U
0
n
0
) n
0
, (26)
u
n0
=
p
Z
+

V
n0
+U
0

V
n0
i

V
n0
i
(U
0
n
0
) n
0
. (27)
Since the above expressions represent two opposite conditions, it comes natural to dene a unique condition
by introducing a blending factor , thus writing:
u
n0
=
p
Z
+

V
n0
U
0

_
(1 ) p/Z

V
n0
_
i
+
(1 ) p/Z

V
n0
i
(U
0
n
0
) n
0
. (28)
The condition = 0 is equivalent to Eq. 26, whereas the condition = 1 is equivalent to Eq. 27. It is
worthwhile to notice that Eq. 27 can be obtained directly by inserting the normal relative velocity, as
dened in Eq. 23, into the impedance denition given in Eq. 7, and by letting the vortex sheet coincide with
the liner surface. In other words, the condition =1 is equivalent to neglect the inuence of the vortex sheet
in the present ow idealization. In the case of a non vibrating surface, Eq. 28 reduces to:
u
n
=
p
Z
U
[(1 ) p/Z]
i
+
(1 ) p/Z
i
(U n) n. (29)
30 of 30
American Institute of Aeronautics and Astronautics
D
o
w
n
l
o
a
d
e
d

b
y

E
h
a
b

F
a
r
e
s

o
n

J
u
n
e

2
4
,

2
0
1
4

|

h
t
t
p
:
/
/
a
r
c
.
a
i
a
a
.
o
r
g

|

D
O
I
:

1
0
.
2
5
1
4
/
6
.
2
0
1
4
-
3
1
0
1

Vous aimerez peut-être aussi