Vous êtes sur la page 1sur 12

48th AIAA Aerospace Sciences Meeting Including the New Horizons Forum and Aerospace Exposition

4 - 7 January 2010, Orlando, Florida

AIAA 2010-655

Towards Development of an Active Single-Layer Acoustic


Liner for Jet Engine Noise Reduction
Michael Perrino1, Jeff Kastner 2 Ephraim Gutmark 3
University of Cincinnati, Cincinnati, OH, 45221
Sivaram Gogineni4
Spectral Energies, LLC, Dayton, OH 45431

Experiments and a Lumped Element Model are used to develop an acoustic liner with an increased degree of
freedom by adding a compliant face sheet to the system. The model is setup to include the neck and cavity
geometry of the liner and a compliant face sheet. The model is then used to predict the impedance of the liner
due to the Helmholtz resonator and the compliant face sheet. A set of experiments were run in a wave tube
apparatus to measure the impedance of various acoustic liners. The parameters studied included varying the
diameter of the hole in the face sheet, the cavity depth, and face sheet thickness. The model and experimental
results were compared, and it is shown that the model is good at predicting the absorption properties for the
various geometries. Depending on the coupling between the liner geometry and the compliant face sheet, the
absorption characteristics were altered. The bandwidth of noise suppression (absorption coefficient > 0.8)
was 500 Hz for a 0.2 diameter hole, a 1 cavity depth, and a material thickness of 0.012.

I. Introduction

he importance of noise suppression was magnified when noise regulations were established by the United
Nations International Civil Aviation Organization (ICAO) Annex 16 and by the United States governments
Federal Aviation Agency (FAA) Part 36 in the late 1960s 1. Even 50 years later, the high noise levels associated
with commercial and military aircraft continues to be a growing concern. The development of engines towards
higher thrust, the sprawl of communities towards once remote runways, and the closeness of ground crews to
aircrafts during take-off on aircraft carriers have all forced serious attention to be focused on addressing the noise
pollution issue. The noise also imparts vibrations on the aircraft structure which can be damaging to the aircraft and
significantly increases the noise levels in the cockpit. The engine noise has two main components: (1) internal noise
and (2) jet noise. The jet noise is due to the high-speed jet exhausting out the rear of the engine. The internal noise
is associated with rotating components for the turbines and the combustion noise.
Acoustic liners are the current technology for attenuating combustion and turbine noise radiated from the jet engines
of aircrafts. Figure 1 illustrates a commercial turbofan engine with acoustic liners. The liners form the flow paths in
the nacelles and are used to suppress the engine noise. Acoustic liners (see Figure 2) typically have a single layer
(Single Degree of Freedom, SDOF) or a double layer design (Two Degree of Freedom, 2DOF) 2,3,4,5. The major
drawback of current acoustic liner technology is that they are passive by nature. A single layer passive acoustic liner
provides excellent absorption characteristics for a very small bandwidth, and as a result, is only effective for specific
flight regimes. Multi-layer acoustic liners can be designed with a wider effective frequency range but they absorb
less sound at a given frequency than a single layer liner over their effective frequency range. An actively controlled
single layer acoustic liner would provide the capability of changing its acoustic characteristics during flight to
provide a wider effective frequency range than a passive single layer acoustic liners without sacrificing the liner
effectiveness.

Ph D. Student, Department of Aerospace Engineering, AIAA Student Member


Research Professor, Department of Aerospace Engineering, AIAA Member
3
Professor, Ohio regents Eminent Scholar, Department of Aerospace Engineering, AIAA Fellow
4
President, Spectral Energies, LLC, AIAA Fellow
1
American Institute of Aeronautics and Astronautics
2

Copyright 2010 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.

Figure 1: Picture of acoustic liners on a turbofan engine


There is also the bulk absorber liner which uses porous bulk materials instead of the honeycomb structures as shown
in Figure 2c 6,7,8,9,10,11,12. These bulk absorber liners are not currently used on aircraft engines due to the increased
weight of the absorber from the undesired absorption of water or fuel. The analytical impedance models and general
formulae of the three different types of acoustic liners (SDOF, 2DOF and bulk absorber liner) are well established as
functions of liner geometries and described by Zorumski and Motsinger & Kraft 13,14,15.

Figure 2: Configurations of three types of acoustic liners


The main focus of this research is to build and test an active acoustic liner with the intention of ultimately building a
smart acoustic system that detects the noise frequency and controls the liner accordingly. The feasibility of the
approach is demonstrated by computational means through modeling the acoustic liner as a Helmholtz Resonator
lumped-acoustic system using the electric circuit analogy similar to the approach used at the University of Florida
16,17,18
. The models are then modified to incorporate a compliant face sheet composed of an isotropic material with
an effective mass and compliance. Several different compliant face sheets are modeled and then tested
experimentally in an impedance tube. The results of these experiments will be used to validate the models. Future
work will include incorporating a piezoelectric element that vibrates due to an electrical excitation.

2
American Institute of Aeronautics and Astronautics

II. Basic Helmholtz Resonator Model


The acoustic liner model begins with a simplified system and then is advanced by building it up to the more
complicated liner system 19. A good simple model for an acoustic liner is the Helmholtz resonator. A schematic of
a Helmholtz resonator is shown in Figure 3a, which includes an external oscillating pressure (Pext()). The job of an
acoustic liner is to suppress this pressure in the engine nacelle. The Helmholtz resonator has two key components
(1) the neck which is represented by the subscript n and (2) the cavity which is represented by the subscript c. This
model assumes that the walls of the liner are rigid and the shortest acoustic wavelength of interest is much larger
than any of the physical dimensions of the resonator. The neck of the resonator works like an airtight piston with a
mass m and length L. This mass is calculated using Equation 1:

mn An L

Eq. (1).

When the external pressure of the cavity is larger than the cavity pressure, the piston is forced into the cavity by a
distance . When the air in the cavity is compressed, the cavity acts like a spring of stiffness s given by Equation 2:

An2
sc c
Vc
2

Eq. (2).

There is also a resistance to the movement of the air through the neck that dampens the movement of the air tight
piston. This resistance Rn is primarily due to skin friction and vortex generation caused by the air flowing from the
neck into the open volume. This resistance is highly dependent on its geometry and the manufacturing process used.
For the current work, the flow resistance will be measured experimentally using a flow resistance apparatus 20.
The equation of motion governing the Helmholtz resonator is a second-order ordinary differential equation:

mn

d 2
d

R
sc SPe jt
n
2
dt
dt

Eq. (3).

The Helmholtz resonator in Figure 3a can be converted into a Lumped Element Model, LEM, as shown in Figure
3b 10,11

(a)

(b)

Figure 3. (a) Schematic of a Helmholtz Resonator and (b) Lumped Element Model of a
Helmholtz Resonator.

3
American Institute of Aeronautics and Astronautics

The LEM represents the external acoustic wave as an A.C. power source P, the neck of the Helmholtz resonator as a
resistance R and inductance L, and the cavity as compliance C. The lumped capacitance and inductance elements are
related to the resonator parameters through Equations 4 (a) to (c):

m
L n2
An

An2
C
sc

Eq. (4a)

Eq. (4b)

Rn
An2

Eq. (4c).

The RLC circuit presented in Figure 3b is a standard representation of a harmonic oscillator. To determine the
acoustic impedance, resonant frequencies, and damping coefficients for this system, it is necessary to perform a
circuit analysis that can be used to predict the impedance of the Helmholtz resonator. The total impedance of the
model can be calculated using Equation 5:

Z RLC

1
L R
C

Eq. (5).

III. Resistance Measurement Setup


The resistance of the face sheet is measured using a flow resistance apparatus at the University of Cincinnati. This
apparatus measures the steady state resistance of a perforated sample for a constant flow velocity. A diagram of the
flow resistance apparatus is shown in Figure 4.

Figure 4: Schematic diagram of a flow resistance apparatus

The two main quantities being measured to calculate the resistance of a porous sheet are the average particle
velocity through the test sample and the static pressure drop across it. The average particle velocity through the neck
hole corrected to standard conditions is given by Equation 6 and the corrected resistance of the neck hole is given by
Equation 7 as described by Motsinger et al 19:

Q P
Un f
An Po

To

Tf

2.5

Eq. (6)

4
American Institute of Aeronautics and Astronautics

Rn Ps

An Ps

Q Pf

Eq. (7).

The standard conditions of pressure and temperature are defined as 14.7psi and 70 F respectively. The flow rate Q is
measured using a Meriam Flow Meter. The three differential pressure measurements are taken using Mensor 6115
pressure transducers. The atmospheric pressure and temperature are measured using a Fisher Scientific digital
barometer. The resistance of the hole is measured for several different particle velocities for flow going both
directions through the sample. A linear fit is applied to the data.

IV. Experimental Impedance Measurement Setup


The results from the computational model can be compared to experiments performed using a wave tube apparatus
at the University of Cincinnati. Figure 5 is a schematic of this wave tube apparatus. The wave tube has a 1.5 x 1.5
inch square cross section and is 25 inches long. An acoustic compression driver (JBL 2446H) is attached at one end
as the sound source, and a sample is placed at the other end with a backing cavity. A signal generator (HP 8904A
Multifunction Synthesizer) is used to generate voltage signals for the acoustic driver. The generator is capable of
generating broadband signals as well as tonal signals at a desired frequency and phase shift.

Liner Face Sheet

Figure 5: Schematic diagram of a wave tube apparatus

The experimental data comes from four acoustic transducers (Endevco model 8506-2) attached to the wall of the
wave tube. The transducers measure the acoustic pressure in the wave tube. The acoustic field in the wave tube is
assumed to be uniform in the transverse direction. This assumption is valid for the plane wave mode, which is below
the first cut-on frequency. For the wave tube currently used, the cut-on frequency of the first order mode is 4500 Hz.
Measurements in the wave tube lose accuracy above this frequency. The measured signals are amplified by signal
conditioners and converted from analog to digital for computer input. The signal processing and analysis are
performed by custom LabView programs. The programs compute the acoustic impedance and the absorption
coefficient.
The impedance of the liner sample is defined as the ratio of acoustic pressure (p) to particle velocity (u). Thus, the
impedance on the surface of the sample is written as

Z( f )x L

p( L, f )
A exp( i k L ) B exp( i k L )

cu( L, f )
A exp( i k L ) B exp( i k L )

(Eq. 8)

where L is the length of the wave tube L = 25 inch and A and B are defined in Figure 5. This allows the impedance
from the computational model (see Eq. 5) can be compared to experimental results (Eq. 8).
5
American Institute of Aeronautics and Astronautics

The overall goal of the current work is to develop an acoustic liner with a compliant face sheet that has improved
absorption properties compared to a passive acoustic liner. The absorption coefficient of a liner is defined as the
fraction of incident acoustic energy absorbed by the test sample and is mathematically defined as:

A B
2

AbsorptionCoefficient

Eq. (9).

Total absorption and no absorption represent the two extremes for the absorption properties of an acoustic liner. If
the acoustic liner absorbs all the acoustic energy, the absorption is equal to one because B = 0. If the liner absorbs
none of the acoustic energy, then A = B and absorption equals zero. Absorption provides a nice metric for
comparing different liners.

V. Results: Helmholtz Resonator


Figure 6 shows the amplitude of the total impedance, the phase of the total impedance and the absorption for the
Helmholtz resonator as predicted by the model (red line) and from experimental results (blue line). The geometric
parameters for this resonator are a neck diameter (Dn) of 0.24, a neck thickness (Ln) of 0.4, and a cavity volume
(Vc) of 2.25 in3. Both the model and the experimental results show the liner having its largest absorption coefficient
around 600Hz. The biggest difference between the experiment and the model results is at the low frequencies. This
is primarily a result of the acoustic drivers inability to generate sufficient noise levels at lower frequencies.

(a)

(b)

(c)

Figure 6. Comparison between model and experiment for a Helmholtz resonator with a 1 cavity. (a)
Total Impedance amplitude, (b) total impedance phase, and (c) absorption coefficient.

The agreement between experiment and model is further validated by changing the depth of the Helmholtz cavity.
Changing the cavity depth (Lc) primarily influences the resonance frequency, but the damping also slightly changes.
Figure 7 compares the absorption coefficient from the model and experiment for Lc = 0.5, 1.0 and 1.8. The
model accurately captures the variation in frequency and breadth of the high amplitude region.

6
American Institute of Aeronautics and Astronautics

(b)

(a)

(c)

Figure 7. Absorption coefficient for a Helmholtz resonator with various cavity depths: (a) L c = 0.5, (b)
Lc = 1.0, and (c) Lc = 1.8

VI. Results: Compliant Face Sheet


The next step towards developing an active acoustic liner requires the simple Helmholtz resonator presented in
Figure 3 to be modified to include a compliant face sheet. Figure 8 shows that by connecting the face sheet
elements in parallel with both the neck and the cavity of the Helmholtz resonator, the basic circuit is modified to
account for the compliant face sheet. This model is similar to, but not exactly the same, to that used by Horowitz et
al. for a compliant back wall on a Helmholtz resonator 11.
The inductance and compliance of the compliant face sheet can be modeled as M p

9L
5rp2

and C p

rp6 1 2
3

16 EL

respectively. L is the thickness, r is the diameter, is the material density, E is the elastic modulus of the material,
and is Poissons ratio of the compliant face sheet.7 To match the computational model to the experiment it is also
necessary to include a Resistance (Rp) which will be found by: R p 2 / C p / M p . The only unknown is the
damping coefficient (), however this is traditionally adjusted based on experimental results. For the present time,
we will also ignore the fact that a voltage source can be used to vibrate the face sheet future work will explore this
application. Instead we will focus on how Pext excites the compliant face sheet and changes the absorption
characteristics of the liner. The total impedance for the model in Figure 8 can be calculated using Equation 10:
Ztotal

1
1
1

j M p R p Z p j M n Rn Z c

Eq. (10).

Figure 9 presents the results for a compliant face sheet for results from both the model and experiments. The model
results came from analyzing the circuit in Figure 8, and the experimental results came from running white noise
through the wave tube. The face sheet only had a thickness of 0.012 compared to the face sheet thickness of 0.40
for the results presented in Figures 6 and 7. The compliant material was stainless steel ( = 7850 kg/m3, E = 7x1010
N/m2, = 0.25). For the model, Simulink was used to determine the impedance across the capacitor for the cavity
(Zc) and piezoelectric element (Zp). The model nicely predicts the two peak frequencies, and it has agreement in
terms of both phase and amplitude of the total impedance. The advantage to using a compliant face sheet is
demonstrated here by increasing the degrees of freedom for the liner while not requiring any increased space.

7
American Institute of Aeronautics and Astronautics

Figure 8. Lumped Element Model for Compliant Face Sheet in Parallel with neck and cavity.

(a)

(b)

(c)

Figure 9. Comparison between experiments and model for (a) amplitude of total impedance, (b) phase of
total impedance, and (c) absorption of liner.

To further demonstrate the increased absorption properties of a compliant face sheet liner and to further support the
validity of the model, the diameter of the hole in the compliant face sheet, the length of the cavity, and the thickness
of the compliant material will all be changed. Figure 10 shows the absorption coefficient for holes with a diameter
of 0.17, 0.24 and 0.35, respectively. The results in Figure 9 were for a face sheet with a hole of 0.24 diameter.
The change in diameter is representative of changing the porosity of a face sheet. The coupling between the two
peak amplitudes in the absorption profile is dependent on the diameter of the hole. For a diameter of 0.17, the
Helmholtz frequency is around 700 Hz and the compliant face sheet is around 1300 Hz. As the diameter increases,
two things happen: (1) the Helmholtz frequency begins to rise and the tone due to the compliant face sheet begins to
decrease. By the time the diameter is increased to 0.35, the two tones are very close in frequency. This is seen in
both the model and the experimental results. Such a case provides a broad range of frequency suppression (~ 500
Hz) where the absorption coefficient is above 0.8.
The results for the liner sample in Figure 10c were further expanded to include cavity depths of 0.5, 1, and 1.8.
Figure 11 shows the absorption profile for the various cavity depths with a 0.34 diameter hole. As the cavity depth
is changed, the coupling between the Helmholtz resonator and the compliant face sheet is varying. For the shortest
cavity depth (Lc = 0.5), the Helmholtz frequency is greater than the resonance frequency for the compliant sheet.
This case does not look very desirable for broadband noise suppression. On the other hand, the larger cavity depths
of 1.0 and 1.8 demonstrate how increasing the cavity depth leads to a drop in the Helmholtz frequency. At the
same time, the compliant sheet parameters do not changing so its peak absorption remains around 1150 Hz. This
demonstrates the ability of this model to aid in the design of future experiments and future liners with compliant face
sheets.

8
American Institute of Aeronautics and Astronautics

(a)

(c)

(b)

Figure 10. Absorption coefficient for a compliant face sheet with various diameter holes. (a) D n = 0.17,
(b) Dn = 0.24, and (c) Dn = 0.305

(a)

(c)

(b)

Figure 11. Absorption coefficient for cavity depths (Lc) of (a) 0.5, (b) 1.0, and (c) 1.8.
The model was also validated for various face sheet thicknesses. This brings out two importance facets. First, the
ability of the model to fully represent the face sheet, and second, the flexibility in the face sheet design. Figure 12
shows the absorption coefficient for a thick face sheet which is not compliant, the thin face sheet that was also to
create Figures 9 to 11, and a very thin face sheet. The very thin face sheet had a thickness of 0.003. All cases had
a cavity depth of 1 and the diameter of the hole in the neck was 0.24. The non-compliant face sheet in Figure 12a
only has one resonance. The compliant face sheets shown in Figure 12b and 12c have similar characteristics in that
they both have two local peak amplitudes. The biggest difference is the amplitude of the two peak values. This is
most likely due to the tuning between the cavity depth and the face sheet thickness.

(a) Noncompliant Sheet

(b) Thin Sheet

(c) Very Thin Sheet

Figure 12. Absorption coefficient for Helmholtz resonators with different face sheets of different
thicknesses.
9
American Institute of Aeronautics and Astronautics

It is well known the Helmoholtz resonator is an over simplification for an acoustic liner. A multi hole geometry was
studied next to bring both the model and experiment results closer in line with a acoustic liner. Figure 13a shows a
schematic of a Helmholtz resonator with three holes. The schematic shows all the holes to be the same size, but by
no means would this be a requirement. Figure 13b shows the LEM for a multi-hole Helmholtz resonator. The model
accommodates the multi-hole geometry by placing all the holes in parallel. Simulink is then used with this geometry
to show how the impedance of the cavity changes due to the modified geometry. Figure 14 shows experimental
results for a compliant face sheet with one hole and with 3 holes. The model has already been shown to match the
single holes case in Figure 11c where the cavity depth was 1.8 and the hole was 0.3 in diameter. The 3 holes were
set such that they had a similar effective area as the single hole, which is why the peak amplitudes have similar
frequencies. On the other hand, by adding multiple holes the absorption profile significantly improves. This seems
to support the idea that as the model and experiments moves from single hole geometries to acoustic liner
geometries, significant improvement will be seen in the absorption properties.

Pext()
Ln

Area: An

Volume:Vc
Pc()

(a)

(b)

Figure 13. Helmholtz resonator with three necks (a) schematic and (b) LEM

Figure 14. Specific acoustic impedance for a Helmholtz resonator with a varying number of holes in
the face sheet. (a) Normalized amplitude and (b) Phase.

10
American Institute of Aeronautics and Astronautics

VII. Results: Resistance Measurements

Face Sheet Resistance (c.g.s. Rayls)

The flow resistance of the face sheet hole was measures independently of all other parameters using the flow
resistance apparatus. The resistances are measured for a range of flow velocities and the results are shown in Figure
15. The resistance value to be used in the lumped acoustic model will be dependent on the amplitude of the sound
waves impinging on the liner face sheet.

250
Dn=0.24in, Ln=0.003in
Dn=0.17in, Ln=0.013in

200

Dn=0.24in, Ln=0.013in
Dn=0.35in, Ln=0.013in

150

100

50

0
0

100

200

300

400

500

600

Average Flow Speed through the Face Sheet (cm/s)


Figure 15. Flow resistance of various compliant face sheets.

VIII. Conclusions
Experiments and computational modeling have been used to develop an acoustic liner with increased degrees of
freedom. The model is set up to include the neck and cavity geometry of the liner and the compliant face sheet.
MATLABs Simulink Toolbox was used to analyze various models through impedance measurements. Such an
analysis has given us the freedom to quickly vary the model and find how the changes may influence an engine
nacelle liner. The model predicts the resonant frequencies and the interaction between the Helmholtz resonator and
the compliant face sheet. A set of experiments were run in a wave tube apparatus. The parameters studied included
varying the diameter of the hole in the face sheet, the cavity depth, and face sheet thickness. The model was then
compared to experimental results and was shown to be very good at predicting the absorption properties for various
liner geometries. Depending on the relationship between the liner geometry and the compliant face sheet, the
absorption characteristics were altered. The bandwidth of noise suppression (absorption coefficient > 0.8) was 500
Hz for a 0.24 diameter hole, a 1 cavity depth, and a material thickness of 0.013. The resistances of the face sheets
were measured and will be implemented in the lumped acoustic model. Future work will continue to utilize the
model and experiments for optimum noise suppression. We also plan to further increase the degrees of freedom by
using a piezoelectric element on the face sheet.

Acknowledgments
The work presented here was supported under a Navy SBIR Phase I (topic number N08-142) and the assistance
from the Technical Monitors, Amy Little and Jesus Suarez, is greatly appreciated. We would also like to thank Dr.
Asif Syed and Chris Porter for their assistance in the wave tube experiments.
11
American Institute of Aeronautics and Astronautics

References
1) Smith, M.J.T., Aircraft noise, Cambridge University Press, 1989
2) Dunn, I.P. and Davern, W.A., Calculation of acoustic impedance of multi-layer absorbers, Appplied Acoustics,
19, pp321-334, 1986
3) Lee, F.C. and Chen, W.H., Acoustic transmission analysis of multi-layer absorbers, J. of Sound and Vibration,
(2001) 248 (4) pp.621-634
4) Bielak et al. Advanced Nacelle Acoustic Lining Concepts Development, NASA CR-2002-211672, August 2002.
5) Hillereau, N., Syed, A.A., and Gutmark, E.J., Measurements of the acoustic attenuation by single layer acoustic
liners constructed with simulated porous honeycomb cores, J. Sound & Vib., V 286, 2005, pp 21-36.
6) Delany, M.E. and Bazley, E.N., Acoustical properties of fibrous absorbent materials, Applied Acoustics 3 (1970)
105-116.
7) Allard, J.F., and Champoux, Y., New empirical equations for sound propagation in rigid frame fibrous materials,
J. Acoust. Soc. Am 1991 (6) pp. 3346-3353
8) Hersh, A.S. and Walker, B., Acoustic behavior of a fibrous bulk material, AIAA-79-0599
9) Kirby, R. and Cummings, A., Prediction of the bulk acoustic properties of fibrous materials at low frequencies,
Applied Acoustics 56 (1999) 101-125.
10) Lee, I., Selamet, A., and Huff, N., Acoustic impedance of perforations in contact with fibrous material, J.
Acoust. Soc. Am 119 (5) May 2006.
11) Voronia, N., Acoustic properties of fibrous materials, Applied Acoustics 42 (1994) pp.165-174
12) Voronia, N., Improved empirical model of sound propagation through a fibrous materials, Applied Acoustics 48
(1996) pp.121-132
13) Zorumski, W.E., Acoustic impedance of curved multilayered duct liners, NASA TN D-7277, 1973
14) Zorumski, W.E., Tester, B.J. Prediction of the acoustic impedance of duct liners, NASA TM X-73951, 1976
15) Motsinger, A. E. and Kraft, R. E. Design and Performance of Duct Acoustic Treatment, Chapter 14 in
Aeroacoustics of Flight Vehicles: Theory and Practice, Volume 2: Noise Control, NASA RP 1258, Vol. 2,
1991
16) Horowitz, S., Nishida, T., Cattafesta, L.N., Sheplak, (2001) Compliant-Backplate Helmholtz Resonators for
Active Noise Control Applications, AIAA Paper No. 2001-0817.
17) Horowitz, S., Nishida, T., Cattafesta, L.N., Sheplak, (2002) Characterization of Compliant-Backplate
Helmholtz Resonators for an Electromechanical Acoustic Liner, AIAA Paper No. 2002-0666.
18) Horowitz, S.B., Nishida, T., Cqattafesta III, L.N., and Sheplak, M., Characterization of a compliant-backplate
Helmholtz resonator for an electromechanical acoustic liner, Int. J. of Aeroacoustics, V 1, N 2, 2002, pp 283205.
19) Kinsler, L.E., Frey, A.R., Coppens, A.B., Sanders, J.V. , Fundamentals of Acoustics Fourth Edition, John Wiley
& Sons, Inc. 2000
20) Motsinger, R. E., Syed, A. A., and Manley, M. B., The Measurement of the Steady Flow Resistance of Porous
Materials, AIAA-83-0779, 8th Aero-Acoustics Conference in Atlanta Georgia, April 1983

12
American Institute of Aeronautics and Astronautics

Vous aimerez peut-être aussi