You are on page 1of 10

AIAA 2016-2848

Aeroacoustics Conferences
30 May - 1 June, 2016, Lyon, France
22nd AIAA/CEAS Aeroacoustics Conference

Unsteady Aerodynamics of High Speed Train Pantograph


Cavity Flow Constrol for Noise Reduction

Hogun Kim1, Zhiwei Hu2 and David Thompson3


University of Southampton, Southampton, SO17 1BJ, United Kingdom

Flow induced noise increase signigicantly with speed, consequently noise reduction has
become an important consideration for high-speed train designs. However, reducing noise
from a train pantograph is more challenging because of the difficulty to shield it using railside
acoustic barriers. The flow behavior around a high-speed pantograph arm and train roof at a
1/10 scale is investigated using computational fluid dynamics. The geometries of the roof and
the pantograph arm are simplified as a square shallow cavity and a straight cylinder,
Downloaded by PURDUE UNIVERSITY on June 6, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-2848

respectively. To resolve the details of the turbulent flow structures and hence enable accurate
noise predictions, the improved delayed detached-eddy simulation is used in near-field
modelling and Ffowcs-Williams & Hawkings aeroacousitc model is employed for far-field
acoustic calculation. In this work, the influence of geometrical alteration at the cavity leading
edge is considered. The results show that the recirculation vortices, which generate highly
turbulent flow, decrease with increasing cavity leading edge roundness, and this reduces wall
pressure fluctuations, which is a major noise source, on the cavity and cylinder surfaces.
Furthermore, aerodynamics performance, such as drag, is improved. Finally, the effect of the
arm positions is studied. No significant effect was found for overall sound pressure level, but
tonal noise levels from the cylinder was reduced. A reduction of interaction between the lower
part of the cylinder surface and cavity shear layers was detected, and it is believed that the
reduction of interaction reduces the peak noise levels.

Nomenclature
c = Sound of speed
Cd = Coefficient of drag
ij = Unit tensor
R = Diameter of cylinder
D = Depth of the cavity
h = Height of the surface on leading edge of cavity
h/D = Ratio of expansion surface of the cavity leading edge
L = Length of the cavity
L/D = Length-to-depth ratio
LE = Leading edge of cavity
m = Mass
Ma = Mach number
n = Surface normal vector
OSPL = Overall sound pressure level
p = Pressure
P = Cylinder position distance
Q = Second invariance of the velocity gradient
Q-norm = Normalized Q
r = Distance from a source point to the observer

1
Graduate student, Aerodynamics and Flight Mechanics Research Group, Faculty of Engineering and the Environment,
hk1g14@soton.ac.uk.
2 Lecturer, Aerodynamics and Flight Mechanics Research Group, Faculty of Engineering and the Environment, Z.Hu@soton.ac.uk

and AIAA member.


3 Professor, Institute of Sound and Vibration Research, Faculty of Engineering and the Environment, djt@isvr.soton.ac.uk.

1
American Institute of Aeronautics and Astronautics

Copyright 2016 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
= Density
SPL = Sound pressure level
t = Time
ij = Viscous stress tensor
ui, uj = Flow velocity
un = Normal velocity of the surface
vn = Normal velocity of the surface
W = Width of the cavity
W/D = Width-to-depth ratio
xi, xj = Flow position

I. Introduction

R ecently, high-speed rail transportation has evolved into a thriving industry. An efficient transport system is
necessary when attempting to compete with air and road transport. A high-speed train traveling at approximately
300 km/h is the most attractive form of transport1. However, in the context of high-speed trains, a pressing
Downloaded by PURDUE UNIVERSITY on June 6, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-2848

environmental challenge due to aerodynamic noise has long been identified. With high speed comes a high levels of
noise; therefore, the issue of noise reduction has become an essential consideration in high-speed train design. Because
of continuously increasing speeds, aerodynamic noise from the train is a major factor when train speeds reach around
260 km/h.2 Hence, noise reduction is critical to the future development, expansion, and operation of high-speed train
systems worldwide. The aerodynamic noise from high-speed trains is believed come from three main sources: the
bogies, intercoach gaps, and the pantograph.3 Noise from the bogies and the intercoach gaps can be eliminated by a
noise barrier; however, it is difficult to shield noise from the pantograph located on the top of the train. Therefore, a
number of research of pantograph noise have been studied. Grosche and Meier4 reported that the noise from
pantograph can be divided into three main sources: the panhead, arm joint and foot. Yu et al.5 have found that the
pantograph aerodynamics noise is generated by the unsteadiness of the wake and by flow separation. Furthermore,
they tested several pantograph noise insulation plates for noise reduction and the insulate cases might cause additional
cavity noise. To address pantograph noise issue, an efficient technique to control unsteady flow and reduce noise
emissions from pantographs is required; noise reduction can be achieved by reducing noise sources such as those
caused by turbulent flow interacting with the roof cavity and the pantograph arm. Therefore, improved understanding
of noise generation mechanism from the cavity flow is required, which can be achieved using a highly accurate
computational fluid dynamics (CFD) solver to resolve details of the turbulent structures and enable sufficient accurate
noise prediction. This paper employs the improved delayed detached-eddy simulation (IDDES) to simulate flow on a
1/10 scaled simplified model of a high-speed train roof cavity and pantograph components to investigate the unsteady
flow aerodynamics characteristics and noise sources in the near field. The Ffowcs-Williams & Hawkings (FWH)
acoustic analogy is involved to predict far-field noise. This study focuses on investigating the effect of geometrical
modifications of the cavity leading edge (LE) and cylinder (e.g., the pantograph arm) positions on noise.

II. Numerical Method


Since the incoming flow simulated in this study is at a low Mach number, the govering equations in flow field
used for the current study are the unsteady, incompressible continuity and Navier-Stoke equations
u j
0 (1)
x j

u i u p ij
u j i (2)
t x j x i x j

The nonconvective form of the FWH formulation is solved for far-field acoustic prediction by using the time-
resolved aerodynamics near-field data as input. The intergration surface is selected on the solid walls, and therefore
there is no monopole sources. The contributions from quadrupoles are neglected. Therefore, the FWH equation for
pressure can be represented by6:

2
American Institute of Aeronautics and Astronautics
1 F 1 F FMi
dS
2
r (1 M r ) ret
dS 2 2
r (1 M r ) ret
1 c
f 0 c f 0

p (3)
4 1 F ( rM r cM r cM r 2 )
c f 0 dS
r (1 M r )
2 3
ret

with F Pij ni u i (u n v n ) (4)

where p is acoustic pressure fluctuation and Mi is the Mach number in the xi direction and Mr is the Mach number in
the observer direction; F is inviscid flux and Pij is the compressive stress tensor, such that
Pij ( p p 0 ) ij ij (5)
Downloaded by PURDUE UNIVERSITY on June 6, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-2848

III. Numerical Model

A. Problem setup
Owing to the complexity of the actual pantograph geometry, it is necessary to simplify the geometry to reduce the
computational resource consumption. To obtain a similarity between the behaviour of flow around the train roof and
the pantograph supporting arm, a closed-cavity with L/D=15, W/D=6.5 (D is the cavity depth D=0.07 m) and a straight
cylinder of diameter R = 0.05 m are used, as shown in Figures 4.1 below. The approaching surface upstream of the
cavity leading edge (LE) is modified by rounding the sharp cavity LE surfaces of a fixed length (1D). The height (h)
of the rounded surface varies from 0.1D to 0.8D. The rounded surface shape is a circular arc of a constant radius (see
Figure 1).
The cylinder, representing the pantograph arm, is placed in the middle of cavity. To investigate influence of the
cylinder position, its position will be changed from 1D to 3D from the centre towards the trailing edge (TE) of the
cavity.

a) Definition of geometry and nomenclature b) Description of cylinder position study

Figure 1. Description of geometry.

Table 1. Summay of cases


h/D Reynolds nunumber Cells Computing prossors
0 0.2 0.4 0.6 0.8 3.4x10 5
Approx. 10 12 million 32

P Reynolds nunumber Cells Computing prossors


0 1D 2D 2.5D 3D 3.4x105 Approx. 12 million 32

3
American Institute of Aeronautics and Astronautics
B. Solver and numerical settings
The geometry is meshed using Star CCM+ embedded mesher. The snappy hex mesh (trimmed mesh) is applied to
all cases, as this kind of mesh has been shown to produce high performance for large domains used for complex
geometry simulations such as airplanes, ships, and trains7. To reduce computation cost and concentrate on resolving
flow details around the cavity and the cylinder, multirefinement regions are used (see Fig. 2). The regions around the
LE and TE of the cavity show a high pressure gradient and recirculation of the fluid, making it necessary to have fine
cells to capture accurate fluid behavior and the cell size in this region is the finest mesh in the domain. The inside of
the cavity region also needs to be refined because most fluid fluctuation occurs in the region. As fluid behaviour
around a cylinder, an excessive adverse pressure gradient and the amount of vorticity, the cell sizes of refinement
cylinder region and its wake region have been refined to be adequate for the flow. The computational domain size
should be large enough to avoid reflection from the external boundary, which may causes higher frequency. The
computational domain covers an area extending 14 D upstream and 18 D downstream of the cavity, with a 14 D height
and a 6.5 D width. For the boundary condition, the velocity inlet condition is defined as the upstream boundary with
Ma = 0.24 (corresponding to 83.3 m/s) and no-slip conditions are defined on the solid wall, including the cylinder and
the cavity. Two sides of the domain boundaries are defined as a periodic condition, whereas a symmetry condition is
imposed on the upper boundary. For the downstream boundary, a pressure outlet condition is applied. Furthermore,
Downloaded by PURDUE UNIVERSITY on June 6, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-2848

using the Spalart-Allmaras turbulence model near the wall region, the y+ values are approximately less than 1 for all
wall regions. All simulations in this paper were run with segregated flow and the implicit unsteady solver. Assuming
incompressible air in the project, the density of air is initialized as 1.225 kg/m3. For stable and accurate simulation,
the underrelaxation factors are 0.2 for pressure and 0.5 for velocity. Moreover, for temporal discretization, second
order is involved. To ensure a stable simulation, overall the Courant number is set as below 0.65.

a) Computational overall Mesh

b) Detail of the mesh in cavity inlet region c) Detail of the cylinder mesh

Figure 2. Overal mesh with refinement region.

IV. Numerical Result

A. Expansion surface cavity leading edge stduy

1. Aerodynamics results

To obtain an overview visualization of the unsteady flow around the cavity and cylinder, the iso-surfaces of the
second invariant of the velocity gradient, called the Q criteria, are plotted at a normalized value of 0.2, and these iso-
surface are colored by velocity magnitude, as shown in Fig. 3. It can be seen that vortices are generated by the cavity
LE, behind the cylinder arm, and cavity TE and most of the highly turbulent wakes are present in the cavity. The
vortices shed from the cavity LE impinge on the bottom surface of the cavity and the lower part of the cylinder arm.
It can be seen that the vortices on the cavity inlet begin to lessen as the expansion surface of the cavity LE (h/D)
4
American Institute of Aeronautics and Astronautics
increases. Hence, the impinged surface area on the cavity and cylinder by vortices is significantly decreased. However,
Fig. 4 c) indicates that the vortices slightly increase again; thus, it can be assumed that separation flow on the cavity
LE might generate the vortices.

Velocity magnitude, m/s


Downloaded by PURDUE UNIVERSITY on June 6, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-2848

a) h/D=0

b) h/D=0.6

c) h/D=0/8

Figure 3. Instantaneous normalized Q criterion (Q-norm = 0.2) colured by velocity magnitude for at 0.6s.

5
American Institute of Aeronautics and Astronautics
Figure 4 shows the contours of the mean velocity magnitude at the midplane of the z-direction (z=0.225 m). The
recirculation region around the LE of the cavity might generate a highly turbulent wake. Fig. 4 a) indicates the baseline
geometry, without any treatment to the LE of the cavity, h/D=0, and it shows the greatest recirculation region. The
impingement point is likely to be around the cylinder arm. It can thus be assumed that a shear layer can influence the
cylinder. Furthermore, it can be observed that as the h/D value increases, the recirculation region is significantly
reduced and the impingement point moves away from the cylinder. A reduction of the region tends to continue until
h/D=0.4, and it stops growing thereafter. The incoming flow does not follow the expansion surface wall, and the
behavior of the flow is likely to be separation flow on the TE of the airfoil. Owing to the separation, it is unnecessary
to significantly increase the h/D values, and the impingement points do not continue to move away from the cylinder.
Figure 5 shows the rms values of the drag coefficients of the cavity and the cylinder for each case. The coefficient of
the drag rms values is lowest at h/D=0.6 for both of cavity and cylinder, decreasing by about 8.3% for the cavity and
10.8% for the cylinder compared to the baseline case. However, the life coefficient is least affected by drag. The
tendencies of the coefficient drag of the cavity and cylinder graphs are similar.
Velocity magnitude, m/s
Downloaded by PURDUE UNIVERSITY on June 6, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-2848

a) h/D=0 b) h/D=0.6

Figure 4. Contour of time-averaged velocity on a plane section (x-y plane) at z=0.225m.

Figure 5. Cavity and cylinder RMS drag coefficient.

2. Acoustic results

The surface wall pressure fluctuations on the surface of the cavity and cylinder are used as inputs for acoustic
calculation. The acoustic receivers are located 2.5 m away from the cylinder in the z-direction and 0.35 m above the
bottom of cavity. The sampling frequency was 100kHz and 20000 samples were acquired. The sound pressure level
6
American Institute of Aeronautics and Astronautics
is calculated by segmental average by using a Hann
window. It is indicated from Fig. 6 that the noise level
decreased with an increased expansion surface ratio
(h/D). The expansion surface shapes of the cavity LE
contribute to the reduction of wall pressure fluctuations
on the cavity and the cylinder. The lowest value of
sound pressure level (SPL) on the cavity points and
cylinder surface is h/D=0.6. Table 1 shows the overall
SPL value and tonal noise level. The level value of the
case without any treatment on the cavity LE is highest
at 95.43 dB, and the SPL with the best case (h/D=0.6)
is 91.41 dB, which is about 4 dB lower than the baseline
case. Furthermore, the tonal noise, caused by pressure
fluctuation on the cylinder surface, was also reduced by
about 5 dB. It can be seen that the shear layer from the
cavity LE can also affect noise from the cylinder.
Downloaded by PURDUE UNIVERSITY on June 6, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-2848

Figure 6. SPL for expasion surface of cavity LE.


Therefore, the noise also can be decreased by reducing
the interaction between the shear layer and the cylinder
surface.
Table 1. Overal Sound Pressure Level (OSPL) and
B. Pantograph arm (cylinder) position study tonal noise level resulting from h/D study.
Another investigation in this project examines the
pantograph arm position. Figure 1b depicts the study,
h/D OSPL (dB) Tonal Noise level (dB)
which is investigated using one of the best results of the 0 95.43 85.96
previous studies (h/D=0.4). The purpose of this study is
to examine the effect of different positions of the arm. 0.2 92.31 85.75
The arm is placed in the middle of the cavity (L/2) in
0.4 91.62 84.65
other studies; however, in this study, the arm will be
moved backward according to the value of P, which 0.6 91.41 81.06
represents the distance between the middle of the cavity
and the arm position. The P value corresponds to the 0.8 82.32 82.32
depth of cavity D, and the position is studied from 1.0
D to 3.0 D.

1. Aerodynamics results

Figure 7 visualises the contour of the mean velocity magnitude on the plane section (x-z plane) at y = 0.02 m from
the bottom of the cavity. The cylinder and the cavity are shown to obtain an overview of the flow characteristic with
different positions. The velocity of the front of the cylinder region has a very slight increment away from the
impingement point. However, the velocity at the side region of the cylinder reduces with increments of distance P.
This is assumed to be caused by the difference of velocity magnitude of the surrounding cylinder for each case. As
shown in Fig. 7 a), the velocity of the surrounding cylinder seemed affected by the cavity inlet shear layer, which may
cause higher velocity surrounding the cylinder. However, when the cylinder is located near the cavity outlet, as shown
in Fig. 7 b), the velocity magnitude of the surrounding cylinder tends to be lower. Therefore, the pattern of flow around
the cylinder is slightly different from the flow pattern in Fig. 7 a). It can be seen that there is less interaction between
the shear layer and the cylinder surface as the cylinder moves away from the impingement point.

2. Acoustic results

Figure 8 shows far-field sound pressure spectrum at the receivers, which are located 2.5 m away from the cylinder
in the z-direction and 0.35 m above the bottom of the cavity with various positions of the cylinder. It is hard to
distinguish the difference between the levels. Especially, it is difficult to see a reduction of broadband noise levels,

7
American Institute of Aeronautics and Astronautics
Velocity magnitude, m/s
Downloaded by PURDUE UNIVERSITY on June 6, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-2848

a) P=0

b) P=2.5D
Figure 7. Contour of time-averaged velocity magnitude overview of cavity (top) around cylinder (bottom) on
plane section (x-z plane) at y=0.02m
which is mainly caused by cavity noise with
increments in the distance of its original position. The
overall sound pressure levels (OSPLs) of P=2.5 D are
reduced by less than 1 dB compared to the cylinders
original position case. The interesting fearture
observed is reduction of the peak noise levels,
occurred by cylinder surface pressure fluctuation. By
comparison, the noise level from the cylinder position
P=2.5 D is about 3 dB less than that from the cylinders
original position case, as shown in Table 2. Hence, it
is believed that far from the impingement point, the
interaction between the shear layer and the cylinder
surface is diminished and it causes reduction of tonal
noise from the cylinder.
Figure 8. Sound pressure level (SPL) for various position
V. Conclusion of cylinder.

In this research study, the flow behavior from a 1/10 scaled simplified pantograph arm and roof of a high-speed
train have been calculated using the IDDES model for near-field flow behavior and FWH analogy for far-field
acoustic noise levels. This project comprised two main investigations: expansion surface of the LE cavity (h/D), and
the pantograph arm position. The h/D study found that the recirculation region and vortices are significantly reduced
with increments of the expansion surface of the cavity LE (h/D), and this behavior continues until h/D=0.4.

8
American Institute of Aeronautics and Astronautics
Furthermore, these modified geometries influence the Table 2. Overal Sound Pressure Level (OSPL) and
surface pressure fluctuations on the cavity and the tonal noise level resulting from cylinder position study
cylinder. The surface pressure fluctuations are also
P OSPL (dB) Tonal Noise level (dB)
weakened on both surfaces. From this acoustic analysis,
a reduction of noise can be seen. Moreover, the 0 91.62 84.65
expansion of the LE surface decreases not only the
noise, but also helps reduce the drag and side force of 1D 91.45 84.63
the cavity and cylinder arm. In the cylinder arm 2D 91.49 82.49
position study, one notable result is the reduction of
peak noise levels. By avoiding the interaction shear 2.5D 90.98 81.58
layer and cylinder surface, the peak noise level can be
decreased. 3D 91.72 82.77
All simulations used a trimmed grid in this study.
However, a structured grid might be more beneficial. The structured mesh usually requires fewer mesh elements and
therefore incurs fewer computational costs and reportedly provides more accurate results, especially in simple
geometry. Owing to limited computational resources, it is difficult to use a finer mesh around the cylinder wake region,
Downloaded by PURDUE UNIVERSITY on June 6, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-2848

but a finer grid is required because many vortices are present in the region.

Acknowledgments
This research project has been supported by the Iridis 4 and Lyceum High Performance Computing Facility at the
University of Southampton.

References
1Jehanno, A., Palmer, D., and James, C., High Speed Rail and Sustainability, International Union of Railways, 2011.
2Martens, A., Wedemann, J., Meunier, N., Leclere., A., Deutsche Bahn, A. G., and Systemtechnik, D. B., High speed train

noise-sound source localization at fast passing trains, Deutsche Bahn AG, Sociedad Espanola de Acoustica, SEA, 2009.
3Talotte, C., Gautier, P. E., Thompson, D. J., and Hanson, C., Identification, modelling and reduction potential of railway

noise sources: a critical survey, Journal of Sound and Vibration, Vol. 267, No. 3, 2003. pp. 447-468.
4Grosche, F. R., and Meier, G. E. A., Research at DLR Gttingen on bluff body aerodynamics, drag reduction by wake

ventilation and active flow control, Journal of Wind Engineering and Industrial Aerodynamics, Vol. 89, No.14, 2001, pp. 1201-
1218.
5Yu, H. H., Li, J.C., and Zhang, H.Q., On aerodynamic noises radiated by the pantograph system of high-speed trains, Acta

Mechanica Sinica, Vol. 29, No. 3, 2013, pp. 399-410.


6Brentner, K. S., and Farassat, F., Analytical comparison of the acoustic analogy and Kirchhoff formulation for moving

surfaces, AIAA journal, Vol. 36, No. 8, 1998, pp. 1379-1386.


7Kynan, M., Best Practices: Volume Meshing, STAR South East Asian Conference, Singapore, 2012.
8Ikeda, M., Mitsumoji, T., Sueki, T., and Takaishi, T., Aerodynamic noise reduction in pantographs by shape-smoothing of

the panhead and its support and by use of porous material in surface coverings, Quarterly Report of RTRI, Vol. 51, No. 4, 2010,
pp. 220-226.
9Kurita, T., Hara, M., Yamada, H., Wakabayashi, Y., Mizushima, F., Satoh, H., and Shikama, T., Reduction of pantograph

noise of high-speed trains, Journal of Mechanical systems for Transportation and Logistics, Vol. 3, No. 1, 2010, pp. 63-74.
10Lee, J., and Cho, W., Prediction of low-speed aerodynamic load and aeroacoustic noise around simplified panhead section

model, Proceedings of the Institution of Mechanical Engineers, Part F: Journal of Rail and Rapid Transit, Vol. 222, No.4, pp.
423-431.
11Ikeda, M., and Mitsumoji, T., Numerical estimation of aerodynamic interference between panhead and articulated

frame, Quarterly Report of Railway Technical Research Institute, Vol. 50, No.4, 2009, pp. 227-232.
12Liu, X., Thompson, D. J., Hu, Z., and Jurdic, V., Aerodynamic noise from a train pantograph, The 21 st International

Congress on Sound an Vibration, Beiging, 2014.


13Brentner, K. S., and Farassat, F., Analytical comparison of the acoustic analogy and Kirchhoff formulation for moving

surfaces, AIAA journal, Vol. 36, No. 8, 1998, pp. 1379-1386.


14Kynan, M., Best Practices: Volume Meshing, STAR South East Asian Conference, Singapore, 2012.
15Baker, C., The flow around high speed trains, Journal of Wind Engineering and Industrial Aerodynamics, Vol. 98, No. 6,

2010, pp. 277-298.


16Moritoh, Y., Zenda, Y., and Nagakura, K., Noise control of high speed Shinkansen, Journal of Sound and Vibration, Vol.

193, No. 1, 1996, pp. 319-334.


17Nagakura, K., Localization of aerodynamic noise sources of Shinkansen trains, Journal of Sound and vibration, Vol. 293,

No. 3, 2006, pp. 547-556.


18Noh, H. M., Choi, S., Hong, S., and Kim, S. W., Investigation of noise sources in high-speed trains, Journal of Rail and

Rapid Transit,Vol. 228, No. 3, 2014, pp. 307-322.


9
American Institute of Aeronautics and Astronautics
19Zhu, J. Y., Hu, Z. W., and Thompson, D. J., Analysis of aerodynamic and aeroacoustic behaviour of a simplified high-speed

train bogie, Noise and Vibration Mitigation for Rail Transportation Systems. Springer, Berlin, Heidelberg, 2015, pp. 489-496.
20Kitagawa, T., and Nagakura, K., Aerodynamic noise generated by Shinkansen cars, Journal of Sound and vibration, Vol.

231, No. 3, 2000, pp. 913-924.


21Takeda, K., and Shieh, C. M., Cavity tones by computational aeroacoustics, International Journal of Computational Fluid

Dynamics, Vol. 18, No.6, 2004, pp. 439-454.


22Tracy, M. B., Plentovich, E. B., and Chu, J., Measurements of fluctuating pressure in a rectangular cavity in transonic flow

at high Reynolds numbers, NASA TM-4363, 1992.


23Khanal, B., A numerical investigation of the aerodynamic noise generation mechanism in transonic cavity flows, Ph.D.

Dissertation, Engineering System and Management Dept., Cranfield Univ., Cranfield, 2011.
24Cattafesta, L. N., Song, Q., Williams, D. R., Rowley, C. W., and Alvi, F. S., Active control of flow-induced cavity

oscillations, Progress in Aerospace Sciences, Vol. 44, No.7, 2008, pp. 479-502.
25Sarohia, V., and Massier, P. F., Control of cavity noise. Journal of Aircraft, Vol. 14, No. 9, 1977, pp.833-837.
26Zhang, X., Chen, X. X., Rona, A., and Edwards, J. A., Attenuation of cavity flow oscillation through leading edge flow

control, Journal of Sound and Vibration, Vol. 221, No.1, 1999. pp. 23-47.
27Lighthill, M. J., On the energy scattered from the interaction of turbulence with sound or shock waves, Proceedings of the

Cambridge Philosophical Society, Vol. 49, No. 3, 1953, pp. 531-551.


Downloaded by PURDUE UNIVERSITY on June 6, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-2848

28Howe, M. S., Acoustics of fluid-structure interactions, Cambridge university press, Cambridge, England, UK, 1998, pp. 317-

410.
29Doak, P.E., Fluctuating total enthalpy as the basic generalized acoustic field, Theoretical and Computational Fluid

Dynamics, Vol.10, No.1-4, 1998, pp. 115-133.


30Kol, V., Vortex identification: New requirements and limitations, International journal of heat and fluid flow, Vol.28,

No.4, 2007, pp. 638-65.


31Winkler, M., Becker, K., Doolan, C., Kameier, F., and Paschereit, C. O., Aeroacoustic effects of a cylinder-plate

configuration, AIAA journal, Vol. 50, No, pp. 1614-1620.

10
American Institute of Aeronautics and Astronautics

You might also like