Sie sind auf Seite 1von 124

DOTTORATO DI RICERCA IN INGEGNERIA MECCANICA E

INDUSTRIALE
CICLO XXIV
Experimental investigation of ow past open and partially covered
cylindrical cavities
Francisco Rodriguez Verdugo
Tutor: Prof. Roberto Camussi
Coordinatore: Prof. Edoardo Bemporad
Thesis submitted in partial fulllment of the requirements for the degree of Doctor of Philosophy
Rome, Italy, March 2012
Abstract
Flows past wall-mounted cavities are found in a wide variety of applications, including side-branches, organ
pipes, automobile sunroofs, inter-car gaps in trains and aircraft bays. Under certain conditions, ow excited
cavities can generate large pressure uctuations, undesirable noise and signicant structural loads. To
date, most of the studies have been focused on rectangular cavities while little attention has been given to
cylindrical cavities despite their widespread use.
Two dierent types of cylindrical cavities were experimentally investigated in low speed wind tunnels: an
open mouth cavity and a deep cavity with a small rectangular opening. The measurements included hot wire
anemometry, particle image velocimetry (PIV) and unsteady surface pressure measurements. Additionally,
numerical analysis of the test section/cavity systems were carried out with the nite element program
COMSOL Multiphysics and with a wave expansion method (WEM) code developed by the Trinity College
Dublin.
Important ow features are described by evaluating the pressure measurements conducted in several
positions over the walls of an open mouth cavity, the PIV measurements performed over horizontal planes
inside the cavity and the hot-wire measurements on the shear layer and on the wake of the cavity.
Pressure Fourier spectra evidence the presence of the rst three shear layer hydrodynamic modes at
frequencies well predicted by classical formulation for rectangular cavities (Rossiter). When the cavity is
open, the acoustic modes of the test section are found to be excited by the ow but when the cavity is
partially covered, the shear layer hydrodynamic modes are more likely to lock on the natural frequencies of
the cavity. The position of the opening has an inuence on the lock-on acoustic modes.
The acoustic energy generated by the shear layer is calculated by applying the vortex sound theory of
Howe: the ow velocity and the vorticity are extracted from the PIV data and the acoustic particle velocity
eld from the WEM calculation. The acoustic sources are localised in space and quantied over an acoustic
period providing insight into the sound production of ow-excited partially covered cylindrical cavities.
i
Acknowledgements
Vorrei ringraziare innanzitutto il prof. Roberto Camussi, per avermi accolto nel suo gruppo di ricerca.
Questi quattro anni passati a Roma Tre, prima per la tesi magistrale e dopo per il dottorato, sono stati
molto formativi, non solo dal punto di vista professionale ma anche dal punto di vista personale.
Un particolare ringraziamento va ai miei cari colleghi della sezione di termouidodinamica e aerodinamica
per i loro aiuti, i loro consigli, e per tutti i momenti passati insieme allinterno del dipartimento e al di fuori.
In ordine alfabetico ringrazio: Giovanni Aloisio, Alessandro Di Marco, Dajana Giulieti, Emanuele Giulietti,
Daniele Grassucci, Silvano Grizzi, Riccardo Moscatelli, Tiziano Pagliaroli, Alessandra Parlato e Francesco
Tomassi.
Ringrazio a tutti gli stagisti e i tesisti che mi hanno circondato durante la mia attivit`a di ricerca. In
particolare ringrazio in ordine cronologico a Chistophe Perge, Stefano Valerio, Alessandro Guerriero, Federico
Gargano, Gabriele Baiocco e Ludovica Penten`e per il loro supporto tecnico in galleria del vento e ad Andrea
Serrani per le sue simulazioni acustiche.
Non mancherei di ringraziare i dottorandi del GRACO per le pause pranzo, per i loro aiuti con L
A
T
E
X,
per le cene, per i loro consigli durante la mia ricerca di lavoro, per i TrovaRoma, per lInternazionale, per le
giornate in montagna, per il calcetto e tantaltro. Loro sono: Alessandro Anobile, Paolo Gradassi, Eugenio
Lombardi, Simone Menicucci, Emanuele Piccione.
This research project has been supported by a Marie Curie Early Stage Research Training Fellowship of
the European Communitys Sixth Framework Programme under Contract number MEST CT 2005 020301.
This nancial support was greatly appreciated. I thank Prof. Aldo Rona for leading the AeroTraNet project,
for bringing us a nice cavity model back in 2007 and also for the useful acoustic measurement equipment that
he lent us. I would like to express my thanks to all the AeroTraNet fellows for the interesting discussions
that we had during the meetings and the conferences. Many thanks to Marco Grottadaurea for the fructose
collaboration.
Parmi les participants de lAeroTraNet, je tiens egalement `a citer Antoine Guitton et Julien Grilliat avec
qui jai partage de magniques moments `a Roma Tre. Je souhaite leur exprimer mes sinc`eres remerciements
pour leurs precieux conseils et lappui quils ont pu me fournir.
I would like to thank my second advisor, Dr. Gareth Bennett for giving me the opportunity to spend
thirteen month in the Trinity College. Dublin was, without doubts, a wonderful experience. Moreover I
iii
would not have been able to complete this work without his guidance and motivation. Many thanks to
Dr. David Stephens who designed the Helmholtz resonator experimental rig that I used. His contribution to
this work has been invaluable.
I also thank Dr. Craig Meskell and Prof. John Fitzpatrick for the interesting discussions about the physical
phenomena studied. Special thanks go to Shane Finnegan whose experimental methodology has stimulated
much of my research. I also want also to acknowledge my Fluids Labs colleagues Miguel Pedroche, Ian
Davis and John Mahon. Furthermore I want to give a special thanks to Eoin King, Donal Lynch, Dorota
Skupi nska, Emer Walsh and Rory Stoney with who I spent many late evenings in the dungeon. I will not
forget to thank my lunch-break mates from the Mechanical and Manufacturing Engineering Department:
Paul Ervine, Paul Harris, Karl Brown, Peadar Golden, Emma Brazel, Stuart Murphy, Kevin Kerrigan and
Robert Smyth.
Many thanks to Prof. Marc Jacob for reading this thesis. I have greatly appreciated his meticulous
corrections and suggestions.
Desde luego, mi mas amplio y sincero agradecimiento a mi familia, cuyo apoyo incondicional hizo posible
la culminaci on mi formacion universitaria en Europa.
Ringrazio Sandro e Ivana che mi hanno generosamente accolto nella loro famiglia facendomi sentire a
casa.
Vorrei inne ringraziare Ambra, per la sua pazienza e il suo amore e per essermi stata vicina in tutti
questi anni.
Grazie a tutti di cuore.
Table of contents
Abstract i
Acknowledgements iii
Table of contents vii
Introduction 1
1 Background 3
1.1 Open mouth cavities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1.1 Rectangular cavities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1.2 Cylindrical cavities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2 Partially covered cavities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.2.1 Helmholtz resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.2.2 Shear layer over the mouth of a Helmholtz resonator . . . . . . . . . . . . . . . . . . . 15
1.3 Noise source characterisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
I Cylindrical cavity with open mouth 19
2 Experimental set-up 21
2.1 Wind tunnel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.2 Test model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.3 Instrumentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.3.1 Pitot-static tube . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.3.2 Hot wire anemometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.3.3 Microphones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.3.4 Particle image velocimetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.3.5 Data acquisition card and processing of the pressure signals . . . . . . . . . . . . . . . 26
2.4 Test conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
v
TABLE OF CONTENTS
2.4.1 Flow conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.4.2 Incoming boundary layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.4.3 Background pressure uctuations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.5 Measurement matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3 Acoustic mode calculation 31
3.1 The acoustic modes of an open-closed cylindrical cavity . . . . . . . . . . . . . . . . . . . . . 31
3.2 The acoustic modes in a wind tunnel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.3 The computational model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.4 The geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.5 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.5.1 Acoustic modes without the cavity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.5.2 Acoustic modes with the cavity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4 Experimental results 43
4.1 Overall aerodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.1.1 Shear layer topology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.1.2 Wake topology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.2 Description of some ow features inside the cavity . . . . . . . . . . . . . . . . . . . . . . . . 44
4.3 Unsteady response to a grazing ow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.3.1 Pressure response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.3.2 Nondimensionalization process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.3.3 Velocity response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.4 Spectral decomposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.4.1 Spectral decomposition and analysis on the symmetry plane . . . . . . . . . . . . . . . 52
4.4.2 Analysis on the cavity walls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5 Conclusion 59
II Cylindrical cavity with partially closed mouth 61
6 Experimental set-up 63
6.1 Overview of the experimental rig . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
6.2 Design of the experimental rig . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
6.3 Opening details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
6.4 Instrumentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
6.4.1 Pitot-static tube . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
6.4.2 Microphones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
vi
TABLE OF CONTENTS
6.4.3 Hot wire anemometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
6.4.4 Particle image velocimetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
6.4.5 Data acquisition card . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
6.5 Phase-averaging technique . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
6.6 Boundary layer characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
7 Acoustic mode calculation 73
7.1 Analytical solution of the Helmholtz equation . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
7.2 Wave Expansion Method (WEM) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
7.2.1 Overview of the method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
7.2.2 Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
7.2.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
7.3 Helmholtz resonance frequency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
7.4 Response of the resonator to an external excitation . . . . . . . . . . . . . . . . . . . . . . . . 80
7.4.1 Acoustic excitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
7.4.2 Boundary layer excitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
8 Experimental results 83
8.1 Response of the resonator to a grazing ow . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
8.1.1 Baseline opening: case L45EU . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
8.1.2 Strength of lock-on . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
8.1.3 Inuence of the location of the opening . . . . . . . . . . . . . . . . . . . . . . . . . . 86
8.2 Shear layer dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
8.2.1 First shear layer mode . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
8.2.2 Second shear layer mode . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
8.3 Acoustic power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
8.3.1 Computation of the acoustic particle velocity . . . . . . . . . . . . . . . . . . . . . . . 89
8.3.2 Time-averaged acoustic power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
9 Conclusion 99
10 Summary 101
A Acoustic power 103
References 105
List of publications 113
vii
TABLE OF CONTENTS
viii
Introduction
Motivation for the research
Flows past wall-mounted cavities are found in a wide variety of applications, including side-branches, organ
pipes, automobile sunroofs, inter-car gaps in trains and aircraft bays. Under certain conditions, ow excited
cavities can generate large pressure uctuations, undesirable noise and signicant structural loads.
To date, most of the studies have been focused on rectangular cavities while little attention has been
given to cylindrical cavities despite their widespread use. In the aerospace sector, cylindrical cavities are
present as pressure relief valve of the fuel vents (gure 1), circular anti-icing vent holes or cylindrical landing
gear wheel wells, just to cite some examples.
The purpose of this thesis was to extend the existing knowledge in this area through an experimental
investigation. Two dierent cases were studied: an open mouth cavity and a partially covered cavity. The
rst cavity has an aspect ratio of 1.357 which was dictated by the need of reproducing typical geometries
present on commercial aircraft. The second cavity (mouth partially covered) was designed in such a way
that dierent acoustic resonances can be excited.
Frameworks
Part of the PhD work presented here was performed in the framework of the AeroTraNet project, an
Early Stage research Training (EST) program funded by Marie-Curie Actions. The AeroTraNet project
was launched in 2006 in order to study the unsteady ow in selected airframe cavities of a wide-body civil
transport aircraft. This four-year European initiative brought together the University of Leicester, the Uni-
versit`a degli Studi Roma Tre, the Politecnico di Torino and the Institut de Mecanique des Fluides de Toulouse
around a common research topic. This successful program brought to an extensive list of publications which
can be found in the AeroTraNet ocial webpage: http://aerotranet.imft.fr.
The Universit`a degli Studi Roma Tre encourages PhD students to spend a period of time in a foreign
research center. Part of the experimental results was therefore obtained during a thirteen months stay
at the Trinity College of Dublin (TCD). The School of Engineering at Trinity was founded in 1841 and
is one of the oldest Engineering Schools in the English-speaking world. The Department of Mechanical
1
Introduction
Figure 1: Photograph of a commercial aeroplane. Detail of circular pressure relief valve on the lower surface of the
wing.
and Manufacturing Engineering has been conducting research for many years on modelling and analysis of
ow/structure interactions and vibro-acoustic problems. The results achieved by the author was an invited
research student at TCD were presented in four dierent international conferences (refer to list of publication
on page 113).
Outline of the thesis
This thesis is organized as follows: the rst chapter introduces the state of the art; the two main parts,
each of which contains four chapters, treat the open mouth case and the partially closed mouth case; the
nal chapter summarises the main results. Hereafter the organization of the two main parts of the thesis is
described: chapters 2 and 6 detail the experimental facilities and the measurement techniques; in chapters 3
and 7 the acoustic modes are calculated analytically and numerically; the experimental results are given in
chapters 4 and 8; chapters 5 and 9 draw the conclusions of each part.
2
Chapter 1
Background
This chapter is not intended to be an exhaustive literature survey on cavity ow studies: special emphasis
is given to research on deep cylindrical cavities with open or partially covered mouths in low Mach number
ows. For general surveys on cavity ow, see for instance Rockwell & Naudascher (1978, 1979). More
recently, Cattafesta et al. (2003) and Rowley & Williams (2006) have reviewed the studies on cavity ow
with a particular focus on the control of ow-induced resonance.
1.1 Open mouth cavities
1.1.1 Rectangular cavities
Early studies and oscillations classication
Anatol Roshko and Krishnamurty Karamcheti, both from California Institute of Technology, performed the
rst mayor studies on rectangular cavities in the 1950s. Roshko (1955) reported the pressure distribution
on the walls of cavities tested in a low speed wind tunnel. Karamcheti (1955, 1956) studied the acoustic
radiation of a rectangular cavity in a transonic wind tunnel by means of schlieren observation, interferometry
and hot-wire anemometry. These pioneer researchers are still largely cited nowadays and have inspired many
studies on ow-excited cavities over the years.
Rockwell & Naudascher (1978) divided the shear layer driven cavity oscillations into three categories:
uid-elastic, uid-dynamic, and uid-resonant. The primary condition for uid-elastic interactions to exist
is the structural elasticity of the cavity walls. Fluid-dynamic oscillations are triggered by the interaction
between the upstream edge of the cavity and the pressure wave generated on the cavitys downstream edge
by the impingement of the shear layer. A uid-resonant interaction may occur if the acoustic wavelength is
of the same order of magnitude than the dimensions of the cavity. The last two cavity oscillation categories
are further analyzed in the next two sections.
3
1.Background
Fluid-dynamic oscillations
The uid-dynamic oscillations, also called self-sustained cavity oscillations or shear layer hydrodynamic
modes are the cavity oscillations in an acoustic-free system. Figure 1.1 shows the essential features involved
in the feedback mechanism which can be summarized in the following way: creation of an instability at the
upstream edge which is amplied as it is convected downstream until it interacts with the downstream edge
of the cavity; the impingement generates an acoustic perturbation which propagates upstream and triggers
the generation of another instability at the upstream edge of the cavity.
Many studies have been conducted in order to predict of the frequencies of the shear layer modes over
a rectangular cavity. Rossiter (1964) proposed a semi-empirical formula based on the vortex shedding
phenomenon:
St =
f L
char
U

n
M +
1

(1.1)
where n is the order of the shear layer mode, L
char
is the characteristic length of the shear layer,
describes the phase delay between the hydrodynamic forcing and the acoustic feedback and is ratio
between the convection velocity in the shear layer and the free stream velocity. This formula correctly tted
Rossiter (1964)s experimental data which were obtained with pressure transducers and ow visualisation
(shadowgraph) for Mach numbers between 0.3 and 1.2.
Even if many dierent values for the convection speed coecient can be found in the literature, the
original value proposed by Rossiter ( = 0.57) is often used. Chatellier et al. (2004) and El Hassan et al.
(2007) argued that there is no need to consider a phase delay when the convection speed is much lower than
the speed of sound and therefore = 0. Usually the selected characteristic length L
char
is the length of the
cavity in the streamwise direction.
Over the years, equation 1.1 has been subjected to small changes introduced after analytical developments:
see for instance Bilanin & Covert (1973), Heller & Bliss (1975), Block (1976) and Howe (1997).
Flow-acoustic coupling
The shear layer hydrodynamic modes can excite dierent types of acoustic modes. Plumblee et al. (1962)
showed that for shallow rectangular cavities the predominant excited acoustic mode is the lengthwise mode
whereas for cavities of aspect ratio higher than unity the depth acoustic mode is the one excited. Similar
ow-acoustic coupling mechanisms exist in the case of coaxial side branches (Arthus & Ziada (2009); Dequand
et al. (2003); Oshkai & Yan (2008); Ziada & Shine (1999)) or in the case of axisymmetric internal cavities
mounted on pipes (Aly & Ziada (2010)).
Ziada et al. (2003) noticed that when a shallow rectangular cavity is mounted in a closed test section
wind tunnel a ow-acoustic coupling can occur between the shear layer modes and the acoustic modes of
the cavity-tunnel. Alvarez & Kerschen (2005) analytically evaluated the inuence of the connement on the
acoustic resonances of a two-dimensional cavity. In Kerschen & Cain (2008) a good agreement was found
4
1.Background
Figure 1.1: Schematic diagram of cavity ow.
between the predicted and the experimental trapped modes. In order to avoid the connement eects, Yang
et al. (2009) studied a deep rectangular cavity mounted at the outlet of a duct.
1.1.2 Cylindrical cavities
Mean ow
As opposed to rectangular cavities for which the ow can be considered two-dimensional when the spanwise
length is large compared to the streamwise length, the ow in a cylindrical cavity is always fully three-
dimensional. The organization of the mean ow depends exclusively on the aspect ratio H/D of the cavity,
where H and D denote respectively the depth and the diameter. Gaudet & Winter (1973), Hiwada et al.
(1983) and Dybenko & Savory (2008) showed that the ow is asymmetric with respect to the central stream-
wise plane for aspect ratios between 0.2 and 0.8. Through long pressure measurements, Hiwada et al. (1983)
identied two dierent dynamics: the ow can either ap for H/D = 0.2 0.4 or switch orientation for
H/D = 0.4 0.7. For aspect ratios such as H/D < 0.2 or 0.8 < H/D, the ow inside the cavity was found
to be stable and symmetric.
The parametric study of Gaudet & Winter (1973) includes a set of oil-ow visualisations at the walls of
9 dierent cavities (H/D between 0.04 and 1.34). The employed experimental technique revealed impor-
tant information about the ow near the walls. Gaudet & Winter (1973) presented such information with
streamline pattern drawing. However, as they mentioned, a certain amount of imagination has been used
in drawing the streamline patterns. An example of these results is shown in gure 1.3: a strong asymmetric
case (H/D = 0.47) as well as a symmetric one (H/D = 1.07) are presented.
Haigermoser et al. (2009) investigated a cylindrical cavity of aspect ratio H/D = 0.5 with stereo and
tomographic particle image velocimetry achieving the rst detailed description of the three-dimensional
macroscopic organisation of the ow inside a cylindrical cavity. Chicheportiche & Gloerfelt (2010), Desvi-
5
1.Background
Figure 1.2: Schematic of a cylindrical cavity
gne et al. (2010), Marsden et al. (2010) and Mincu et al. (2009), within a Fondation de Recherche pour
lAeronautique & lEspace (FRAE) project called AEROCAV, studied numerically a cylindrical cavity (H
= D = 10 cm). Their results were compared to experiments performed in two dierent wind tunnels: the
Ecole Centrale de Lyons anechoic wind tunnel and the ONERAs F2 closed section wind tunnel (experi-
mental details can be found in Marsden et al. (2008) and Mincu et al. (2009)). Even if the main objective of
the AEROCAV project was the investigation of the ow unsteadiness and the noise generated by cylindrical
cavities, some important results on the mean ow were also obtained. Some results are reported in gure 1.4.
Grottadaurea (2009) simulated the ow over a deep cylindrical cavity (H/D = 1.4) with a Detached Eddy
Simulation (DES). One of the important results of this study, from the aerodynamic point of view, is the
analysis of the wake behind the cavity. Specic features are described as for example the counter-rotating
convective eddies generated by the cavity downstream edge. Figure 1.5 gives an example of the results
obtained.
Shear layer hydrodynamic mode
An important question for the aeroacoustic viewpoint is if Rossiters formula can be used to estimate the
frequencies of the shear layer modes on cylindrical cavities. Bruggeman et al. (1991) assumed that circular
side branches can be treated as rectangular side branches for the prediction of the shear layer modes as long
as an eective length W
eff
is used:
W
eff
=
D
4
(1.2)
This eective length represents the streamwise dimension of a rectangular opening with the same surface
area and the same spanwise dimension (D) as the original circular opening. If the upstream edge of the side
branch is not sharp, an additional term should be added to W
eff
(see Bruggeman et al. (1991) or Tonon
et al. (2011)).
Czech et al. (2006) recently studied an array of circular vent holes in a wind tunnel. In order to accom-
modate Rossiters formula, another expression for the eective length, based on an equivalent streamwise
6
1.Background
(a) (b)
(c) (d)
(e)
Figure 1.3: Oil-lm ow visualisation and streamline patterns of a cylindrical cavity of aspect ratio H/D = 0.47 (a,
b) and H/D = 1.07 (c, d), Gaudet & Winter (1973). Legend (e).
7
1.Background
(a) (b) (c)
Figure 1.4: Numerical results: (a) Non-dimensionalized mean streamwise U velocity proles, in black lines, and vector
plot of in-plane velocity (U, W) as grey arrows in the middle spanwise plane of the cavity; (b) Non-dimensionalized
mean cross-stream V velocity proles and vector of in-plane velocity (V, W) in the middle streamwise plane of the
cavity. Free-stream ow velocities of 70 m/s, Marsden et al. (2010). (c) Flow dynamics using Q criterion applied to
the mean ow, iso-surfaces of Q = 0.25(U/D)
2
, Mincu (2010)
Figure 1.5: Grottadaurea numerical simulation of a H/D = 1.375 cylindrical cavity. ReD = 34800 and M = 0.235.
(a) View of the cavity with streamlines over streamwise planes. The mean velocity contours are also plotted over to
the walls. A color-bar range from blue to red is used (0.1 < u/u < 0.5). Image taken from Rodrguez Verdugo
et al. (2010)
8
1.Background
length for a square vent hole of the same area, has been introduced:
L
eff
=

D
2
(1.3)
Figure 1.6 shows the equivalent rectangular openings according to Bruggeman et al. (1991) and Czech
et al. (2006).
Figure 1.6: Schematic representing the original circular opening and the equivalent rectangular opening according
to Bruggeman et al. (1991) and Czech et al. (2006).
Mery et al. (2009) found a better agreement with Block (1976) s formula in the prediction of the shear
layer modes over a cylindrical cavity:
St =
n
1
k
R
+M(1 +
0.514
L/D
)
(1.4)
where k
R
is the real part of the wave number of the disturbance travelling downstream. This formula
includes the eect of the bottom-reected acoustic wave originated at the downstream edge.
Acoustic modes of an open mouth cylindrical cavity
When studying the acoustic resonances of a cavity without mean ow, it is fundamental to distinguish the
open and the closed mouth cases: the boundary conditions inuence indeed drastically the acoustic modes.
The presence of a mean ow, especially a shear layer over the mouth of the cavity, is believed to change
the characteristics (shape and frequency) of the modes. In Rona (2007), an extensive discussion about
which boundary condition should be imposed at the opening of a cavity when solving the acoustic eigenvalue
problem for low Mach number ows is given. Rona (2007) chose a simple acoustic reecting boundary
condition for the cavity open end while admitting that this represents a strong hypothesis in his model.
In the next section, important experimental results will show that the acoustic modes of an open mouth
cylindrical cavity excited by a low Mach number ow are better estimated when a no reection condition
is imposed at the open end. Therefore, in the following the acoustic modes of a cylindrical cavity with an
open end condition are described.
9
1.Background
The rst longitudinal mode of a deep cylindrical cavity, sometimes referred as the depth acoustic mode or
as the quarter-wavelength, is the lowest tone which can resonate in the cavity. Euler and Lagrange assumed
the pressure at the mouth of an open pipe to be zero and therefore the wave-length of the rst longitudinal
mode to be exactly 4H. Rayleigh (1894) introduced a correction for the open end of a tube with an innite
ange. The general expression for the frequency of the rst longitudinal mode of a H-long open-closed pipe
is:
f
c
=
c
4(H +R)
(1.5)
where c is the speed of sound and is the end correction factor and R = D/2 is the radius of the pipe.
The original value calculated by Rayleigh (1894) for an innite ange tube is = 0.8242. Nomura et al.
(1960) studied the sound radiation from a anged circular pipe with Weber-Schafheitlin type integrals and
Jacobis polynomials. They reported how the end correction varies with the frequency of the incident plane
wave and found lim
ka0
= 0.8217 which is close to the coecient calculated by Rayleigh. Unfortunately,
Nomura et al. (1960) did not propose a simple expression tting their results for . Norris & Sheng (1989)
studied the same problem with Greens functions and proposed a rational function to approximate their
numerical solution:

0.82159 0.49(kR)
2
1 0.46(kR)
3
(1.6)
where k is the wavenumber of the incident acoustic wave. According to Norris & Sheng (1989), their own
results graphically match the Nomura et al. (1960) ones. Another function that approximates correctly the
numerical results of Nomura et al. (1960) and that removes the singularity of equation 1.6 is (Dalmont et al.
(2001)):
0.8216
_
1 +
(0.77kR)
2
1 + 0.77kR
_
1
(1.7)
Figure 1.7: Correction coecient as a function of kR predicted by equation 1.7.
10
1.Background
The frequencies of higher order longitudinal modes can be predicted by the generalization of equation 1.5.
f
c
=
c(2q + 1)
4(H +R)
(1.8)
where q N and 2q +1 is the order of the mode. This expression guarantees a pressure node at a distance
of R from the opening.
The longitudinal modes are not the only resonant acoustic modes of a cylindrical cavity: there are also
the azimuthal and the radial modes and all the combinations of these three types of modes. Marsden et al.
(2008), while testing a cylindrical cavity in an anechoic wind tunnel, found a peak on the pressure spectra
whose frequency was much higher than the frequencies of the quarter wave length: The peak at 2160 Hz
[...] is more strongly visible on the cavity oor, and it exhibits a strong symmetry with respect to the ow
direction, almost disappearing for = . This peak is not a multiple of one of the two main frequencies
present, nor of the cavity resonant frequency. Its origin is at present not clearly identied.. Mincu et al.
(2009) performed a numerical simulation of a cylindrical cavity with exactly the same dimensions. They
demonstrated that the 2160 Hz peak reported by Marsden et al. (2008) corresponds in fact to the rst
azimuthal mode. Furthermore, they described four other modes present in the pressure spectra. The shape
of the modes are given in gure 1.8.
Figure 1.8: Shape of the acoustic modes (pressure amplitude) inside a cylindrical cavity at the half depth plane
given by Large Eddy Simulation (Mincu (2010)). For the coordinate system see gure 1.4.
The acoustic modes can be calculated analytically from the Helmholtz equation which is an eigenvalue
problem: the eigenvectors represent the pressure distribution of the acoustic modes of the cavity and the
eigenvalues represent the square of the natural frequencies. In section 3.1, the acoustic modes of an open
mouth cylindrical cavity are calculated analytically whereas in section 7.1 the closed case is treated.
11
1.Background
Acoustic-hydrodynamic coupling
Similarly to rectangular cavities, the hydrodynamic modes of a shear layer spanning a cylindrical cavity can
be amplied when their frequencies match the acoustic modes of the cavity. Parthasarathy et al. (1985)
performed a series of experiments on deep cylindrical cavities at low Mach numbers (M = 0.12 - 0.24). They
reported a strong whistle when the quarter wave length matches the rst shear layer mode. Marsden et al.
(2008) found a strong coupling mechanism between the rst two shear layer modes and the quarter wave
length of a cylindrical cavity tested in an anechoic wind tunnel. Dybenko & Savory (2008) did not nd a
uid acoustic coupling for any of their cylindrical cavities (H/D = 0.20, 0.47 and 0.70) tested at 27 m/s
because, as it is mentioned in the paper, the rst shear layer mode was expected at 145.5 Hz and the quarter
wave length at 1156 Hz for the deepest cavity studied.
1.2 Partially covered cavities
Cavities whose openings are partially covered, commonly called Helmholtz resonators, are found not only in
the transportation industry (window bueting in cars/trains, aircraft landing gear wheel well) but also in
musical instruments (jug bands) or in duct applications (side branches).
When these systems are exposed to a grazing ow, periodic pressure uctuations can be generated inside
the cavity. However, as noticed by Rayleigh (1894) in 310, Panton & Miller (1975a) and De Metz et al.
(1977), two dierent oscillatory mechanisms can be excited: either periodic compressions of the volume of
uid inside the cavity, commonly called the Helmholtz resonance, or the excitation of a standing acoustic wave
in the resonator. Both mechanisms fall into the Rockwell & Naudascher (1978) category of uid-resonant
oscillations.
1.2.1 Helmholtz resonance
The gravest mode of vibration of a cavity communicating with the exterior space though an orice is the
Helmholtz resonance. This acoustic mode can be excited either by acoustic pressure uctuations or by uid-
dynamic pressure uctuations. The response of a Helmholtz resonator acoustically excited is presented rst,
followed by the examination of the ow excited case.
The classical Helmholtz theory
The easiest way to model the Helmholtz resonance is with the mechanical analogy of a spring-mass system.
The air in the opening acts as lumped mass and the air volume inside the cavity acts as spring. If an external
perturbation (incident acoustic waves for example) displace the slug of air in the neck, the volume inside
the cavity undergoes an adiabatic transformation (compression or dilatation). By applying Newtons second
law, one can nd the second order dierential equation governing the displacement of the slug of air in the
neck. The natural frequency of vibration of the system, Helmholtz resonance, takes the owing expression:
12
1.Background
(a) (b)
Figure 1.9: Two dierent types of Helmholtz resonator: (a) thin opening cavity and (b) long neck resonator
f
HR
=
c
2

S
V l
eq
(1.9)
where c is the sound speed, V the volume of the resonator, S is the cross-section area of the orice, l
eq
is the equivalent neck length. For the extended development, refer any textbook of acoustics, for example
Rienstra & Hirschberg (2011). This generic formula can be applied to dierent types of Helmholtz resonators
by correctly estimating the equivalent neck length which is the height of the slug of air. Figure 1.9 gives a
schematic drawing of two types of cavities: the slug of air is represented with a dashed line.
Conceptually, l
eq
is the actual length of neck plus an interior and an exterior end correction (l
int
and
l
ext
):
l
eq
= l + l
int
+ l
ext
(1.10)
Dierent expressions for the end corrections have been developed depending on the shape of the aperture,
on the ange and on the dimensions of cavity. For circular openings, the Rayleigh end correction formerly
introduced in equation 1.5 is often used for the exterior part (l
ext
= 0.8242a). In order to apply the
Rayleigh end correction for non-circular openings, an eective radius a
eff
= 1.06 S
3/4
U
1/2
, where U is the
perimeter of the opening, has proven to give reasonable agreement with experiments (Crighton et al. (1994)).
Ingard (1953) proposed dierent expressions for the interior end correction, based on the shape of both the
cavity and the opening and on the position of the opening. Interior and exterior end corrections are often
grouped in a single parameter 2l.
Inuence of the geometry
Some researchers have shown that the classical Helmholtz theory presented so far does not correctly predict
the frequency of the Helmholtz resonance because it does not take into account the geometrical characteristics
13
1.Background
of the cavity. Selamet et al. (1995), for example, investigated the eect on the resonance frequency of changing
the aspect ratio of a cylindrical cavity while maintaining its volume. Their computational and experimental
results showed that the resonance frequency depends on the cavitys aspect ratio: the deeper the cavity is,
the lower the frequency is.
Panton & Miller (1975b) proposed a transcendental equation for the resonant wavenumbers k of a cylin-
drical Helmholtz resonator based on one-dimensional wave propagation inside the cavity:
l
eq
A
H S
kH = cot(kH) (1.11)
where A is the cavity cross section area. In Panton & Miller (1975b)s development, the slug of air in
the neck was treated as a lumped mass. A more sophisticated analysis was performed by Tang & Sirignano
(1973) who assumed one-dimensional wave propagation not only inside the cavity but also on its neck. The
rst-order approximation gives:
tan(kl) tan(kH) = S/A (1.12)
This expression yields to equation 1.11 when the cavity has a short neck (kl 1) because in this
case tan(kl) kl. Furthermore, when the wavelength is much larger than the dimensions of the cavity
(kH 1), equation 1.11 reduces to the classical formula for the Helmholtz resonance (equation 1.9) because
cot(kH) 1/kH. It is interesting to note that Selamet et al. (1995), through a dierent conceptual develop-
ment, found equation 1.12 as well. The trend of the Helmholtz resonance frequency observed experimentally
is correctly predicted by equation 1.12 even if some quantitative discrepancies have been noticed for small
aspect ratios H/D (Selamet et al. (1995)). Tang & Sirignano (1973) recommended the addition of the end
corrections to the length of the orice and the inclusion of the ow contraction eects to the model for small
aspect ratio cavities in order to improve the prediction of the resonant modes.
The model of Panton & Miller (1975b) and Tang & Sirignano (1973) does not only predict the frequency
of the Helmholtz resonance but also the longitudinal modes of the cavity. In 1.10 a graphical resolution of
equations 1.11 and 1.12 is given.
Panton & Miller (1975b) also proposed an explicit expression for the Helmholtz resonator that improves
the classical Helmholtz resonance formula (equation 1.9) by retaining two terms in equation 1.11:
f
HR
=
c
2

S
V l
eq
+H
2
S/3
(1.13)
This formula has been used by dierent authors to predict the Helmholtz resonance (Panton (1990) or
Mechel (2002)).
14
1.Background
(a) (b)
Figure 1.10: Graphical resolution of two dierent transcendental equations for the resonant wavenumbers. (a) Panton
& Miller (1975b) : blue line cot(kH) and red line
leq A
H S
kH. (b) Tang & Sirignano (1973): black line cot(lkH/H) S/A
and magenta line tan(kH)
Flow eect on the resonance frequency
Several researchers have reported an increase of the Helmholtz resonance frequency when the cavity is excited
by a ow. Anderson (1977), while testing side branch Helmholtz resonators of dierent sizes in a circular
duct with fully developed turbulent ow, reported an increase of the Helmholtz resonance frequency for
ow velocities higher than 30 m/s. Their 53.1 mm long cavity underwent a 100 Hz shift of the resonance
frequency from 250 Hz at 0 m/s to 350 Hz at a ow velocity of 80.7 m/s. For ow velocities lower than 30
m/s the resonant frequency did not change. Phillips (1968) reported an increase of f
HR
between 20 m/s and
80 m/s for a partially covered cavity tested in a wind tunnel. The experiments of Panton & Miller (1975a)
on the fuselage of a glider with a free-stream speed of 30 m/s also indicated an increase in the frequency of
the fundamental acoustic mode of dierent Helmholtz resonators. Their hot wire measurements assessed a
fully developed boundary layer in the orice region.
Zoccola (2000), reported in his PhD thesis a decrease of the resonances frequency of ow excited
Helmholtz resonators. He tested three dierent cavities and measured the frequency f
HR
for a purely acous-
tic excitation (0 m/s) and for a 6.9 m/s ow: 346 Hz (333 Hz), 636 (554 Hz) and 436 Hz (416 Hz). Zoccola
(2000)s experiments were done at a ow speed inferior than the velocity range analysed by Anderson (1977),
Phillips (1968) and Panton & Miller (1975a). Unfortunately Zoccola (2000) did not give the characteristics
of his boundary layer.
1.2.2 Shear layer over the mouth of a Helmholtz resonator
Of particular interest is the understanding of the shear layer organization when the cavity is subject to a uid-
resonant oscillations. Elder (1978) experimentally investigated a deep cylindrical cavity with a rectangular
opening. He characterized the shear layer with a hot wire probe sampled simultaneously with a microphone
15
1.Background
at the bottom wall the cavity: after phase-averaging the velocity, shear layer proles at dierent phases
during an acoustic cycle were presented. Elder (1978) also introduced the concept of the interface wave
surface which is the locus of the inection points of the shear layer proles along the opening. He proposed
a model to predict interface wave surface based on an initial displacement wave generated by the traverse
acoustic particle motion. Good agreement was found between the experiments and the proposed model.
Nelson et al. (1981) used two-component Laser Doppler Anemometry (LDA) and a ow visualisation
technique (stroboscopic light) to characterize the shear layer spanning the rectangular opening of a cuboidal
Helmholtz resonator. The free stream velocity chosen for the experiments was 22 m/s which corresponds to
the rst shear layer mode strongly coupled with the Helmholtz resonance of the cavity. At the upstream lip
they observed the shedding of a single vortex per acoustic cycle.
Ma et al. (2009) studied the shear layer over a rectangular Helmholtz resonator with PIV. They analysed
the same resonant case as Nelson et al. (1981): the rst shear layer mode coupled with the Helmholtz
resonance. However Ma et al. (2009) reported phase-averaged contours of the spanwise vorticity for three
dierent velocities corresponding to one strongly resonant case and two weakly resonant cases. They found
that the shear layer has a sheet-like character in the region closer to upstream edge and tends to roll up into
a single discrete vortex in the downstream portion of the opening (gure 1.11).
Figure 1.11: Contours of phase averaged vorticity for three dierent ow speeds. In all the cases the predominant
acoustic mode is the Helmholtz resonance and the amplied shear layer instability is the rst hydrodynamic mode
(Ma et al. (2009)).
16
1.Background
1.3 Noise source characterisation
Aeroacoustic source characterization has been facilitated by the development of full eld velocity mea-
surement techniques, as indicated by the number of publications in the last decade: Geveci et al. (2003),
Haigermoser (2009), Velikorodny et al. (2010) and Finnegan et al. (2010) just to cite a few. Morris (2010)
has recently provided a review on the use of PIV to examine how shear layer instabilities and turbulence lead
to radiated sound. The common strategy is to apply an acoustic analogy to the experimental data, either
the Curle (1955) acoustic analogy or the vortex sound theory introduced by Powell (1964) and extended
by Howe (1975), in order to identify the spatial distribution of the sound sources. Both analogies were
applied by Koschatzky et al. (2010) to estimate the cavity sound emission: the overall sound pressure level
was correctly predicted by the two methods even if the vortex sound theory appears to predict better the
amplitude of the tonal component.
Howe (1975, 1980) estimated that with the low Mach number, inviscid, constant entropy approximation,
the total ow velocity can be decomposed into an incompressible vorticity-bearing velocity and an irrotational
acoustic velocity. The generation of acoustic power by the vortical eld can be calculated by the following
formula:
=
0
_
V
( w v) u
acoust
dV (1.14)
where
0
is the uid density, v and w are the uid velocity and vorticity, and u
acoust
is the acoustic
particle velocity. The sign of determines if vorticity acts as a source or sink for the acoustics. For the
details on the derivation of equation 1.14 see Appendix A.
Interesting results have been found by applying equation 1.14 to strongly resonant cases: Velikorodny
et al. (2010), for example, estimated the distribution of the acoustic sources and sinks on a duct with coaxial
side branches (gure 1.12).
17
1.Background
Figure 1.12: Patterns of time-averaged acoustic power on a duct with coaxial side branches corresponding to the
second hydrodynamic oscillation mode (Velikorodny et al. (2010)).
18
Part I
Cylindrical cavity with open mouth
19
Chapter 2
Experimental set-up
2.1 Wind tunnel
The experimental investigation was conducted in a closed circuit low speed wind tunnel designed by the
Mechanical and Industrial Engineering Department (DIMI) of Roma Tre University. The facility is located
in the Italian National Agency for New Technology Energy and Environment (ENEA) research center of
Casaccia, 28 km from Rome. The closed test section is 2.49 m long (Lx) and has a 0.89 1.16 m
2
cross
section (Ly Lz). Other important dimensions of the wind tunnel are given in gure 2.1. The fan is able to
generate a ow ranging from 0 to 90 m/s in the centreline of the test section with a relative turbulence level
of 0.1% at a velocity of 40 m/s. Further details about the wind tunnel properties can be found in Camussi
et al. (2006a).
2.2 Test model
The test model was designed within the framework of the AeroTraNet project and manufactured by the Uni-
versity of Leicester. A Perspex cylindrical pipe with an interior diameter (D) of 210 mm was mounted ush
to the bottom wall of the test section, 1780 mm downstream the end of the convergent section (gure 2.2).
A at Perspex disk sealed the cylinder from underneath, creating a 285 mm deep (H) cavity (gure 2.3).
These dimensions lead to a cavity aspect ratio of 1.357 . According to the above mentioned literature studies
(section 1.1.2), the selected aspect ratio should guarantee the symmetry of the ow with respect to the
central streamwise wall-normal plane.
21
2.Experimental set-up
Figure 2.1: Schematic drawing of the ENEA Casaccia wind tunnel (top view).
(a) (b)
Figure 2.2: Schematic drawing of the test section with the cylindrical cavity studied. Dimensions of the test section
are: Lx = 2.49 m, Ly = 0.89 m, Lz = 1.16 m. An azimuthal angle has been introduced to follow the pressure at
the walls of the cavity.
22
2.Experimental set-up
Figure 2.3: Photography of the cavity before the preliminary characterization campaign. During the experiments,
one microphone was mounted in one of the orices. See gure 2.5 for further details on the set-up of the microphones.
2.3 Instrumentation
2.3.1 Pitot-static tube
The ow velocity in the test section was monitored with a Pitot tube connected to a Kavlico pressure
transducer model P592. The probe was introduced in the wind tunnel through the breather, a small slot
around the perimeter at the downstream end of the test section. The purpose of this vent is to keep the test
section close to atmospheric pressure and to isolate it from vibration that could otherwise be transmitted
through the diuser. The calibration of the pressure transducer was done with a pressure pump and a U-tube
manometer.
2.3.2 Hot wire anemometry
Velocity in the shear layer and in the wake of the cavity was measured with a 55P11 single component Dantec
probe connected to a constant temperature hot-wire anemometer (A.A. Lab System AN-1003). To reach the
desired positions on a given yz-plane, the probe was mounted on a two axis traverse system equipped with
stepping motors (Rexroth Compact Module CKK). The calibration of the probe was done with the Pitot
tube.
Because the hot wire anemometry is an intrusive technique, a verication that the presence of the hot-
wire probe does not aect the ow signicantly was performed. A microphone in the downstream cavity
23
2.Experimental set-up
wall was used as controller. A hot-wire probe was introduced in the test section and placed at 16 dierent
locations above the cavity. Figure 2.4 presents two pressure spectra corresponding to two dierent cases:
the hot-wire probe in the shear layer (y/H = 0) and outside the shear layer, far above the cavity (y/H =
0.35). Signicant dierences are not observed, leading to the conclusion that the presence of the hot-wire
probe in the shear layer does not strongly inuence the ow.
Figure 2.4: Auto-spectra of the pressure signal with a hot-wire probe in the shear layer (black line) and outside
the shear layer (grey line). The hot-wire probe was located in the center plane at 0.05D of the downstream edge.
U = 40 m/s.
2.3.3 Microphones
Two 1/4-inch Br uel&Kjr free-eld microphones (type: 4939 and 4135) were used to measure the uctuating
pressure at the cavity walls. The microphones were connected to pre-ampliers B&K 2670 and to a signal
conditioner Br uel&Kjr NEXUS 2692. Even if several holes for the microphones are available on the walls
of the cavity, a toothed graduated ring mechanism (precision of 1

) was designed in order to reach any


azimuthal angle by rotating the cavity (gure 2.3).
The microphones were connected to the interior of the cavity through 1 mm diameter pinholes. The
geometry of the pinholes is the same as that adopted in Camussi et al. (2006a,b, 2008). By using pinholes
the spatial averaging eects are minimized. The main drawback of this layout is the presence of a non
negligible volume between the microphone diaphragm and the pinhole. This volume can act as a Helmholtz
acoustic resonator. In the present case the corresponding cut-o frequency has been accurately calculated
to 3347 Hz with equation 1.9. The expression l
eq
= l + 1.64a was taken for the equivalent neck length
24
2.Experimental set-up
(a) (b)
Figure 2.5: Detail on the microphone set-up. The volume of air between the microphone diaphragm and the interior
surface of the cavity represents a Helmholtz resonator which dimensions are: length and cross surface area of the
pinhole l = 1.3 mm and S = 0.5
2
mm
2
respectively and volume of the resonator V = 99 mm
3
.
(Blake (1986)), where a represents the pinholes radius, which is a simplication of a double Rayleigh end
correction. The maximum frequency of interest (shear layer modes) is expected to be around 300 Hz for the
velocity range explored, thus the Helmholtz resonator should not aect the frequency range of interest.
2.3.4 Particle image velocimetry
To investigate the ow inside the cavity, Particle Image Velocimetry (PIV) was used. The planes of interest
were illuminated by a BigSky Twins Ultra/CFR 200 Nd:YAG laser with the maximum energy of 50 mJ
per pulse. The images were taken with a PCO Pixely VGA (1280 1024 pixels) camera equipped with a
Docter Optics Tevidon 1.8/16 lens. Synchronization between the laser and the camera was achieved with a
BNC digital pulse/delay generator 575 controlled with an in-house Labview program.
Velocity elds over horizontal streamwise planes (Oxz) were explored by introducing the laser light sheet
horizontally through the cavity sidewall and placing the camera under the cavity. The measurements were
performed at dierent depths by moving the laser along the y-axis. A picture and a schematic of the
experiment are given in gure 2.6.
The PIV processing was done with PIVDEF, a software developed by the Istituto Nazionale per Studi ed
Esperienze di Architettura Navale (INSEAN) also known as the Italian Ship Model Basin. This software
is based on a standard PIV cross-correlation algorithm (Cotroni et al. (2000)), a recursive window oset
(Westerweel et al. (1997)) and a multiplication between adjacent correlation tables (Hart (1998)). A window
deformation techniques similar to the one developed by Lecordier et al. (1999) is also used. Further details
on the processing algorithm are given in Di Florio et al. (2002).
For each plane of interest and for each ow velocity investigated, 600 couples of images were recorded.
The output velocity elds obtained with PIVDEF were exported to Matlab for further processing.
25
2.Experimental set-up
(a) (b)
Figure 2.6: (a) Photograph taken from the exterior of the test setion showing, from left to right, the laser
beam, a cylindrical and a spherical lens, the generated laser sheet, the cavity and the camera underneaths.
(b) Schematic of the experimental rig showing the position of the PIV apparatus.
2.3.5 Data acquisition card and processing of the pressure signals
Signals from all the instruments were acquired using a National Instrument SCXI-1600 Data Acquisition
Module. The signal from the hot wire was sampled at a frequency of 40 kHz, acquired for 5 seconds and
low-pass ltered at a cut-o frequency of 10 kHz to avoid aliasing. The pressure signals were acquired for 4
seconds using a sample rate of 25 kHz and were band-pass ltered (between 20 Hz and 10 kHz) by the signal
conditioner.
The pressure signals were processed in Matlab using a Fast Fourier Transform (FFT) technique. The
number of points per segment was 25000 and therefore the frequency resolution was 1 Hz. The 4 data sets
were then averaged.
2.4 Test conditions
2.4.1 Flow conditions
The wind tunnel velocity was set manually by rotating a knob which controls the speed of the fan: the RPM
can be then read on a digital display located in the control room. Before every measurement, enough time is
waited in order to obtain stable ow conditions. Upstream of the wind tunnels contraction, the temperature
was kept at 21

C 1

C by a heat exchanger. For the aeroacoustic characterization of the cavity, the ow


speed was increased from 5 to 55 m/s (M = 0.015 - 0.161) by steps of 1 m/s.
26
2.Experimental set-up
2.4.2 Incoming boundary layer
The boundary layer over the test section oor was characterized by hot-wire anemometry. The properties of
the incoming boundary layer at 46 mm from the upstream corner of the cavity are summarized in table 2.1 for
three velocities (20, 30 and 40 m/s). The length scale of the boundary layer (momentum thickness) represents
less than 2 % of the diameter of the cavity thus this parameter should not inuence the aerodynamics and
acoustics of the cavity. Velocity proles are given in gure 2.7 jointly with the 1/7 power proles. The
velocity measurements match the theoretical turbulent boundary layer trend.
U

(m/s) (mm)

(mm) (mm)
19.9 35.2 4.9 3.8
30.4 34.4 4.7 3.6
40.2 31.7 4.3 3.3
Table 2.1: Boundary layer properties for 3 free stream velocities (20, 30 and 40 m/s). Boundary layer thickness (99 %),
displacement thickness

and momentum thickness .


Figure 2.7: Velocity proles of the boundary layer measured at 0.22D from the upstream corner of the cavity at 3
dierent ow velocities: U = 20, 30 and 40 m/s.
2.4.3 Background pressure uctuations
The background pressure uctuations in the test section were explored in order to check if their intensity is
lower than the ow-excited cavity unsteady pressure level. Camussi et al. (2000) have already claried most
27
2.Experimental set-up
of the background unsteady pressure features of the ENEA Casaccia wind tunnel. Therefore the results from
this previous study were used to identify some of the frequency peaks in the pressure spectra.
The measurements were done by means of an in-ow microphone tted with a nose cone and located 150
mm (y) above the cavity. The mouth of the cavity was covered with a wooden plate. The pressure spectra
obtained for ow velocities between 5 and 55 m/s are given in gure 2.8 with the theoretical blade passing
frequency and its harmonics superimposed. The pure tone at 300 Hz, present for all the ow velocities, has
been ascribed by Camussi et al. (2000) to the constant speed fan installed for cooling the wind tunnel blower.
Figure 2.8: Background pressure uctuations in the test section without the cavity. The white dots represents the
blade passing frequency and its rst two harmonics.
2.5 Measurement matrix
The uctuating pressure was measured at the side and bottom walls of the cavity. A total of 325 dierent
positions were surveyed over 6 dierent depths (y), 4 dierent radii (r) and 36 dierent azimuthal angles ().
Figure 2.9 shows the positions adopted by the microphone. In this schematic drawing, the lateral wall has
been unrolled and the circular bottom reproduced underneath. With this plane representation, the reader
can have a panoramic view of the cavity walls. To achieve this measurement matrix, a single microphone
was moved from one pinhole to another and the cavity rotated by steps of 10 degrees. Results can be found
in section 4.4.
The aerodynamic campaign consisted in a three-dimensional grid of 2520 points distributed over 8 dif-
ferent streamwise (x), 7 dierent vertical (y) and 45 dierent spanwise (z) positions. A schematic drawing
indicating the hot wire probe positions adopted for the velocity measurements is given in gure 2.10: a top
28
2.Experimental set-up
and a lateral view of the experimental rig are represented. In section 4.1, the results from this campaign are
discussed.
For the PIV measurements, three dierent depths inside the cavity were explored: y/H = -0.25, -0.50
and -0.75. The obtained mean velocity elds are reported in section 4.2.
Figure 2.9: Schematic drawing of the wall pressure measurements grid, see gure 2.2 for the reference frame adopted.
The lateral wall has been unrolled and the circular bottom reproduced underneath.
29
2.Experimental set-up
(a)
(b)
Figure 2.10: Schematic drawing of the velocity measurements grid, see gure 2.2 for the reference frame adopted.
Positions taken by the hot wire probe for the aerodynamic campaign: top view (a) and lateral view (b)
30
Chapter 3
Acoustic mode calculation
3.1 The acoustic modes of an open-closed cylindrical cavity
The aim of this section is to nd an analytical expression for the natural acoustic modes of a cylindrical cavity
with an open mouth. The starting point is the three-dimensional wave equation that is the second-order
linear partial dierential equation governing the acoustic pressure p:

2
p =
1
c
2

2
t
2
p (3.1)
where c is the speed of sound. The pressure can be decomposed into a sum of orthogonal Fourier
components p
h
e
i
h
t
which individually satisfy the wave equation at every moment t. Therefore equation 3.1
becomes the Helmholtz equation:

2
p
h
+k
2
h
p
h
= 0 (3.2)
where k
h
=
h
/c the wave number of the corresponding Fourier component.
In cylindrical coordinates (r, and y, see gure 3.1) the Helmholtz equation becomes:

2
p
h
r
2
+
1
r
p
h
r
+
1
r
2

2
p
h

2
+

2
p
h
y
2
+k
2
h
p
h
= 0 (3.3)
This equation can be solved analytically by the separation of variables technique (details are not given
here for sake of brevity and because it can be found in many acoustic textbooks). The general solution of
equation 3.3 takes the following form:
p
h
(r, , y) = J
m
(k
r
r) e
im
(C
1
e
ikyy
+C
2
e
ikyy
) (3.4)
where J
m
is the Bessel function of the rst kind of order m, C
1
and C
2
two constants.
31
3.Acoustic mode calculation
Figure 3.1: Denition of the cylindrical coordinates for the analytical resolution of the Helmholtz equation.
In order to nd the acoustic eigenfrequencies of the cavity, the boundary conditions have to be specied:
the wall-normal derivative of p has to vanish at the walls (bottom and side) and an assumption has to be
made for the mouth of the cavity (y = H). As mentioned in section 1.1.2, physically the pressure node is not
exactly at the opening of the cavity. A simple way to proceed is to take an end correction: the additional
boundary condition adopted here is that the pressure has to vanish at y = H +R. The general expression
for an eigenfrequency is:
f
m,n,q
=

2
=
c
2
_
_
j

m,n
R
_
2
+
_
(2q

+ 1)
2(H +R
_
2
_
1/2
(3.5)
where j

m,n
is (n + 1)
th
positive zero of J

m
and m, n and q are the azimuthal, radial and longitudinal
order of the mode and q = 2q

+ 1 where (m, n, q

) N
3
. It is interesting to notice that if m = n = 0,
equation 3.5 reduces to equation 1.8.
With the dimensions of the present cavity (H and D) and the temperature and the ambient pressure
during the tests (for the calculation of c), the frequencies of the acoustic modes of the cavity can be calculated.
The end correction coecient that depends on k
y
can be evaluated with equation 1.7 (gure 3.2). The
predicted frequencies are given in table 3.1 and the shapes of the rst eight modes are reported in gure 3.3.
Mode /4 3/4 AZ1(/4) AZ1(3/4) 5/4 AZ1(5/4) AZ1R1(/4) AZ1R1(3/4)
(m, n, q) (0,0,1) (0,0,3) (1,0,1) (1,0,3) (0,0,5) (1,0,5) (1,1,1) (1,1,3)
Frequency 236 760 994 1246 1333 1663 2792 2904
Table 3.1: Frequencies, in hertz, of the acoustic modes of the cavity. Mode order (m, n, q): azimuthal, radial, longitudinal.
32
3.Acoustic mode calculation
Figure 3.2: Correction coecient predicted by equation 1.7. The red cross represent the quarter wavelength ( =
0.7536), the black dot the mode 3/4 ( = 0.5153) and the green star the mode 5/4 ( = 0.3562).
3.2 The acoustic modes in a wind tunnel
When performing aeroacoustic experiments in a closed test section, it is important to study the acoustic
behaviour of the facility in addition to its aerodynamic performance. If the walls of the test section are not
acoustically treated, the pressure measurements may be aected by the acoustic resonant modes of the wind
tunnel. Furthermore the presence of a model in the test section can modify the acoustic resonances of the
wind tunnel and, more important, excite strongly localized resonant modes with zero radiation loss called
trapped modes.
In order to predict the frequency and the shape of the resonant acoustic modes of an experimental rig,
nite element analysis has become popular in the past decade: Aly & Ziada (2010) used the commercial
software ABAQUS to study the trapped modes of a ducted axisymmetric internal cavity, Finnegan et al.
(2010) calculated the acoustic eld around a blu body with ANSYS and Oshkai et al. (2008) predicted
the acoustic modes of a coaxial side branch resonator with COMSOL. Some advantages of the use of a
commercial software for the calculation of the acoustic modes are the promptness and the reliability of the
results when the boundary conditions are properly set.
3.3 The computational model
The acoustic modes of the test rig were studied using the commercial software COMSOL Multiphysics
4.0. This package includes a nite element solver able to evaluate partial dierential equations of dierent
disciplines. Even if several physical models can be solved jointly, a no-ow case was studied. The default
eigenfrequency formulation available within this software was used for the calculation:

p
_
+

2
p
c
2
= 0 (3.6)
33
3.Acoustic mode calculation
(a) (0,0,1) (b) (0,0,3)
(c) (1,0,1) (d) (1,0,3)
(e) (0,0,5) (f) (1,0,5)
(g) (1,1,1) (h) (1,1,3)
Figure 3.3: Shape of the rst height acoustic modes of a cylindrical cavity with a closed bottom wall and an open
top surface. The innite ange has not been represented. The color bar gives the normalized real pressure values.
See table 3.1 for the frequency of each mode (m, n, q).
34
3.Acoustic mode calculation
(a) (b)
Figure 3.4: Overview (a) and detail (b) of the mesh. The spatial domain chosen for the calculation includes the
contraction and the test section of the wind tunnel.
where is the air density, c the speed of sound, p the pressure and an eigenvalue. From this equation, it
is clear that the uid is modeled as a lossless medium. Once the equation has been solved, the eigenfrequency
can be evaluated through:
= i2f (3.7)
3.4 The geometry
A preliminary investigation brought into evidence that the simulation of the test section without the con-
vergent nozzle does not predict correctly the frequencies measured experimentally. Thereby a simplied
contraction was included into the calculation leading to a ve-meter-long geometry.
As already mentioned in the previous chapter, the frequency range of interest is [0 - 300 Hz]. In order to
resolve correctly the highest frequency of this range, every mesh element does not have to exceed 0.2 meters.
This is dictated by the ve-mesh-elements-per-wavelength rule. The wind tunnel section was meshed with
approximately 53 000 tetrahedrons and the cavity with 6 000 elements. Figure 3.4 gives an idea of the
meshed geometry.
All the walls (contraction, test section and cavity) were modelled with rigid boundaries (called sound-
hard in the COMSOL Multiphysics) meaning that the normal derivative of the pressure is set to zero. At
the end of the contraction and at the end of the test section soft boundary conditions were imposed: the
acoustic pressure vanishes.
35
3.Acoustic mode calculation
Without cavity With cavity
Frequency Mode number Mode number Frequency Dierence
Hz Hz %
42.64 1 1 42.63 0.02
66.72 2 2 66.67 0.07
99.95 3 3 99.72 0.23
102.42 4 4 102.42 0.00
118.07 5 5 118.07 0.00
130.86 6 6 130.62 0.18
136.11 7 7 136.11 0.00
152.05 8 8 152.06 -0.01
154.98 9 9 154.81 0.11
160.25 10 10 160.28 -0.02
165.67 11 11 165.47 0.12
167.84 12 12 167.84 0.00
181.13 13 13 181.19 -0.03
183.76 14 14 183.76 0.00
186.95 15 15 186.87 0.04
191.16 16 16 191.16 0.00
196.16 17 17 196.16 0.00
198.32 18 19 198.66 -0.17
199.83 19 18 197.78 1.03
204.28 20 20 204.3 -0.01
212.2 21 21 211.96 0.11
215.95 22 22 212.78 1.47
217.59 23 23 217.61 -0.01
223.86 24 24 223.86 0.00
225.32 25 26 225.32 0.00
226.96 26 25 224.6 1.04
232.04 27 27 232.06 -0.01
232.52 28 28 232.53 0.00
241.29 29 29 236.77 1.87
244.71 30 30 244.72 0.00
do not exist 31 246.76
249.47 31 32 249.47 0.00
250.08 32 33 250.13 -0.02
251.21 33 34 251.29 -0.03
253.97 34 35 256.64 -1.05
258.69 35 36 258.69 0.00
260.52 36 37 260.52 0.00
262.13 37 38 263.08 -0.36
Table 3.2: Frequency of the acoustic modes of the wind-tunnel with (right) and without (left) cavity.
36
3.Acoustic mode calculation
3.5 Results
3.5.1 Acoustic modes without the cavity
In the range of frequency [0 - 263 Hz], 37 dierent eigenmodes were found and their frequencies, sorted from
smallest to largest, are reported in table 3.2. Six modes were identied with a zero isobar central horizontal
plane but without a zero isobar central vertical plane. The frequencies of these modes are highlighted in
the table 3.2 and their pressure distributions over the walls of the test section are plotted in gure 3.5. The
shapes of six modes not having the characteristics previously described are given in gure 3.6.
3.5.2 Acoustic modes with the cavity
When the cavity is considered, 38 dierent eigenmodes are found in the same frequency range. The eigenfre-
quencies are reported in table 3.2 jointly with the corresponding eigenfrequencies without the cavity. Two
minor changes can be noticed: small frequency shifts and two order permutations (modes 18-19 and 25-26).
Furthermore, the addition of the cavity into the geometry brings one mayor change: another eigenfrequency
appears. The discussion of this extra mode will be detailed later.
As in the previous case, the modes can be divided in two categories according to the pressure pattern on
the horizontal and vertical planes of the test section. The second group (31 eigenmodes) is weekly inuenced
by the introduction of the cavity: the biggest frequency shift is 0.23 %. When comparing gure 3.6 to
gure 3.8 it can be seen that the shapes of the modes have not changed signicantly. Another observation
is that the pressure inside the cavity is constant and equal to zero. It is important to point out that most
of the modes of this group have a central vertical plane where the pressure is equal to zero (pressure node).
The shape of these modes would has probably changed if the cavity had been added in a plane other than
the central one.
Figures 3.7 gives the shapes of the 6 acoustic modes for which the central horizontal plane (0xz) was
a pressure node and the central vertical plane was not a zero isobar before the introduction of the cavity.
The frequency shifts are reported in table 3.2: the maximum dierence is of 1.87 % which is a much higher
percentage than for the former category. It can be seen, in gures 3.7, that the pressure amplitude in the
cavity is not constant. All the modes of this class have a pressure pattern similar to the cavity quarter-wave
length: the bottom of the cavity is a pressure anti-node and the uctuations at the orice are minimal. This
common shape can be illustrated by the mode 35 (gure 3.10). At the cavity orice, the pressure pattern is
not uniform as it should be for an unconned case: this mode, as the 6 others, are indeed a combination of
a cross-ducted mode with the quarter-wave length of the cavity.
The only eigenfrequency added by the cavity is 246.76 Hz (pressure reported in gure 3.9). This value
is close to 326 Hz, the frequency of the quarter-wave length of the cavity analytically calculated (table 3.1).
This result is relevant for the present analysis since the signature of this mode will be evidenced by the wall
pressure measurements presented in section 4.4 of the next chapter.
37
3.Acoustic mode calculation
(a) Mode 19 (b) Mode 22
(c) Mode 26 (d) Mode 29
(e) Mode 34 (f) Mode 37
Figure 3.5: Pressure distribution (Pa) on the walls for six dierent eigenmodes (without cavity).
38
3.Acoustic mode calculation
(a) Mode 25 (b) Mode 27
(c) Mode 30 (d) Mode 31
(e) Mode 32 (f) Mode 35
Figure 3.6: Pressure distribution (Pa) on the walls for six dierent eigenmodes (without cavity).
39
3.Acoustic mode calculation
(a) Mode 18 (b) Mode 22
(c) Mode 25 (d) Mode 29
(e) Mode 35 (f) Mode 38
Figure 3.7: Pressure distribution (Pa) on the walls for six dierent eigenmodes (with cavity).
40
3.Acoustic mode calculation
(a) Mode 26 (b) Mode 27
(c) Mode 30 (d) Mode 32
(e) Mode 33 (f) Mode 36
Figure 3.8: Pressure distribution (Pa) on the walls for six dierent eigenmodes (with cavity).
41
3.Acoustic mode calculation
Figure 3.9: Extra mode: acoustic mode 31
Figure 3.10: Isosurface plot of the acoustic pressure distribution (Pa) inside the cavity at 256Hz. This mode has a
quarter-wave length shape
42
Chapter 4
Experimental results
4.1 Overall aerodynamics
A single component hot-wire probe was used to measure the velocity above and downstream of the cavity
in order to obtain a general understanding of the shear layer and of the wake topology. The incoming free
stream velocity was xed to U

= 40 m/s. The ow velocity at a given position (x, y, z) in the test section


is

U =

U
x
+

U
y
+

U
z
. As the wire was orientated along the z direction, the velocity measured by the hot-wire
is U
hw
= ||

U
x
+

U
y
||. Proles of the non-dimensional mean velocity (U
hw
/U

) and of the non-dimensional


standard deviation of the measured velocity (
hw
/U

) are given in gure 4.1 and gure 4.2 respectively.


According to the literature, because H/D > 0.8 both quantities are expected to be symmetric with respect to
the central streamwise plane (Oxy) within the range of experimental uncertainties. The shear layer topology
is rst presented followed by the wake description.
4.1.1 Shear layer topology
Three dierent streamwise planes are reported in gure 4.1: x/D = 0, 0.25 and 0.50. For each one of these
planes, dierent heights were explored. In the region close to the vertical symmetry plane (z/D = 0) a mean
velocity defect is clearly observed as an eect of the shear layer generated over the cavity opening. Because
the geometry is cylindrical, the greatest shear layer eects are observed in the central region. The turbulent
velocity uctuations (
hw
) conrm this trend since an increase of the turbulent kinetic energy is observed in
the region close to the plane z/D = 0. When moving downstream, the shear layer grows: the mean velocity
decreases and the velocity uctuations increase. This eect is shown both for the cases at x/D = 0.25 and
0.5.
It is interesting to note that when a given longitudinal position x/D is considered, the perturbation
induced by the presence of the cavity is restricted to the wall proximity region. Indeed, in the region close
to the downstream side of the cavity and for around y/D = 1/10, U
hw
and
hw
proles are almost at and
43
4.Experimental results
the signature of the cavity shear layer is no longer observed.
Closer to the wall, further interesting features are detected as an eect of the interaction between the
shear layer and the downstream edge of the cavity. In the plane x/D = 0.5, wiggles at about z/D = 0.25
and z/D = 0.25 are present in the y/D = 0.015 and y/D = 0.025 proles. These localized bumps are present
both in the mean velocity and in the root-mean-square velocity uctuations proles at the same positions.
According to surface oil-lm pattern visualisations presented by Gaudet & Winter (1973), these bumps can
be ascribed to the eect of the reattachment of the shear layer in the downstream cavity wall region. The
shear layer impacting the wall indeed leads to a localized variation of the mean velocity magnitude and,
consequently, to a relevant increase of the velocity uctuations.
4.1.2 Wake topology
The presence of wiggles is also detected in the velocity and turbulence proles in the wake region (gure 4.2).
According to the physical features described above, the signature of the shear layer is still present in the
vicinity of the symmetry plane even though for x/D > 0.65 its eect declines rapidly. Similar results have
been obtained in a rectangular deep cavity by Ukeiley & Murray (2005) as an eect of the reattachment of
the shear layer on the downstream lip. A decrease of the turbulent velocity uctuations is indeed observed in
the wake near the wall, while for larger distances from the wall, the turbulence level increases as an eect of
the wake development. The amplitude of the velocity defect and the corresponding increase of the turbulence
level gradually disappear when moving downstream from the cavity, thus the eect of wiggles on the velocity
proles becomes more evident in the far region. With respect to the velocity bumps observed in the shear
layer region above the cavity, the wiggles are now observed at spanwise positions of about z/D = 0.5 and
z/D = 0.5, and for larger distances from the wall, up to y/D 0.1 for x/D = 2.5. These results seem to
suggest that the trend observed in the wake region is due to physical mechanisms dierent from the one
observed in the region above the cavity. With the support of the Gaudet & Winter (1973) visualisations,
an interpretation can be addressed: the observed wiggles can be ascribed to the perturbation caused by
two vortical structures convected from the cavity downstream lip, having streamwise vorticity and detaching
symmetrically with respect to the vertical plane at z = 0, from the cavity wall. The signature of the two
vortices is indeed not present in the region above the cavity thus suggesting that they are formed downstream
of the cavity opening. With the current measurements it is impossible to draw denite conclusions about
the topology of the tip vortices. However, preliminary results from a numerical simulation of a cylindrical
cavity (Grottadaurea & Rona (2008)), conrm the existence of longitudinal tip vortices.
4.2 Description of some ow features inside the cavity
Before discussing the ow inside a cylindrical cavity, it is important to describe rst the well documented
ow inside a rectangular cavity. For a deep rectangular cavity, the ow detaches at the upstream edge
44
4.Experimental results
(a) (b)
(c) (d)
(e) (f)
Figure 4.1: Spanwise proles of non-dimensional mean velocity U
hw
/U (left) and of the non-dimensional root-
mean-square velocity uctuations
hw
/U (right) in the shear layer at three dierent streamwise positions (x/D =
0, 0.25 and 0.5) and for dierent heights (legend in the left column).
45
4.Experimental results
(a) (b)
(c) (d)
(e) (f)
Figure 4.2: Spanwise proles of non-dimensional mean velocity U
hw
/U (left) and of the non-dimensional root-
mean-square velocity uctuations
hw
/U (right) in the wake at three dierent streamwise positions (x/D = 0.64,
1.5 and 2.5) and for dierent heights (legend in the left column).
46
4.Experimental results
creating a shear layer over the mouth of the cavity. This shear layer hits the downstream-wall and one part
of the ow turns downwards into the cavity. This turned down jet behaves like a decelerating stagnation
ow while approaching the cavity the bottom. Afterwards the ow mainly recirculates inside the enclosure.
This description is also valid in the symmetric plane (z/D = 0) of a deep cylindrical cavity (Mincu et al.
(2009) and Grottadaurea & Rona (2008)). However, because for circular based cavity the ow is strongly
3-dimensional, the knowledge of the ow in a single plane gives only a local understanding and not a global
description of the ow.
Results from PIV measurements taken at three dierent horizontal planes are analysed in this section.
At each plane, 600 couples of images were captured and a mean velocity eld was calculated (gure 4.3).
The velocity eld in an upper plane inside the cavity (y/H = 0.25) is presented in gure 4.3(a). Near
the downstream-wall, the turned down jet gives rise to two vertical counter-rotating vortices (x = 0.07 m,
gure 4.3(a)). The main reasons for the generation of these vortices are the curvature of the wall and the
opposite spanwise velocity components at each side of the cavity. The negative streamwise component of
the velocity indicates that the jet detaches from the vertical wall as it goes down. By analyzing the plane
at half depth (y/H = 0.5, gure 4.3(b)), one can notice that the jet has spread and that the vortices have
moved apart from each other.
As expected for a deeper plane, at y/H = 0.75 the uid was found to ow in an opposite direction
than for an shallower plane: on the downstream half part of the cavity (x/D > 0), the velocity vectors are
orientated upstream (gure 4.3(c)). For the half plane x/D < 0 however, velocity eld is governed by the
sink point located at x = 0.04 m over the x-axis. These 2D PIV measurements do not give information
about the vertical velocity component (U
y
). However it is easy to imagine that the sink point corresponds
to the center of the turn up jet over the upstream wall.
4.3 Unsteady response to a grazing ow
4.3.1 Pressure response
In order to characterize the cavity pressure response to a turbulent boundary layer, dierent ow velocities
were investigated. The pressure spectra at three dierent ow velocities are reported in gure 4.4: a case
with no relevant resonance (30 m/s), a case exhibiting two spectral peaks of almost the same intensity (41
m/s) and a strong resonating case (53 m/s). When the velocity increases, the pressure uctuations at the
walls also increase by several orders of magnitude, thus justifying the use of a logarithmic scale. As already
mentioned in chapter 2, the blowers cooling system generates a tone at 300 Hz that can be noticed in all
the spectra. For comparison, the spectra of the background pressure uctuations in the test section are also
reported in gure 4.4 (see section 2.4.3 for experimental details).
In order to have a global view of the pressure response for velocities ranging from 4 to 56 m/s, a total of 53
spectra are presented together in gure 4.5. Three dierent spectral peaks can be seen, evolving in frequency
47
4.Experimental results
(a)
(b)
(c)
Figure 4.3: Mean velocity eld for 3 dierent planes: (a) y/H = 0.25; (b) y/H = 0.50; (c) y/H = 0.75
for a free stream velocity in the test section of 10 m/s.
48
4.Experimental results
Figure 4.4: Pressure uctuations measured with a microphone mounted on the downstream wall of the cavity
( = 180

, y/H = -0.1875) and with a in-ow microphone (cavitys mouth covered) for dierent velocities: U =
30, 41 and 53 m/s.
and in amplitude with the wind tunnel velocity: they are the signature of the shear layer hydrodynamic
modes. The agreement with the Rossiter (1964) prediction is very good. For the present data, the values
showing the best agreement with the Rossiter semi-empirical equation are = 0 and = 0.53 as can be
seen in gure 4.5. It should be noted that the characteristic length chosen was the diameter of the cavity
because it ts better the data compared the other expressions as those used by Bruggeman et al. (1991) and
Czech et al. (2006).
The shear layer modes follow a linear trend for velocities up to 30 m/s. For higher ow speeds, a stepwise
evolution characterizes the progression of the modes: the second and third hydrodynamic modes start to
increase by steps after reaching 197 Hz (gure 4.5). This phenomenon can be ascribed to the fact that the
cavity model is installed in a closed test section, an environment propitious to the establishment of a ow-
acoustic coupling. The shear layer instabilities are responsible for the initial excitation of the acoustic modes
of the wind-tunnel/cavity system. The acoustic eigenmodes of the system dictate the lock-on frequencies. As
can be seen in gure 4.5, lock-on is generated only at 6 frequencies. When comparing to the nite element
study (chapter 3), these lock-on frequencies correspond to acoustic modes that have a quarter-wave length
shape. A stepwise evolution of the shear layer modes was also observed by Czech et al. (2006) for an array of
cylindrical cavities in a closed section. As a matter of fact, Marsden et al. (2008) performed measurements
in a cavity installed in an open test section and reported a linear evolution of the shear layer modes.
The coupling between a hydrodynamic mode and an acoustic mode not only generates lock-on but also
49
4.Experimental results
(a)
(b)
Figure 4.5: Three-dimensional and plan view of the pressure amplitude on a logarithmic scale (dB) as a function of
the ow speed and the frequency. The microphone was located at: ( = 90

, y/H = -0.25). Solid lines represent the


Rossiter (1964) prediction of the shear layer modes with [, ] =[0 , 0.53 ]. The frequencies of the acoustic modes
calculated with the FEM simulation are also plotted (dotted and dashed lines and diamonds). The dash lines are the
eigenvalues having a quarter-wave length shape pressure distribution inside the cavity. The depth mode (diamonds)
has a frequency of 246 Hz.
50
4.Experimental results
amplication of the pressure oscillations. At 53 m/s, for example, the shear layer mode SL2 is amplied
because coupled with an acoustic mode of the wind-tunnel/cavity: the peak in the pressure spectrum raises
40 dB over the background noise levels (gure 4.5). This should be compared to a non-resonating state, at
30 m/s for example, where the peaks corresponding to hydrodynamic modes are only 10 dB higher than the
overall sound pressure level.
4.3.2 Nondimensionalization process
The pressure spectra are now presented in non-dimensional form in gure 4.6. The Strouhal number based on
the cavity diameter is the selected dimensionless frequency. It has been shown previously that the pressure
uctuations increase with the ow velocity: in order to take into account the hydrodynamic eect on the wall
measurement, the pressure is normalized with respect to the dynamic pressure 0.5U
2

. The dimensionless
pressure spectra are presented on a logarithmic scale. For Strouhal numbers up to 0.8, a satisfactory collapse
of the spectra is observed. The signature of the rst hydrodynamic shear mode is a bump in the spectra at
St = 0.57. In contrast, the signature of the second hydrodynamic mode (St = 1.00) is either a bump (U

= 15, 22 and 32 m/s) or a sharp peak (U

= 42 and 52 m/s) depending whether or not is it coupled with


an acoustic modes. For Strouhal numbers higher than 1.1, a collapse of the spectra is no longer observed.
Figure 4.6: Pressure spectra for dierent velocities: U = 15, 22, 32, 42 and 52 m/s. The frequency and the pressure
intensity have been normalized.
For the sake of completeness, the pressure response for velocities between 4 and 56 m/s, is reported in
the form of plane contour plot with non dimensional frequencies in gure 4.7. The frequencies of the three
Rossiter modes and the 246 Hz mode predicted with the FE simulation are also plotted. This quarter wave
51
4.Experimental results
length-like mode has the propriety to be the only trapped mode generated by the cavity itself.
Figure 4.7: Plan view of dimensionless pressure amplitude on a logarithmic scale as a function of ow speed and the
Strouhal number based on the diameter of the cavity. Lines represent the Rossiter (1964) prediction and diamonds
the 246 Hz mode found with the FEM calculation.
4.3.3 Velocity response
A similar study was conducted from the velocity measurements in the shear layer. The velocity spectra
at dierent ow speed are reported in gure 4.8. The plane contour plot representation is used and the
velocity amplitudes are normalized by the ow speed. This gure is similar to the pressure response given
in gure 4.5: the three rst shear layer hydrodynamic modes are identiable as well as their lock-on with
the test section/cavity acoustic modes.
4.4 Spectral decomposition
4.4.1 Spectral decomposition and analysis on the symmetry plane
In order to extend the spectral pressure study, several measurement positions have been analysed. Figure 4.9
shows the curvilinear coordinate s adopted in order to locate the 19 pinholes on the vertical central plane of
the cavity. The pressure spectra at three dierent points (s = 0.93, 0.5 and 0.07) are presented in gure 4.10
for a ow speed of 40 m/s. The signature of the three shear layer modes can be identied (peaks at f
= 100, 195, 256 Hz). When considering the spectra quantitatively, pressure amplitudes are higher on the
downstream wall (s = 0.93) than on the upstream wall (s = 0.07) for all the frequencies except for f = 256
52
4.Experimental results
Figure 4.8: Plane view of the dimensionless velocity amplitude on a logarithmic scale as a function of ow speed
and frequency. The hot-wire was located in the shear layer region near the downstream edge (x/D = 0.45, y/H =
0.14, z/D = 0).
Hz. For this specic frequency, the pressure uctuations on the downstream wall are as important as the
pressure uctuations on the upstream wall. When the microphone is located at the bottom of the cavity (s
= 0.5), the spectrum is entirely dominated by the 256 Hz peak.
An original signal post-processing has been applied to the pressure data in order to explain the variations
of spectral content from one position to another. The idea is to evaluate the importance of some frequencies
to the overall uctuating pressure level which is dened as:
OAFPL = 10 log
_
_
_
_
_
+
_
f=0
W
p
(f) df
P
2
ref
_
_
_
_
_
(4.1)
where W
p
is the power spectral density of the wall pressure and P
ref
is the reference sound pressure in
air (20 Pa) . The uctuating pressure level of the frequency f can be estimated by:
FPL
f
= 10 log
_
_
_
_
_
_
f+10Hz
_
f

=f10Hz
W
p
(f

) df

P
2
ref
_
_
_
_
_
_
(4.2)
By this procedure, four dierent quantities have been calculated: the FPL
30Hz
from the low frequencies
and FPL
100Hz
, FPL
195Hz
and FPL
256Hz
from the shear layer hydrodynamic modes (SL1, SL2 and SL3).
53
4.Experimental results
Figure 4.9: Schematic drawing of the cavitys middle plane with the curvilinear coordinate s adopted. The positions
of the pressure transducers are represented with red circles. The recirculation pattern present inside a deep cavity is
also shown: the shear layer impacts on the downstream wall (s 1), then a near-wall ow moves towards the bottom
and impacts the bottom wall of the cavity (s 0.6).
Figure 4.10: PSD measured at three dierent positions (s = 0.07, 0.50 and 0.93) along the cavity wall for a ow
velocity of 40 m/s. The rst three hydrodynamic modes are annotated.
54
4.Experimental results
(a) (b)
Figure 4.11: (a) Power spectral density of the pressure uctuations at the bottom wall (s = 0.5). The four stacked
areas represent the integrals in FPL30Hz, FPL100Hz, FPL195Hz and FPL256Hz. (b) Fluctuating pressure levels of some
frequencies (LF, SL1, SL2 and SL3) along the curvilinear coordinate s for 40 m/s.
The selected frequencies are shown in gure 4.11(a) and the related uctuating pressure levels are given in
gure 4.11(b). When considering the low frequency uctuations, one can notice that FPL
30Hz
evolves as
the ow progresses inside the cavity: it is maximal where the shear layer hits the downstream wall (near
s = 1) and decreases inside the cavity as a consequent loss of energy of the internal recirculating ow. The
detachment of the turned down jet from the downstream wall and the second impingement on the bottom
of the cavity (at s = 0.6) can be detected with FPL
30Hz
. Such a ow organization has been observed,
for example, in the Detached Eddy Simulation (DES) of Grottadaurea & Rona (2008), where streamlines
indicate a detachment of the turned down jet and an impact on the bottom wall. The mean pressure
measurements of Hering et al. (2006) for a circular cavity (H/D = 0.7, U

= 27 m/s), also indicate that the


maximum pressure region on the bottom is not exactly in the corner but more towards the bottom center
leading to a recirculation region in the downstream corner of the cavity. The FPL
100Hz
trend is similar
to the FPL
30Hz
. The explanation for that is given by the dimensionless pressure study (gure 4.6): for
low frequencies (St < 0.7), the pressure spectra collapse because there is no strong coupling between the
uid-dynamics and the acoustics inside the cavity.
A very dierent evolution is found for FPL
256Hz
: its maximum is reached on the cavity bottom and its
minimum at the cavity mouth. This trend is related to an acoustic mode having a pressure anti-node at the
cavity bottom and a node at its opening. The velocity presented (40 m/s) corresponds indeed to a case in
which the third shear layer mode excites a trapped mode (f = 256 Hz) which shape is similar to the quarter
wavelength mode of the cavity.
55
4.Experimental results
4.4.2 Analysis on the cavity walls
The uctuating pressure levels are further analyzed by considering their azimuthal variations. The quantities
OAFPL, FPL
256Hz
and FPL
30Hz
over the side wall and the bottom of the cavity are given in gure 4.12.
When comparing the spatial distribution of OAFPL and FPL
256Hz
, it appears that the acoustic eects
related to the 256 Hz mode dominate the wall pressure uctuating almost everywhere inside the cavity. The
pressure level of the low frequencies gives an indication of the ow proprieties inside the cavity. In order
to facilitate the description of the ow, the side wall was divided in ve dierent regions based on FPL
30Hz
amplitude and on Gaudet & Winter (1973) observations. Arrows representing the ow direction were also
included (gure 4.12(c)). High values of FPL
30Hz
on the downstream wall (A) and on the bottom wall
correspond for example to impingement areas. The interface between a region where the air ows towards
the cavity bottom (B), and a region where it ows towards the top of the cavity (D) is characterized by
medium FPL
30Hz
values (C).
56
4.Experimental results
(a) (b)
(c)
Figure 4.12: Fluctuating pressure levels over the internal walls of the cavity (side and the bottom) for a ow speed
of 40 m/s. (a) overall uctuating pressure levels OAFPL, (b) uctuating pressure level of the third shear layer mode
FPL256Hz, and (c) uctuating pressure level of the low frequencies FPL30Hz. The arrows represent the ow velocity
direction in the dierent regions (A, B, C, D) of the side wall.
57
4.Experimental results
58
Chapter 5
Conclusion
An open mouth cylindrical cavity was characterized experimentally by means of velocity and wall-pressure
measurements.
Velocity and turbulence proles revealed that the ow is symmetric as expected. Through the analysis
of the cavitys wake, some three-dimensional eects were identied: the signature of a vortex pair detaching
from the cavity downstream lip and the shear layer reattachment on the downstream wall.
Concerning the ow inside the cavity, PIV measurements gave important insight into the three-dimensional
characteristics of the mean ow. The turned down jet was found to generate two vertical symmetric vortices
near the downstream wall.
The spectral analysis of both the wall-pressure and velocity signals revealed the presence of three shear
layer hydrodynamic modes. The frequencies of these modes are correctly predicted by the classical empirical
formulation for rectangular cavities. A ow-acoustic coupling has been found to occur for some specic
ow velocities. A nite element simulation of the experimental rig showed that the lock-on frequencies
correspond to some acoustic modes of the test section/cavity. These acoustic resonances have a common
pressure distribution inside the cavity: a quarter-wave length shape.
Through a processing of the wall pressure uctuations, which calculates the contribution of some specic
frequencies to the overall uctuating pressure level, some important ow features were claried. After the
impingement of the shear layer on the downstream wall, a fraction of the uid enters into the cavity creating
a turn down jet along the downstream sidewall. Before it impinges on the bottom, the wall-jet moves away
from the sidewall.
59
5.Conclusion
60
Part II
Cylindrical cavity with partially
closed mouth
61
Chapter 6
Experimental set-up
6.1 Overview of the experimental rig
Experiments were performed in the Fluids Laboratory of the Mechanical and Manufacturing Engineering
Department at the Trinity College in Dublin. The rig consists of a 493 mm (H) deep cylindrical cavity with
a 238 mm internal diameter (D) mounted on the lateral wall of a 335 mm long test section with a 125 mm
125 mm cross section (W W). A section view of the experiment is given in gure 6.1. A rectangular
orice connects the interior of the cavity to the test section of the elliptical bell-mouth inlet draw-down wind
tunnel. More details about the opening are given in section 6.3. The ow was generated by a centrifugal
blower driven by a motor. A picture of the Perspex cavity is given in gure 6.2. Additionally, a schematic
of the wind tunnel (gure 6.3) gives an overview of the experimental rig.
6.2 Design of the experimental rig
As the available facility for the testing was a small low speed wind tunnel, a preliminary analysis was required
in order to optimally design the experimental rig. One of the goals of this research is to demonstrate that
acoustic resonant modes other than longitudinal modes can be excited in deep cylindrical cavities. Therefore
the cavity had to be designed in such a way that the frequencies of the desired acoustic modes match the
frequencies of one of the shear layer hydrodynamic modes within the velocity range of the wind tunnel.
The geometrical parameters that could be chosen were the length of the opening L and the dimensions
of the cavity (diameter and depth). The constraints were mainly the dimensions of the test section (335 mm
125 mm 125 mm) and the ow speed range (0 - 53 m/s). The shear layer hydrodynamic modes were
estimated with the Rossiter (1964)s formula (equation 1.1) and only the rst three orders (n = 1, 2 and 3)
were taken into account. The frequencies of the acoustic modes were computed analytically: see section 7.1
for more details.
63
6.Experimental set-up
Figure 6.1: Schematic of the wind tunnel test section and the cavity. Section taken in the central plane. Only one
microphone is represented for simplicity. Not to scale.
Figure 6.2: Photograph of the experimental rig. One can see the digital camera for the PIV measurements and two
rings of eight microphones each.
64
6.Experimental set-up
Figure 6.3: Schematic of the wind tunnel, adapted from Finnegan (2011).
Figure 6.4: Design of the experimental rig. The depth of the cavity was set to H = 492.5 mm. Coloured curves
represent the rst three shear layer hydrodynamic modes (red I, blue II and green III): the thickness of the curves
is proportional to the opening length L: 10 mm (thin), 40 mm (normal) and 70 mm (thick). The horizontal lines
represent the acoustic modes of the cavity (H1, AZ1 and AZ1H1): the thickness of the lines is proportional to the
diameter of the cavity D: 150 mm (thin), 200 mm (normal) and 250 mm (thick).
65
6.Experimental set-up
According to Rossiter (1964)s formula, the frequencies of the hydrodynamic modes do not depend on
D nor on H; they depend only on the length of the opening (L). Secondly, the frequencies of the acoustic
modes do not depend on L; they depend only on the dimensions of the cavity: the bigger the cavity is,
the lower the modal frequencies are. These two considerations are the main concepts for the design of the
experimental rig.
Figure 6.4 shows some of the results of the parametric study performed during the design stage. For
simplicity, H was set to 492.5 mm and only L and D were allowed to variate. Three dierent opening lengths
(10, 40 and 70 mm) and three dierent cavity diameters (150, 200 and 250 mm) were analysed. In order to
keep the gure as simple as possible, the second and the third longitudinal modes were voluntarily omitted.
The lines representing the acoustic modes have been represented with a thickness which is proportional to
D. The thickness of the hydrodynamic mode curves is proportional to L.
The rst longitudinal mode H1 depends only on the cavity depth which is constant in the example given
in gure 6.4: the three curves are indeed one over the other at 348 Hz.
If the rig had not been correctly designed, for example D = 150 mm and L = 70 mm, the frequencies of
the shear layer modes would not have been high enough (SL3 at 904 Hz for U

= 53.5 m/s) to excite the


rst azimuthal mode (1340 Hz).
6.3 Opening details
In order to describe the eect of the opening length and location, a parametric study was performed.
Because of the small dimensions of the test section, no technical solution was found to gradually change
both the position and the streamwise length of the opening. Six dierent covering plates each one with
one opening were instead manufactured. By rotating a plate of 180
deg
, the position of the opening can be
changed. Therefore, with the six dierent plates, ten dierent cases can be studied. Figure 6.5 and table 6.1
represents and summarizes all the possible openings.
The covering plates were manufactured from a 7.75 mm thick Perspex panel. A single rectangular orice
with sharp (45

) chamfered leading and trailing edges was machined into each plate. All the openings have
a spanwise length of 40 mm and were centred in the spanwise direction. Their position is given by , the
distance between the trailing edge of the opening and the downstream edge of the cavity (gure 6.5).
L (mm) 40 45
Case L40EU L40HU L40CC L40HD L40ED L45EU L45HU L45CC L45HD L45ED
(mm) 189 159 99 39 9 184 157 97 37 9
Table 6.1: Cavity opening streamwise length and position.
66
6.Experimental set-up
Figure 6.5: Schematic of the opening
(a) (b)
Figure 6.6: Sketch showing the dierent rectangular openings. (a) L = 40 mm and (b) L = 45 mm. Cases: EU,
HU, CC, HD and ED. Refer to table 6.1 for the values of .
67
6.Experimental set-up
6.4 Instrumentation
6.4.1 Pitot-static tube
The ow speed was measured with a Pitot-static tube located at the end of the test section and connected
to a micromanometer Furness Controls FCO 510. The probe was introduced in the wind tunnel through
a small hole on the upper wall of the diuser section. During the PIV measurements, the Pitot tube was
removed from the test section in order to avoid ow disturbances. The micromanometer internally calculates
the velocity; a verication of the calculated mean velocity was done with a pressure transducer connected to
a liquid column manometer. The mean velocity given by the FCO 510 was found in good agreement with
the liquid column manometer measurements.
6.4.2 Microphones
Two 7 mm diameter G.R.A.S microphones (model 40PR) and fourteen Sennheiser KE4 electret microphones
were used to measure the sidewall pressure uctuations. The GRAS microphones were connected to an
amplier PCB Piezotronics 482A16. The Sennheiser microphones have a 20 - 20 000 Hz range and integrated
pre-ampliers (Jordan et al. (2002)). They were calibrated before the experimental campaign in an impedance
tube with a white noise produced by a B&K Noise Generator (type 1405) up to 20 kHz. The impedance tube
has a radius of 25 mm: its cut-o frequency is therefore around 4 kHz. All the microphones were calibrated
with respect to one of the two G.R.A.S microphones. The transfer function is plotted in gure 6.7.
Figure 6.7: Transfer functions. Calibration of the 14 Sennheiser microphones.
68
6.Experimental set-up
6.4.3 Hot wire anemometry
A Dantec CTA Module 90C10 with a Dantec 50P11 single component hot-wire (HW) probe was used for
the boundary layer characterization. The velocity measurements are automatically compensated for ow
temperature changes through a thermocouple based system.
The HW probe was directly calibrated inside the test section and the procedure is described hereafter.
The probe mounted on its holder was introduced in the test section at the position desired for the BL
characterization. A Pitot tube was also introduced in the wind tunnel and placed near the HW probe. The
location of the Pitot tube in the same plane that the HW probe was not always possible because of the nite
number of holes to introduce the Pitot tube. After the calibration, the Pitot tube was removed from the
wind tunnel and the BL characterization was done straight after. The main advantage to this procedure is
that the calibration and the measurements are done with the same conditions (temperature, pressure, HW
orientation).
6.4.4 Particle image velocimetry
The ow in the orice region was explored with a low speed LaVision PIV system. The seeding particles of
Di-Ethyl-Hexyl-Sebacat (DEHS), which had a typical particle size of 1 m, were produced by an LaVision
aerosol generator. A double pulsed Nd:YAG laser was used to illuminate the plane of interest. Images were
taken with a digital Flow Master CCD camera equipped with a 1279 1023 pixel CCD sensor and a 28 mm
focal length lens. The images were processed using Davis 7.2 software. The computed velocity elds were
then exported into Matlab for further post processing. The PIVMat Toolbox (www.fast.u-psud.fr/pivmat)
developed by Frederic Moisy from the Universite Paris-Sud was used for the production of the vector eld
gures.
6.4.5 Data acquisition card
The output signals from the instruments were acquired using a National Instrument PXI-4472B Data Ac-
quisition Card. The three signals were sampled at a frequency of Fs = 40 kHz for 8 seconds. The pressure
signals were processed in Matlab using a Fast Fourier Transform (FFT) technique. The number of points
per segment was 8192, therefore the frequency resolution was 2.44 Hz. The 19 data sets were then averaged.
6.5 Phase-averaging technique
The highest recording rate of the camera is 4.03 Hz. This value is far below the range of frequency of the
shear layer modes [0 - 1500 Hz] predicted in Section 6.2. In order to provide an average view of the velocity
uctuations, a phase-averaging technique was used. This method is applicable only when the pressure signal
has a single dominant peak in the spectrum. To identify the phase of the PIV images, the wall pressure
69
6.Experimental set-up
Figure 6.8: Schematic the PIV conguration
uctuations and the laser trigger signal were recorded simultaneously. The = 0

condition was chosen


at an arbitrary location during the pressure evolution (one quarter of period previous to the position of
the local maxima). The phase in degrees is dened as = 360

/T (gure 6.9(a)) where is the relative


time at which the velocity is measured and T is the local period of the pressure uctuations. In practice,
the signal from the microphone was ltered to reduce temporal disturbances. For every velocity studied,
1000 PIV images were acquired and divided in 8 dierent bins according on their phases. Each bin contains
between 100 and 150 PIV realizations as reported in gure 6.9(b). All the velocity elds of each bin are then
averaged in order to obtain a phase-averaged velocity eld.
6.6 Boundary layer characterization
The knowledge of the boundary layers nature is of special interest when dealing with wall bounded ows.
For the boundary layer characterization, the panel with a rectangular orice was replaced by a panel without
orice. The boundary layer was measured in a position corresponding to the upstream edge of the openings
L40EU and L45EU. Two dierent free stream velocities were investigated: 7.2 m/s and 48.3 m/s. Velocity
proles are reported in gure 6.10. At low velocities, the boundary layer was found to be laminar whereas
for high ow speed it becomes turbulent. Table 6.2 gives the characteristics of the boundary layer.
70
6.Experimental set-up
(a) (b)
Figure 6.9: (a) Temporal evolutions of the pressure signal and the laser trigger electric signal. The relative time
of the PIV realization and the local period of the pressure uctuations T. (b) Histogram of 1000 PIV realizations
distributed into height dierent phase bins
Figure 6.10: Boundary layer proles for two dierent ow velocities: 7.2 and 48.3 m/s. The Blasius (plain line) and
1/7 power proles (dashed line) are also represented.
71
6.Experimental set-up
U

(m/s) (mm)

(mm) (mm)
7.2 3.98 1.34 0.50
48.3 3.6 0.51 0.40
Table 6.2: Boundary layer properties for 2 free stream velocities: 7.2 and 48.3 m/s. Boundary layer thickness (99 %),
displacement thickness

and momentum thickness .


72
Chapter 7
Acoustic mode calculation
7.1 Analytical solution of the Helmholtz equation
In chapter 3 the acoustic modes of an open-closed cylindrical cavity were estimated by analytically solving
the Helmholtz equation (eq. 3.3). In the case of a closed-closed cavity, the boundary conditions are simpler:
the wall-normal derivative of p has to vanishes at all the walls. Therefore the general expression for the
mode frequency is:
f
m,n,q
=

2
=
c
2
_
_
j

m,n
R
_
2
+
_
q
H
_
2
_
1/2
(7.1)
where j

m,n
is (n + 1)
th
positive zero of J

m
and m, n and q are the azimuthal, radial and longitudinal
order of the mode ((m, n, q) N
3
). The frequencies and the shapes of the rst seven modes, calculated with
equation 7.1 and 3.4, are reported in table 7.1 and gure 7.1 respectively.
Mode H1 H2 AZ1 AZ1H1 H3 AZ1H2 AZ1H3
(m,n,q) (0,0,1) (0,0,2) (1,0,0) (1,0,1) (0,0,3) (1,0,2) (1,0,3)
Analytical (without opening) 350 699 846 915 1049 1097 1347
WEM 365 706 842 930 1054 1111 1354
Speaker 375 712 852 930 1054 1110 1350
Flow 377 706 851 927 1050 1102 1345
Panton & Miller (1975b) 364 699 1049
Tang & Sirignano (1973) 364 706 1052
Table 7.1: Frequencies, in hertz, of the acoustic modes of a cylindrical cavity (D = 238 mm and H = 493 mm) at 22

C.
Mode order (m, n, q): azimuthal, radial, longitudinal.
73
7.Acoustic mode calculation
(a) H1 (b) H2
(c) AZ1 (d) AZ1H1
(e) H3 (f) AZ1H2
(g) AZ1H3
Figure 7.1: Shape of the rst seven acoustic modes of a cylindrical cavity with a closed bottom wall and a closed
top surface. The color bar gives the normalized real pressure values.
74
7.Acoustic mode calculation
7.2 Wave Expansion Method (WEM)
In the previous section, the rectangular opening was neglected intentionally in order to calculate the modes
analytically. However, the presence of the opening modies the shape and the frequency of the modes. In
order to characterize the acoustic behaviour of our experimental rig (cavity and test section), a numerical
simulation was performed. A highly ecient nite dierence method originally introduced by Caruthers
et al. (1996) was used for the analysis. The approach uses wave functions which are exact solutions of the
governing dierential equation. The wave expansion method (WEM) code used for this study was developed
by Ruiz & Rice (2002) and has been examined by Bennett et al. (2009) for its applicability in ducts.
7.2.1 Overview of the method
The framework of this method is the Helmholtz equation:

2
p +k
2
p = 0 (7.2)
The aim of the WEM is to solve this equation numerically for a nite number of point in a domain. The
pressure p at a given position x
0
can be approximated locally by a combination of m plane waves
p( x
0
) =
m

j=1

j
e
i k

dj x0
(7.3)
where k is the wavenumber,

d
j
is the unit propagation direction vector of the j
th
plane wave with complex
amplitude
j
.
By using the matrix form, equation 7.3 can be re-written as
p
0
= h
0
(7.4)
where h
0
is a row vector (1 m) of the plane wave functions evaluated at x
0
and is a column vector
(m 1) of the wave strengths. The same approximation can be applied to other positions leading to
p = H (7.5)
where p is an (n 1) vector of the pressures at each position i and H
ij
= e
i k

dj xi
By combining equations 7.4 and 7.5,
p
0
= h
0
H
+
p (7.6)
where H
+
is a pseudo-inverse of matrix H. Equation 7.6 represents a so-called nite element template,
i.e. a parametrized algebraic forms that reduces to specic nite elements by setting numerical values to the
free parameters (Felippa & O nate (2007)). The raw vector (1 n) formed by the product h
0
H
+
is called
local stiness
0
.
75
7.Acoustic mode calculation
Figure 7.2: Computational domain meshed for the WEM calculation. Detail of the mesh in the opening area (section
taken in the middle plane).
Once the templates are formed (after the inclusion of the boundary conditions, Ruiz & Rice (2002)), an
overall sparse equation system may be assembled with each template independently contributing a row:
Kp = f (7.7)
where K is the overall stiness matrix and f represents a source vector.
7.2.2 Implementation
A three-dimensional unstructured mesh encompassing the wind tunnel test section and the cavity was gen-
erated with the commercial software Gambit, resulting in approximately 240 000 tetrahedral elements (g-
ure 7.2). As the largest cell in the mesh measures 0.0185 m, the highest frequency of interest (1838 Hz) is
therefore resolved with at a minimum of 10 grid points per wavelength. In the opening region the mesh was
rened in order to have at least the same resolution as the PIV measurements.
The system was excited by a monopole source located in the center of the opening. A preliminary
parametric study showed that the location of the monopole source does not have any inuence on the shape
and the frequency of the modes. The inlet and the outlet of the test section were modelled with radiation
boundary conditions.
The response of the system as a function of frequency was determined by running the code in a loop over
1000 dierent frequencies in the range [46 Hz - 1838 Hz].
76
7.Acoustic mode calculation
7.2.3 Results
In the studied frequency range, 17 dierent acoustic modes were found. The shape of the rst seven is
reported in gure 7.3. Contrarily to the open mouth case (chapter 3), here the test section does not have
an inuence on the acoustic resonances of the cavity. The WEM study reveals indeed that there are no test
section/cavity modes excited by the acoustic source. The shapes of modes calculated by the WEM are very
similar to the shapes of the modes inside a completely closed cylinder (gure 7.1). In order to appreciate the
dierences between these two cases, the real part of the complex pressure is analysed in the middle plane of
the cavity (gure 7.4). Unsurprisingly, the opening does not only modify the pressure in the orice vicinity
as shown by the rst longitudinal mode but in some cases the pressure pattern changes everywhere inside
the cavity: the third longitudinal mode clearly illustrates this shape distortion.
The frequencies of the WEM modes is given in table 7.1: small dierences can be noticed when they
are compared with the analytically case. These dierences are generated by the distortion of the pressure
pattern. The pressure vanishes in the test section which induces a pressure adaptation in proximity to the
orice. The eects of the opening are similar to a reduction of the cavity depth. Consequently the frequencies
of the modes increase. An illustration of this is given by the mode AZ1H1 (gure 7.4(b)): there is a clear
virtual reduction of the cavity depth compared to the closed case.
7.3 Helmholtz resonance frequency
The frequency of the Helmholtz resonance was calculated with the classical Helmholtz theory and with
the Panton & Miller (1975b) and Tang & Sirignano (1973) transcendental equations which also predict the
frequencies of the longitudinal modes of the cavity. Table 7.1 summarizes the results for the longitudinal
modes and table 7.2 the frequency of the Helmholtz resonance. The Rayleigh end correction with an eective
radius was used to estimate the eective depth of the opening (neck length) for all the calculations:
l
eq
= l + 2l = l + 2 0.8242 1.06 S
3/4
U
1/2
(7.8)
Method Equation Frequency (Hz)
Prediction: Classical Helmholtz theory Eq. 1.9 73.2
Prediction: Improved Helmholtz theory Eq. 1.13 68.4
Prediction: Panton & Miller (1975b) Eq. 1.11 68.3
Prediction: Tang & Sirignano (1973) Eq. 1.12 68.3
Experimental: Speaker 84.2
Table 7.2: Frequencies, in hertz, of the Helmholtz resonance of the cylindrical cavity (D = 238.5 mm and H = 492.5 mm)
with a rectangular opening (L = 45 mm) at 22

C. Eective length leq = 45.9 mm.


77
7.Acoustic mode calculation
(a) H1 (b) H2
(c) AZ1 (d) AZ1H1
(e) H3 (f) AZ1H2
(g) AZ1H3
Figure 7.3: Shape of the rst seven acoustic modes. The color bar gives the normalized real pressure values.
78
7.Acoustic mode calculation
(a) (b) (c)
(d) (e) (f)
Figure 7.4: Normalized real pressure on the middle plane of the cavity. WEM results (rst row) and analytical
solutions (second row). Three acoustic resonance are given: (a and d) rst longitudinal mode H1; (b and e) rst
azimuthal-longitudinal mode AZ1H1; (c and f) third longitudinal mode H3. The position of the microphone for the
PIV campaign (section 8.2) is indicated by a black square.
79
7.Acoustic mode calculation
7.4 Response of the resonator to an external excitation
7.4.1 Acoustic excitation
The cavity response to an acoustic excitation was experimentally investigated in order to validate the numer-
ical acoustic simulation. For this purpose, a small loudspeaker radiating broadband noise was used. After
trying dierent locations (inlet and outlet of the wind-tunnel), the lateral wall of the test section opposite
to the cavity was removed and replaced by the loudspeaker. A disadvantage of such a conguration is that
the inuence of the closed test section on the acoustics could not be studied. However, as shown previously
by the numerical simulation, the test section/cavity mode should not be excited.
The transfer function between the speakers input signal and the pressure measured with a GRAS mi-
crophone on the walls of the cavity was calculated and its magnitude is plotted in gure 7.5. The WEM
transfer function between the monopole source and a location corresponding to the experimental microphone
position is also given. There is a good qualitative agreement between the experimental and the numerical
data. Table 7.1 gives the frequencies of the peaks in the experimental transfer function.
Figure 7.5: Response of the cavity to dierent excitations: monopole acoustic source (WEM, red), broadband noise
(experimental, black) and ow (experimental, blue).
7.4.2 Boundary layer excitation
The resonator was exposed to a ow at room temperature through the wind tunnel. The unsteady pressure
inside the cavity was measured with a ush mounted microphone. Figure 7.5 shows the pressure autospectrum
at a ow velocity of 21 m/s. The only conclusion given here is that the grazing ow excites all the acoustic
80
7.Acoustic mode calculation
modes of the cavity. A deepest analysis is made is the next chapter.
81
7.Acoustic mode calculation
82
Chapter 8
Experimental results
8.1 Response of the resonator to a grazing ow
8.1.1 Baseline opening: case L45EU
An automated velocity sweep of the tunnel was performed by controlling the centrifugal blower motor speed
with a LabView program. Results are presented in gure 8.1 together with of the shear layer modes calculated
according to the Rossiter (1964)s formula (equation 1.1). Values of = 0 and = 0.46 have been used here
as a best t to the measured data.
In the ow speed range [4 - 9 m/s], the rst shear layer mode is found to excite the Helmholtz resonance.
The pressure uctuations over the walls of the cavity are reported in gure 8.2(a) for U

= 7.4 m/s. The


frequency of the dominant peak of the spectra is 67 Hz, which, if the frequency resolution of the sprectrum
(2.44 Hz) is considered, is in good agreement with the frequency found with equation 1.12. The spectral
analysis reveals that the pressure signal contains, apart from the harmonics of the Helmholtz resonance, the
acoustic modes of the cavity (H1, AZ1, ...).
For velocities between 28 m/s and 48 m/s, the rst longitudinal mode H1 is excited by the shear layer
mode I. The pressure uctuations on the sidewall of the cavity are strong for the whole ow speed range.
The case U

= 46.3 m/s is reported in gure 8.2(b). The oscillations are ten times higher than in the former
case (gure 8.2(a)). The pressure signal contains also the harmonics of the H1 mode: the sharp peak at 751
Hz, between H2 and AZ1, corresponds to the rst harmonic of H1. The intensity of the second harmonic is
surprisingly as high as the rst harmonic: this is mainly due to the fact that the frequency of the second
harmonic is very close to the AZ1H2 mode.
When the ow velocity reaches 48 m/s, the rst hydrodynamic mode (SL1) is overtaken by SL2. This
switch occurs because the shear layer instability II gets closer to the frequency of a dierent acoustic reso-
nance: in this case, the combination of the rst azimuthal and the rst longitudinal mode (AZ1H1). It is
interesting to notice that two eigenmodes are skipped (H2 and AZ1). Figure 8.2(c) gives the pressure both
83
8.Experimental results
Figure 8.1: Acoustic response inside the cavity as a function of tunnel ow-speed. Superimposed on the plot are the
theoretical shear layer modes (SL), the WEM acoustic modes (H1, AZ1...) and the Helmholtz resonance (HR). The
rst longitudinal cavity/test section mode (HW1) calculated analytically is also reported. Frequencies are given in
non-dimensional form (Helmholtz number = 2Rf/c).
in the time and in the frequency domain for a ow velocity of 48.4 Hz. Even if AZ1 is just 11 Hz higher
than the theoretical frequency of the shear layer instability II, the AZ1H1 mode, that is 89 Hz higher, is
the one coupled with the shear layer mode II. The vortex sound theory developed by Howe (1975) is a good
start in order to explain the predilection for certain specic eigenmodes for a ow-acoustic coupling. In the
Howes acoustic analogy, the Coriolis density forces
0
w u are identied as the principal source of sound.
The acoustic power generated by the vortical eld can be calculated by equation 1.14 which states that
the is proportional to the triple u
acoust
( wu). From this formula it is clear that if the acoustic particle
velocity at the opening has the same orientation as the velocity or the vorticity, there is no acoustic power
generated. Let us now take the example of the rst azimuthal mode (AZ1). It has been shown numerically,
in section 7.2.3, that this mode is symmetrical about plane Oyz (it could has been symmetrical about any
other vertical plane; however the opening location imposes the symmetry plane) and the 16 simultaneous
wall pressure measurements have conrm this (results not given here). The mode AZ1 does not radiate
because the acoustic particle velocity has the same orientation as the mean ow. This is a very simplistic
explanation especially because it assumes the directions of u
acoust
, w and u to be known a priori.
The hydrodynamic mode II remains locked on AZ1H1 mode until 51 m/s. When the ow speed reaches
51.5 m/s there is again a switch of predominant acoustic resonance. Figure 8.2(d) shows that the third
longitudinal mode dominates the spectrum. It has been found that for the highest blower rotational speed
(corresponding to U

= 52 m/s), this particular ow-acoustic coupling persisted. It is expected that for


84
8.Experimental results
(a) (b)
(c) (d)
Figure 8.2: Wall pressure recorded with a microphone (top) and the corresponding spectrum (bottom). The ow
velocity was xed to 7.4 m/s (a), 46.3 m/s (b), 48.4 m/s (c) and 51.5 ms/s (d). The frequencies of the acoustic modes
numerically predicted (table 7.1) are also reported (H1, AZ1, ...). The frequency of the Helmholtz resonance (68.3
Hz). The rst three hydrodynamic shear layer modes (I, II and III) calculated with the Rossiter formula (equation 1.1,
[, ] = [0, 0.42]) are plotted in red.
85
8.Experimental results
(a) (b)
Figure 8.3: Pressure amplitudes at 73 Hz (a) and 376 Hz (b) as a function of the wind tunnel ow speed. Opening
L45ED.
higher ow speeds the rst shear layer will lock-on the second longitudinal eigenfrequency of the cavity.
8.1.2 Strength of lock-on
The receptivity of the cavity to shear layer excitation was quantied using a strength of lock-on parameter,
as suggested by Mendelson (2003) and described by Yang et al. (2009). The parameter chosen was the
amplication of the cavity pressure level above the background noise level, as dened by a linear scaling in
log-log space. For each frequency, a linear t was made to the spectral density in decibels. The magnitude
of the pressure above this linear t was designated as the strength of lock-on (SoL). At the ow speed
U

, the SoL of the frequency f can be calculated through:


SoL(M

, f) = p
dB
(M

, f) [p
dB
(1, f) + 20nlog(M

)] (8.1)
where p
dB
denotes the pressure amplitude (in dB) function of the frequency f and the Mach number of
the ow (M

= U

/c) and n the exponential at which the broadband turbulent noise grows. Two examples
of the linear tting are given in gure 8.3. The same procedure was applied to all the frequencies of the
spectra and the results are given in gure 8.4 in the form of a colour map.
8.1.3 Inuence of the location of the opening
The eect of the opening location was quantied using the SoL parameter previously introduced. Three
dierent positions were explored: = 99, 39 and 9 mm (L40CC, L40HD and L40ED respectively). The
results are summarized in gure 8.5, where the contour lines encircle values of SoL higher than 13 dB. The
86
8.Experimental results
(a) (b)
Figure 8.4: Acoustic response (a) and the corresponding strength of lock-on (b) inside the cavity as a function of
tunnel ow-speed. The opening L40CC.
threshold criterion is an useful and straightforward technique to identify the resonance conditions in a ow
excited cavity.
There are some interesting dierences between the three opening analysed. It is clear that the resonance
lock-on for H1 is much stronger when the opening is in the center of the cavity (L40CC): the rst shear
layer hydrodynamic mode remains locked-on H1 for a wider range of velocities than for the other two orice
positions. Especially noteworthy is the cut-on of the azimuthal mode AZ1H1 at velocities above 45 m/s for
L40ED and L40HD, which does not occur for L40CC. When the opening is o-center, the shear layer pressure
uctuations tend to excite AZ1H1 because they are closer to an acoustic anti-node. On the contrary, the
central location is a pressure node for this acoustic mode as seen in chapter 7. The third longitudinal mode
is excited by the shear layer mode II from 50.3 m/s to 52.3 m/s for L40HD while this resonance occurs only
for the highest tested ow speed for L40ED.
Chanaud showed, rst numerically and then experimentally (Chanaud (1994, 1997)), that a displacement
of the orice away from the center results in decrease of the resonant frequency . In the present experiments,
a change of frequency is not observed. However, from gure 8.5, the Helmholtz resonance is more likely to
be excited when the opening is centred or half centred.
For low velocities (U

< 15 m/s), the amplitudes of the second and third shear layer modes are higher
when the opening is at = 39 mm. This observation could not be explained by the author and its
interpretation is a start of future studies.
87
8.Experimental results
8.2 Shear layer dynamics
Velocity measurements in the orice region allow greater insight into the uid dynamics of the shear layer
to be obtained. The four dierent velocities explored (7.4, 46.3, 48.4 and 51.5 m/s) have already been
presented in section 8.1.1. They correspond to: the rst shear layer mode locked on the Helmholtz resonance
(gure 8.6), the rst shear layer mode locked on the rst longitudinal mode (gure 8.7), the second shear
layer mode locked on AZ1H1 (gure 8.8) and the second shear layer mode locked on H3 (gure 8.9) for the
baseline opening (L45EU). The phase-averaged velocity and vorticity elds are reported. Height dierent
phases were calculated during the processing but 45

, 135

, 225

and 315

were omitted for brevitys sake.


8.2.1 First shear layer mode
Helmholtz resonance
At a velocity of 7.4 m/s, the rst shear layer hydrodynamic mode locks on the Helmholtz resonance (g-
ure 8.2(a)). The phase average velocity and vorticity elds are reported in gure 8.6. In the upstream
portion of the opening (1 < x/L < 0), the shear layer appears to ap whereas in the downstream portion
(0 < x/L < 1) it rolls up into a single vortex. As the vortex is convected downstream along the cavity
opening, it grows and when it reaches the downstream edge, it splits in two parts: one part is captured by
the cavity while the other part escapes from it. The splitting mechanism is not very clear mainly because
the area under the edge is not correctly illuminated by the laser.
Acoustic mode H1
The phase averaged velocity and vorticity elds are given in gure 8.7 for U

= 46.3 m/s. During an acoustic


cycle, a single vortex is generated as in the previous case. In fact, the 7.4 m/s and the 46.3 m/s cases are
very similar even if the nature of the resonant acoustic mechanism is dierent: periodic compression of the
uid inside the cavity in the former case as opposed to a standing acoustic wave in the second case. For
both velocities however the predominant instability in the shear layer is the rst hydrodynamic mode.
8.2.2 Second shear layer mode
Acoustic mode AZ1H1
The phase averaged velocity and vorticity elds are given in gure 8.8. The dynamics is less clear than
for the two velocities formerly presented where the shear layer was rolling up into one single vortex per
acoustic cycle. Here instead, it seems to roll-up into two smaller vortices. The apping noticed for the rst
hydrodynamic mode is not discernible here. Furthermore, the air does not appear to enter into the cavity,
not even near the downstream edge.
88
8.Experimental results
Acoustic mode H3
In gure 8.9 the phase averaged velocity and vorticity elds are reported. During an acoustic cycle, two
vortices are generated in the shear layer. From the velocity elds, the rolling-up appears to take place in
the rst half of the orice, sooner than when the shear layer mode I is dominant. Again, the ow does not
seem to enter into the cavity at any phase of the cycle. The main dierence between AZ1H1 and H3 is the
phase of the velocity elds: the = 180

for a ow speed of 48.4 m/s corresponds to the = 0

at 51.5
m/s. This change of phase is due to the position of the microphone (gure 7.4): the acoustic oscillations at
the opening are in phase with the uctuations at the microphone location for the mode AZ1H1 and out of
phase for the mode H3.
8.3 Acoustic power
In this section the Howe (1975) vortex sound theory is applied to the opening region for the baseline orice
(L45EU). The generation of acoustic power by the vortical eld is calculated through the Howes integral
(equation 1.14). The velocity and the vorticity elds were extracted from the PIV data and the particle
velocity numerically from the WEM simulation. An important assumption in the development followed is
that the hydrodynamic and to acoustic elds can be computed independently. This assumption is motivated
by the fact that the orice, i.e the region where the acoustics is generated, is compact compared to the
acoustic wavelengths.
8.3.1 Computation of the acoustic particle velocity
As seen in section 7.2, the acoustic simulation gives a complex pressure eld whose amplitude depends on
the intensity of the monopole source. In order to scale the pressure, the experimental data acquired with a
ush-mounted microphone was used. It is assumed that the pressure inside the cavity has a simple harmonic
behaviour. This assumption is good for strongly resonant states. Therefore, the pressure P
acoust
at any
point inside the cavity (x, y, z) and at any moment of time (t) can be expressed as:
P
acoust
(x, y, z, t) = cst P
WEM
(x, y, z, f
acoust
) sin(2f
acoust
t) (8.2)
where, P
WEM
is the pressure from the numerical simulation, f
acoust
is the frequency of the dominant
acoustic mode during the experimental testing and cst a constant necessary to match the experimental
pressure measurements. The calculation of the particle velocity eld was done by integrating the linearized
Euler momentum equation:

0
u
acoust
t
+P
acoust
= 0 (8.3)
89
8.Experimental results
The integration is straightforward because the eld is time harmonic. As an example, pressure and
velocity elds in the orice region are given in gure 8.10 for the modes H1, AZ1H1 and H3. The area of
interest for this study is delimited by a rectangle. Even if inside the cavity the acoustic modes have very
dierent shapes (gure 7.4), locally, around the opening the acoustic velocity elds are very similar for the
three cases (gure 8.10).
8.3.2 Time-averaged acoustic power
Three free stream velocities were chosen for the study: 46.3, 48.4 and 51.5 m/s. These velocities correspond
respectively to the rst shear layer mode (SL1) locked-on with the rst longitudinal mode (H1), the second
shear layer mode (SL2) exciting the rst azimuthal-longitudinal mode (AZ1H1) and the second shear layer
mode (SL2) amplied by the third longitudinal mode H3. The shape of these three acoustic modes is
displayed in gure 7.4 and 8.10. The instantaneous acoustic power was found for 8 dierent phases by
computing the integrand of equation 1.14. This intermediate result, even if essential for understanding of
the sound production, is not presented here for sake of brevity.
The acoustic power generated by the vortices in the orice region over an entire acoustic period can be
obtained by averaging the computed instantaneous acoustic powers:
< >=<
0
_
V
u
acoust
( w u) dV > (8.4)
where <> denotes the time averaging over one period of oscillation.
The net acoustic energy E = <> /f produced by a free stream during an acoustic cycle of 46.3 m/s
(SL1-H1) is given in gure 8.11(a). The spatial distribution of the acoustic energy is characterized by a
distinct source-sink pair. This corresponds to the fact that when the rst shear layer mode predominates, a
single large-scale vortex is generated during an acoustic period.
The two cases for which the second shear layer mode is dominant are presented in gure 8.11(b) and
8.11(c). It is interesting to observe the degree of similarity between these two patterns: two source-sink
pairs are found above the opening. Again, the shear layer rolling up into large-scale vortices generates this
pattern: the second shear layer mode produces two vortices during an acoustic cycle. Oshkai et al. (2008)
(alternatively Velikorodny et al. (2010)) also reported two source-sink pairs when calculating the acoustic
power for coaxial side branches for a dominant second shear layer mode (gure 1.12).
90
8.Experimental results
(a) L40CC (b) L40HD
(c) L40ED (d)
(e)
Figure 8.5: Contour of strength of lock-on higher than 13 dB. Three dierent orice position are given: L40CC,
L40HD and L40ED.
91
8.Experimental results
Figure 8.6: Phase-averaged velocity (right) and vorticity (left) elds at the opening of the cavity. Four phases are
given: 0

, 90

, 180

and 270

. Flow speed 7.4 m/s. Predominant shear layer mode: SL1. Predominant acoustic
mode: Helmholtz resonance.
92
8.Experimental results
Figure 8.7: Phase-averaged velocity (right) and vorticity (left) elds at the opening of the cavity. Four phases are
given: 0

, 90

, 180

and 270

. Flow speed 46.3 m/s. Predominant shear layer mode: SL1. Predominant acoustic
mode: the rst longitudinal resonance (H1).
93
8.Experimental results
Figure 8.8: Phase-averaged velocity (right) and vorticity (left) elds at the opening of the cavity. Four phases are
given: 0

, 90

, 180

and 270

. Flow speed 48.4 m/s. Predominant shear layer mode: SL2. Predominant acoustic
mode: AZ1H1.
94
8.Experimental results
Figure 8.9: Phase-averaged velocity (right) and vorticity (left) elds at the opening of the cavity. Four phases are
given: 0

, 90

, 180

and 270

. Flow speed 51.5 m/s. Predominant shear layer mode: SL2. Predominant acoustic
mode: the third longitudinal resonance (H3).
95
8.Experimental results
(a) (b)
(c) (d)
(e) (f)
Figure 8.10: Acoustic pressure eld Pacoust (left), acoustic velocity eld uacoust in green and streamlines in blue
(right) at the orice region calculated with the WEM simulation and scaled by the experimental data. From top to
bottom: H1, AZ1H1 and H3. Pressure elds are given at = 90

whereas velocity elds at = 0

.
96
8.Experimental results
(a)
(b)
(c)
Figure 8.11: Distribution of the net acoustic energy generated per acoustic cycle on the orice. (a): 46.3 m/s,
SL1-H1. (b): 48.4 m/s, SL2-AZ1H1. (c): 51.5 m/s, SL2-H3.
97
8.Experimental results
98
Chapter 9
Conclusion
A cylindrical Helmholtz resonator was experimentally studied by wall pressure measurements and 2D PIV.
The designed experiment allows dierent resonant modes to be excited depending on both ow speed and
orice location.
At low ow speeds, the Helmholtz resonance was found to be excited by the rst shear layer hydrodynamic
mode at the frequency predicted by an improved Helmholtz resonance formulation.
For higher velocities, lock-on between the rst two shear layer hydrodynamic modes and dierent eigen-
modes of the cavity was observed. The location of the cavitys opening was found to be a major factor in
determining which acoustic mode is excited. Specically, a combination mode (azimuthal and longitudinal)
was found to generate lock-on only when the opening was located o-center. This was ascribed to the fact
that the main axis of the cavity is a pressure node for the rst azimuthal mode.
The dynamics of the shear layer were then explored through PIV measurements for four ow speeds
corresponding to strongly resonant cases. Phase averaged PIV allowed the coherent structures present in
the shear layer to be examined and the interaction between the cavity resonances and the shear layer to be
analysed. The rst hydrodynamic mode is characterized by a apping movement in the upstream portion
of the resonators opening and by the rolling up of the shear layer in a single vortex in the downstream
portion of the orice. The number of vortices generated per acoustic cycle increases to two when the
second hydrodynamic instability becomes dominant. This observation is consistent with previous studies.
The experimental results show that the order of the dominant acoustic resonant mode does not aect the
organization of the shear layer. A possible explanation is that in the opening region, the acoustic velocity
elds are similar.
The vortex sound theory of Howe was applied in order to characterize the energy transfer mechanisms
on the orice region. The velocity and the vorticity elds were extracted from the PIV data and the particle
velocity was calculated using a Wave Expansion Method (WEM) simulation. Three dierent ow conditions
generating acoustic resonance were analysed. For each case, the acoustic sources were localized revealing
99
9.Conclusion
that the spacial organization of the sound production depends exclusively on the predominant shear layer
hydrodynamic mode.
100
Chapter 10
Summary
The main objective of this research was to increase the understanding of mechanisms governing ow-induced
resonance in cylindrical cavities. The dissertation was divided into two main parts corresponding to the
two cases studied: an open mouth cylindrical cavity and a cylindrical Helmholtz resonator with a rectan-
gular orice. Specic conclusions have been provided in chapters 5 and 9 and a summary of the major
accomplishments is given hereafter.
By using dierent measurement techniques (hot-wire anemometry, Particle Image Velocimetry and wall
mounted microphones), the shear layer hydrodynamics modes were identied and characterised. Extensive
data were obtained over a broad subsonic speed range for the two cavities studied. In both cases, the
classical formulation for the prediction of the frequency of the sheartones proved to be in agreement with the
experimental data. The constant accounting for the phase delay in the Rossiter (1964)s feedback mechanism
was found to be negligible for the low Mach numbers tested. The description of the shear layer over the
Helmholtz resonator was achieved by phase averaging the PIV data. The rst shear layer mode was found
to correspond to one single vortex whereas the second mode to two vortices as described by previous studies.
The rst hydrodynamic mode is also characterized by a apping movement of the shear layer especially on
the upstream portion of the opening. An important result is that the dominant acoustic mode does not have
an inuence on the shear layer morphology.
The acoustic resonances of the two test rigs were investigated numerically and good agreement was found
with pressure measurements. When the cavity opening is partially covered, the shear layer hydrodynamic
modes tend to couple with specic eigenmodes of the cavity. In particular, lock-on was detected at three
dierent acoustic modes using a strength of lock-on parameter. The position of the orice has an inuence
on the ow-acoustic coupling: a combination of the rst longitudinal and the rst azimuthal modes is strongly
excited when the opening is o-center of the cavity axis.
The eect of the connement has a substantial inuence on the acoustics of the ow excited open mouth
cavity: lock-on between the shear layer modes and the acoustic resonances of the test section was reported.
101
10.Summary
The common characteristic of the lock-on modes is the quarter wave length shape inside the cavity.
A semi-empirical approach was used to estimate the acoustic power generated by a ow-excited Helmholtz
resonator. The vortex sound theory of Howe (1975) was used to quantify the energy transfer between the
acoustic eld and the turbulent ow. Velocity eld measurements in conjunction with a numerical simulation
of the acoustic eld were used for the calculation of the physical quantities in the Howes integral. The spatial
distribution of the acoustic power generated over the orice was found to depend only on the predominant
shear layer hydrodynamic mode: the rst sheartone produces one acoustic source-sink pair whereas two
source-sink pairs were identied when the second sheartone is dominant. Results show this approach to be
promising for the understanding of ow-induced acoustic resonance.
The presented experimental work also led to the description of some ow features of an open mouth
cylindrical cavity past by a grazing ow. These results were obtained using hot-wire anemometry, 2-D PIV
and wall pressure measurements. The three-dimensionality and the symmetry of the ow, expected for the
aspect ration studied, was conrmed. The velocity measurements, obtained from hot-wire anemometry using
a single wire probe on the shear layer and the wake of a relatively deep open mouth cylindrical cavity (H/D
= 1.357), are the rst of this kind in the literature because they concern planes normal to the streamwise
direction. They provide a robust database for CFD validation. Other original results that can be used for
comparison, are the velocity elds at dierent horizontal planes inside the cavity (PIV) and the uctuating
pressure levels at the walls of the cavity.
102
Appendix A
Acoustic power
By following the vortex sound theory of Howe (1975, 1980), an expression for the acoustic power generated
is derived bellow.
The total uid velocity u can be decomposed in an irrotational part and a rotational part u
rot
=

:
u = +

(A.1)
A homentropic ow satises Croccos form of the Euler momentum equation:
u
t
+B = u (A.2)
where B is the total enthalpy and where the vorticity is dened by = u. Note that the friction
and heat transfer are neglected.
By using the decomposition of u, equation A.2 becomes:
u
rot
t
+

t
+B = u
rot
(A.3)
By taking the scalar product of u
rot
and by simplifying:
1
2

t
u
2
rot
+u
rot

_

t
+B
_
= u
rot
( ) (A.4)
This equation can be simplies further by recalling that (u
rot
) = 0 :
1
2

t
u
2
rot
+
_
u
rot
_

t
+B
__
= u
rot
( ) (A.5)
By integrating over a volume:
1
2

t
_
V
u
2
rot
dV +
_
V

_
u
rot
_

t
+B
__
dV =
_
V
u
rot
( )dV (A.6)
103
A.Acoustic power
The divergence theorem states that:
_
V

_
u
rot
_

t
+B
__
dV =
_
S
(u
rot
n)
_

t
+B
_
dS (A.7)
The volume V chosen for the integration is a large sphere whose radius can be set arbitrarily. For a large
spheres radius, the normal component of u
rot
vanishes on the surface S and therefore:
1
2

t
_
V
u
2
rot
dV =
_
V
u
rot
( )dV (A.8)
Furthermore, by assuming small acoustic perturbation, the unsteady irrotational part of the velocity eld
can be dened as the acoustical velocity: u
acoust
= .
1
2

t
_
V
u
2
rot
dV =
_
V
u
rot
( u
acoust
)dV (A.9)
By multiplying this equation by
0
and by rearranging its right hand side:

0
1
2

t
_
V
u
2
rot
dV =
0
_
V
u
acoust
( u
rot
)dV (A.10)
The left hand side of the equation gives the rate at which the the vorticity-bearing part of the velocity eld
generates kinetic energy. Right hand side represents the work performed by the lift
0
u
rot
experienced by
vortex elements in the velocity eld of the sound. Equation A.10 quanties
abs
, the rate at which acoustic
energy is absorbed by the vortical eld. Therefore the acoustic power generated by the ow is given by:
=
0
_
V
u
acoust
( w u
rot
) dV (A.11)
104
References
Alvarez J.O., Kerschen E.J. (2005). Inuence of wind tunnel walls on cavity acoustic resonances, 11th AIAA/CEAS
Aeroacoustics Conference, 2005-2804, Monterey, California.
Aly K., Ziada S. (2010). Flow-excited resonance of trapped modes of ducted shallow cavities, Journal of Fluids and
Structures 26, pp. 92120.
Anderson J.S. (1977). The eect of an air ow on a single side branch Helmholtz resonator in a circular duct, Journal
of Sound and Vibration 52(3), pp. 423431.
Arthus D., Ziada S. (2009). Flow-excited acoustic resonances of coaxial side-branches in annular duct, Journal of
Fluids and Structures 25, pp. 4259.
Bennett G.J., OReilly C.J., Liu H. (2009). Modelling multi-modal sound transmission from point sources in ducts
with ow using a wave-based method, 16th Congress on Sound and Vibration, Krakow, Poland.
Bilanin A.J., Covert E.E. (1973). Estimation of possible excitation frequencies for shallow rectangular cavities, AIAA
Journal 11, pp. 347351.
Blake W.K. (1986). Mechanics of ow induced sound and vibrations, Academic Press, Orlando, FL.
Block P.J.W. (1976). Noise response of cavities of varying dimensions at subsonic speeds, Tech. Rep. TN D-8351,
NASA.
Bruggeman J.C., Hirschberg A., van Dongen M.E.H., Wijnands A. (1991). Self-sustained aero-acoustic pulsations in
gas transport systems: experimental study of the inuence of closed side branches, Journal of Sound and Vibration
150, pp. 371393.
Camussi R., Guj G., Barbagallo D., Prischich D. (2000). Experimental characterization of the aeroacoustic behavior
of a low speed wind tunnel, 6th AIAA/CEAS Aeroacoustics Conference, 2000-1986, Lahaina, Hawaii.
Camussi R., Guj G., Di Marco A., Ragni A. (2006a). Propagation of wall pressure perturbations in a large aspect
ratio shallow cavity, Experiments in Fluids 40, pp. 612620.
Camussi R., Guj G., Ragni A. (2006b). Wall pressure uctuations induced by turbulent boundary layers over surface
discontinuities, Journal of Sound and Vibration 294, pp. 177204.
105
REFERENCES
Camussi R., Robert G., Jacob M. (2008). Cross-wavelet analysis of wall pressure uctuations beneath incompressible
turbulent boundary layers, Journal of Fluid Mechanics 617, pp. 1130.
Caruthers J.E., Engels R.C., Ravinprakash G.K. (1996). A wave expansion computational method for discrete fre-
quency acoustics within inhomogeneous ows, 2nd AIAA/CEAS Aeroacoustic Conference, State College, Pennsyl-
vania.
Cattafesta L., Williams D., Rowley C., Alvi F. (2003). Review of active control of ow-induced cavity resonance,
33rd AIAA Fluid Dynamics Conference, 2003-3567, Orlando, Florida.
Chanaud R. (1994). Eects of geometry on the resonance frequency of Helmholtz resonators, Journal of Sound and
Vibration 178(3), pp. 337348.
Chanaud R. (1997). Eects of geometry on the resonance frequency of Helmholtz resonators, Part II, Journal of
Sound and Vibration 204(5), pp. 829834.
Chatellier L., Laumonier J., Gervais Y. (2004). Theoretical and experimental investigations of low Mach number
turbulent cavity ows, Experiments in Fluids 36, pp. 728740.
Chicheportiche D., Gloerfelt X. (2010). Direct noise computation of the ow over cylindrical cavities, 16th
AIAA/CEAS Aeroacoustics Conference, 2010-3775, Stockholm, Sweden.
Cotroni A., Di Felice F., Romano G.P., Elefante M. (2000). Investigation of the near wake of a propeller using particle
image velocimetry, Experiments in Fluids 29, pp. S227S236.
Crighton D., Dowling A., Ffowcs Williams J., Heckl M., Leppington F. (1994). Modern methods in analytical acoustics:
lecture notes, Springer-Verlag, 1994.
Curle N. (1955). The inuence of solid boundaries upon aerodynamic sound, Proceedings of the Royal Society of
London. Series A, Mathematical and Physical Sciences 231(1187), pp. 505514.
Czech M.J., Crouch J.D., Stoker R.W., Strelets M.K., Garbaruk A. (2006). Cavity noise generation for circular and
rectangular vent holes, 12th AIAA/CEAS Aeroacoustics Conference, 2006-2508, Cambridge, Massachusetts.
Dalmont J.P., Nederveen C.J., Joy N. (2001). Radiation impedance of tubes with dierent anges numerical and
experimental investigations, Journal of Sound and Vibration 244, pp. 505534.
De Metz F.C., Farabee T.M., David W. (1977). Laminar and turbulent shear ow induced cavity resonances, 4th
AIAA Aeroacoustics Conference, 77-1293, Atlanta, Georgia.
Dequand S., Hulsho S.J., Hirschberg A. (2003). Self-sustained oscillations in a closed side branch system, Journal
of Sound and Vibration 265, pp. 359386.
Desvigne D., Marsden O., Bogey C., Bailly C. (2010). Calcul direct du bruit rayonne par un ecoulement laminaire `a
Mach 0.2 aeurant une cavite cylindrique, Congr`es Francais dAcoustique, Lyon.
106
REFERENCES
Di Florio D., Di Felice F., Romano G.P. (2002). Windowing, re-shaping and re-orientation interrogation windows
in particle image velocimetry for the investigation of shear ows, Measurement Science and Technology 13, pp.
953962.
Dybenko J., Savory E. (2008). An Experimental Investigation of Turbulent Boundary Layer Flow over Surface-
Mounted Circular Cavities, Proc. IMechE Vol. 222 Part G: J. Aerospace Engineering, vol. 117.
El Hassan M., Labraga L., Keirsbulck L. (2007). Aero-acoustic oscillations inside large deep cavities, 16th Australasian
Fluid Mechanics Conference, Australia.
Elder S.A. (1978). Self-excited depth-mode resonance for a wall-mounted cavity in turbulent ow, Journal of Acoustical
Society of America 64, pp. 877890.
Felippa C., O nate E. (2007). Nodally exact Ritz discretizations of 1D diusion?absorption and Helmholtz equations
by variational FIC and modied equation methods, Computational Mechanics 39(2), pp. 91111.
Finnegan S., Oshkai P., Meskell C. (2010). Experimental methodology for aeroacoustic source localisation in ducted
ows with complex geometry, 16th AIAA/CEAS Aeroacoustic Conference, Stockholm, Sweden.
Finnegan S.L. (2011). Resonant aeroacoustic source localisation in ducted blu body ows, Ph.D. thesis, University
of Dublin.
Gaudet L., Winter K.G. (1973). Measurements of the drag of some characteristic aircraft excrescences immersed in
turbulent boundary layers, Tech. Rep. Aero 1538, Royal Aircraft Establishment.
Geveci M., Oshkai P., Rockwell D., Lin J.C., Pollack M. (2003). Imaging of the self-excited oscillation of ow past a
cavity during generation of a ow tone, Journal of Fluids and Structures 18, pp. 665694.
Grottadaurea M. (2009). Aerodynamics and near-eld acoustics of a subsonic cylindrical cavity ow by parallel CFD,
Ph.D. thesis, University of Leicester.
Grottadaurea M., Rona A. (2008). The radiating pressure eld of a turbulent cylindrical cavity ow, 14th AIAA/CEAS
Aeroacoustics Conference, 2008-2852, Vancouver, British Columbia Canada.
Haigermoser C. (2009). Application of an acoustic analogy to PIV data from rectangular cavity ows, Experiments
in Fluids 45, pp. 145157.
Haigermoser C., Scarano F., Onorato M. (2009). Investigation of the ow in a circular cavity using stereo and
tomographic particle image velocimetry, Experiments in Fluids 46, pp. 517526.
Hart D.P. (1998). High-speed PIV analysis using compressed image correlation, Journal of Fluids Engineering 120,
pp. 463470.
Heller H.H., Bliss D.B. (1975). The physical mechanism of ow induced pressure uctuations in cavities and concepts
for suppression, 2nd AIAA Aeroacoustics Conference, 75-491, Hampton, VA.
Hering T., Dybenko J., Savory E. (2006). Experimental Verication of CFD Modeling of Turbulent Flow over Circular
Cavities using FLUENT, CSME 2006 Forum.
107
REFERENCES
Hiwada M., Kawamura T., Mabuchi I., Kumada M. (1983). Some characteristics of ow pattern and heat transfer
past a circular cylindrical cavity, Bulletin of the JSME 26, pp. 17441752.
Howe M.S. (1975). Contributions to theory of aerodynamic sound, with application to excess jet noise and the theory
of the ute, Journal of Fluid Mechanics 71,4, pp. 625673.
Howe M.S. (1980). The dissipation of sound at an edge, Journal of Sound and Vibration 70, pp. 407411.
Howe M.S. (1997). Edge, cavity and aperture tones at very low Mach numbers, Journal of Fluid Mechanics 330, pp.
6184.
Ingard U. (1953). On the theory and design of acoustic resonators, Journal of Acoustical Society of America 25(6),
pp. 10371061.
Jordan P., Fitzpatrick J.A., Vali`ere J.C. (2002). Measurement of an aeroacoustic dipole using a linear microphone
array, Journal of Acoustical Society of America 111(3), pp. 12671273.
Karamcheti K. (1955). Acoustic radiation from two-dimensional rectangular cutouts in aerodynamic surfaces, Tech.
Rep. TN 3487, NASA.
Karamcheti K. (1956). Sound radiation from surface cutouts in high speed ow, Ph.D. thesis, California Institute of
Technology.
Kerschen E.J., Cain A.B. (2008). Aeroacoustic mode trapping for a wind tunnel with a cavity in the wall, 39th
Plasmadynamics and Laser Conference, AIAA 2008-4213, Seattle, Washington.
Koschatzky V., Westerweel J., Boersma B.J. (2010). Comparison of two acoustic analogies applied to experimental
PIV data for cavity sound emission estimation, 16th AIAA/CEAS Aeroacoustic Conference, Stockholm, Sweden.
Lecordier B., Lecordier J.C., Trinite M. (1999). Iterative sub-pixel algorithm for the cross-correlation PIV measure-
ments, The third international Workshop on PIV 99, pp. 37-44, Santa Barbara, USA.
Ma R., Slaboch P.E., Morris S.C. (2009). Fluid mechanics of the ow-excited Helmholtz resonator, Journal of Fluid
Mechanics 623, pp. 126.
Marsden O., Bogey C., Bailly C. (2010). Numerical investigation of ow features and acoustic radiation around a
round cavity, 16th AIAA/CEAS Aeroacoustics Conference, 2010-3988, Stockholm, Sweden.
Marsden O., Jondeau E., Souchotte P., Bogey C., Bailly C., Juv`e D. (2008). Investigation of ow features and acoustic
radiation of a round cavity, 14th AIAA/CEAS Aeroacoustics Conference, 2008-2852, Vancouver, British Columbia
Canada.
Mechel F. (2002). Formulas of acoustics, Springer, 2002.
Mendelson R.S. (2003). Methods of measuring lock-in strength and their application to the case of ow over a
cavity locking into a single sidebranch, 9th AIAA/CEAS Aeroacoustics Conference, 2003-3106, Hilton Head, South
Carolina, USA.
108
REFERENCES
Mery F., Mincu D.C., Casalis G., Sengissen A. (2009). Noise generation analysis of a cylindrical cavity by LES and
global instability, 15th AIAA/CEAS Aeroacoustics Conference, 2009-3205, Miami, Florida.
Mincu D.C. (2010). Aeroacoustique des cavites cylindriques, Ph.D. thesis, Universite de Provence.
Mincu D.C., Mary I., Redonnet S., Manoha E., Larcheveque L. (2009). Numerical simulations of the sound generation
by ow over surface mounted cylindrical cavities including wind tunnel installation eects, 15th AIAA/CEAS
Aeroacoustics Conference, 2009-3314, Miami, Florida.
Morris S.C. (2010). Shear-layer instabilities: Particle Image Velocimetry measurements and implications for acoustics,
Annual Review of Fluid Mechanics 43, pp. 529550.
Nelson P.A., Halliwell N.A., Doak P.E. (1981). Fluid dynamics of a ow excited resonance, part I: experiment, Journal
of Sound and Vibration 78(1), pp. 1538.
Nomura Y., Yamamura I., Inawashiro S. (1960). On the acoustic radiation from a anged circular pipe, Journal of
the Physical Society of Japan 15, pp. 510517.
Norris A.N., Sheng I.C. (1989). Acoustic radiation from a circular pipe with an innite ange, Journal of Sound and
Vibration 135, pp. 8593.
Oshkai P., Yan T. (2008). Experimental investigation of coaxial side branch resonators, Journal of Fluids and Struc-
tures 24, pp. 589603.
Oshkai P., Yan T., Velikorodny A., VanCaeseele S. (2008). Acoustic power calculation in deep cavity ow: a semiem-
pirical approach, Journal of Fluids Engineering 130(5).
Panton R.L. (1990). Eect of orice geometry on Helmholtz resonator excitation by grazing ow, AIAA Journal 28,
pp. 6065.
Panton R.L., Miller J.M. (1975a). Excitation of a Helmholtz resonator by turbulent boundary layer, Journal of
Acoustical Society of America 58(4), pp. 800806.
Panton R.L., Miller J.M. (1975b). Resonant frequencies of cylindrical Helmholtz resonators, Journal of Acoustical
Society of America 57(6), pp. 15331535.
Parthasarathy S.P., Cho Y.I., Back L.H. (1985). Sound generation by ow over relatively deep cylindrical cavities,
Journal of Acoustical Society of America 78(5), pp. 17851795.
Phillips B. (1968). Eects of high-wave amplitude and mean ow on a Helmholtz resonator, Tech. Rep. TM X-1582,
NASA.
Plumblee H.E., Gibson J.S., Lassiter L.W. (1962). A theoretical and experimental investigation of the acoustic
response of cavities in aerodynamic ow, Tech. Rep. WADD-TR-61-75, Wright-Patterson Air Force Base, Dayton,
Ohio.
Powell A. (1964). Theory of vortex sound, Journal of Acoustical Society of America 36, pp. 177195.
109
REFERENCES
Rayleigh J.W.S. (1894). Theory of sound, vol. II, Macmillan, London, 2nd ed.
Rienstra S.W., Hirschberg A. (2011). An Introduction to Acoustics, Eindhoven University of Technology, April 28.
Rockwell D., Naudascher E. (1978). Review - Self-sustained oscillations of ow past cavities, Journal of Fluids
Engineering 100, pp. 152165.
Rockwell D., Naudascher E. (1979). Self-sustained oscillations of impinging free shear layers, Annual Review of Fluid
Mechanics 11, pp. 6794.
Rodrguez Verdugo F., Guitton A., Camussi R., Di Marco A., Grottadaurea M. (2010). Investigation of the ow and
the acoustics generated by a cylindrical cavity, 16th AIAA/CEAS Aeroacoustics Conference, 2010-3776, Stockholm,
Sweden.
Rona A. (2007). The acoustic resonance of rectangular and cylindrical cavities, Journal of Algorithms and Computa-
tional Technology 1-3, pp. 329355.
Roshko A. (1955). Some measurements of ow in a rectangular cutout, Tech. Rep. TN 3488, NASA.
Rossiter J.E. (1964). Wind-tunnel experiments on the ow over rectangular cavities at subsonic and transonic speeds,
Tech. Rep. R&M 3438, Aeronautical Research Council.
Rowley C.W., Williams D.R. (2006). Dynamics and control of high-Reynolds-number ow over open cavities, Annual
Review of Fluid Mechanics 38, pp. 251276.
Ruiz G., Rice H.J. (2002). An implementation of a wave-based nite dierence scheme for a 3-D acoustic problem,
Journal of Sound and Vibration 256, pp. 373381.
Selamet A., Dickey N.S., Novak J.M. (1995). Theoretical, computational and experimental investigation of Helmholtz
resonators with xed volume: lumped versus distributed analysis, Journal of Sound and Vibration 187(2), pp. 358
367.
Tang P.K., Sirignano W.A. (1973). Theory of a generalized Helmholtz resonator, Journal of Sound and Vibration
26(2), pp. 247262.
Tonon D., Hirschberg A., Golliard J., Ziada S. (2011). Aeroacoustics of pipe systems with closed branches, Interna-
tional Journal of Aeroacoustics 10, pp. 201276.
Ukeiley L., Murray N. (2005). Velocity and surface pressure measurements in an open cavity, Experiments in Fluids
38, pp. 656671.
Velikorodny A., Yan T., Oshkai P. (2010). Quantitative imaging of acoustically coupled ows over symmetrically
located side branches, Experiments in Fluids 48, pp. 245263.
Westerweel J., Dabiri D., Gharib M. (1997). The eect of a discrete window oset on the accuracy of cross-correlation
analysis of digital PIV recordings, Experiments in Fluids 23, pp. 2028.
110
REFERENCES
Yang Y., Rockwell D., Lai-Fook Cody K., Pollack M. (2009). Generation of tones due to ow past a deep cavity:
eect of streamwise length, Journal of Fluids and Structures 25, pp. 364388.
Ziada S., Ng H., Blake C.E. (2003). Flow excited resonance of a conned shallow cavity in low Mach number ow
and its control, Journal of Fluids and Structures 18, pp. 7992.
Ziada S., Shine S. (1999). Strouhal numbers of ow-excited acoustic resonance of closed side branches, Journal of
Fluids and Structures 13, pp. 127142.
Zoccola P.J. (2000). Experimental investigation of ow-induced cavity resonance, Ph.D. thesis, Catholic University
of America.
111
REFERENCES
112
List of publications
Journal Publications
1. Rodriguez Verdugo F., Guitton A., Camussi R., Experimental investigation of a cylindrical cavity in a
low Mach number ow, Journal of Fluids and Structures 28, pp. 1-19, January 2012.
Conference Proceedings
1. Rodriguez Verdugo F., Camussi R., Bennett G.J., Aeroacoustic source characterization technique ap-
plied to a cylindrical Helmholtz resonator, International Conference on Sound and Vibration, Rio de
Janeiro, 10-14 July 2011.
2. Rodriguez Verdugo F., Bennett G.J., Stephens D.B., Dynamics of the shear layer in the orice of a
cylindrical Helmholtz resonator using PIV, XVIII A.I.VE.LA. National Meeting, Rome, Italy, 15-16
December 2010.
3. Bennett G. J., Rodriguez Verdugo F., Stephens D. B., Shear layer dynamics of a cylindrical cavity
for dierent acoustic resonance modes, 15th Int. Symp. Appl. Laser Techn. Fluid Mech., Lisbon,
Portugal, 05 - 08 July 2010.
4. Stephens D. B., Rodriguez Verdugo F., Bennett G. J., Shear layer driven acoustic modes in a cylindrical
cavity, 16th AIAA/CEAS Aeroacoustics Conference, Stockholm, Sweden, 07 - 09 June 2010.
5. Rodriguez Verdugo F., Guitton A., Camussi R., Di Marco A., Grottadaurea M., Investigation of the
ow and the acoustics generated by a cylindrical cavity, 16th AIAA/CEAS Aeroacoustics Conference,
Stockholm, Sweden, 07 - 09 June 2010.
6. Verdugo F.R., Camussi R., Guitton A., Experimental characterisation of a cylindrical cavity in a low
Mach number ow, XX AIDAA Congress, Milano, Italy, July 2009.
7. Rodriguez Verdugo F., Guitton A., Di Marco A., Camussi R., Aeroacoustic characterization of a
cylindrical cavity, XVII A.I.VE.LA. National Meeting, Ancona, Italy, 26 - 27 November 2009.
113
List of publications
8. Rodriguez Verdugo F., Guitton A., Camussi R. Grottadaurea M., Experimental investigation of a
cylindrical cavity, 15th AIAA/CEAS Aeroacoustics Conference, Miami, USA, 11 - 13 May 2009.
114

Das könnte Ihnen auch gefallen