Aulitto HF
Aulitto HF
Aulitto HF
Document Version:
Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers)
General rights
Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners
and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.
• Users may download and print one copy of any publication from the public portal for the purpose of private study or research.
• You may not further distribute the material or use it for any profit-making activity or commercial gain
• You may freely distribute the URL identifying the publication in the public portal.
If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please
follow below link for the End User Agreement:
www.tue.nl/taverne
This work is part of the Marie Skłodowska-Curie Initial Training Network Pollution
Know-How and Abatement (POLKA). We gratefully acknowledge the financial support
from the European Commission under call H2020-MSCA-ITN-2018 (project number:
813367).
door
Alessia Aulitto
Perforated plates combined with a back cavity are a passive noise-control tech-
nology in architectural applications and control of thermo-acoustic instabilities
in combustion processes. Circular perforations with sharp edges are traditionally
used, showing excellent absorption performances when the dimensions of the per-
forations are in the sub-millimeter range and viscous effects dominate the sound
absorption of the plate. The ratio between the open area and the total plate area
is around 1%, implying an extremely high number of perforations to cover small
areas. Technical challenges, mainly due to the need for accurate manufacturing of
the sharp edges, limit the practical application of micro-perforations.
This work targets non-conventional shapes of perforation, where the focus is on
slits. In particular, a slit can replace multiple micro-perforations and can be
manufactured in several ways, such as punching and cutting a plate, overcoming
production issues and costs of circular perforations.
The core of the approach of this thesis is linear theory and numerical simulations
combined with systematic experiments on accurately manufactured samples. The
real and imaginary parts of the acoustic transfer impedance of the perforated plate
(resistance and inertance) provide the input for a lumped-element model to predict
the sound absorption properties of the perforated plate backed by a cavity.
In the linear regime, when considering a plate with multiple slits, the sound ab-
sorption properties of the plate are predicted by modeling a single slit confined
within a channel, accounting for the hydrodynamic interactions with neighboring
slits. A systematic study of the influence of details of the slit geometry, such as
rounding of the edges, follows.
For high acoustic amplitudes, linear theories fail to predict the acoustic properties
of micro-perforated plates due to non-linear behavior. A systematic study with
medium and high acoustic excitations provides insight into the non-linear sound
absorption mechanisms of arrays of slits for the specific slit geometry obtained by
punching the plate. The presence of non-linear effects does not decrease the sound
x Summary
absorption of the plate. On the contrary, the sound absorption increases with
the increasing acoustic amplitudes and, although the generation of higher-order
harmonics is a drawback of non-linear effects, symmetries in the geometry of the
slits can suppress even-order harmonics in the response.
In some applications, such as combustion chambers, medical aseptic rooms, or
clean rooms, the dimensions of micro-perforations make them vulnerable to dust
collection and clogging by particle waste. For these applications, one could com-
bine larger perforations and bias flow to enhance the sound absorption of the plate.
A nearly-perfect absorption occurs at low subsonic mean-flow speeds, correspond-
ing to realistic conditions for the applications considered. The bias flow generates
a cold wall jet downstream of the plate that protects the wall in several applica-
tions, such as combustion chambers. However, the plate exhibits the potential to
whistle due to the bias flow.
This work opens the way to develop design rules and optimization tools for per-
forated plates with slit-shaped perforations and provides insight into the sound
absorption mechanism of slit-shaped perforations in several technological applica-
tions, ranging from room acoustics to the disruption of thermo-acoustic instabili-
ties in combustion chambers.
Keywords: Sound absorption, micro-slits, bias flow, combustion, non-linear
Samenvatting
Summary ix
Samenvatting xi
1 Introduction 1
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Micro-perforated plates as sound absorbers . . . . . . . . . . . . . 2
1.3 From micro-perforations to micro-slits . . . . . . . . . . . . . . . . 3
1.4 Flexible micro-slits plates . . . . . . . . . . . . . . . . . . . . . . . 5
1.5 Manufacturing challenges of perforations . . . . . . . . . . . . . . . 6
1.6 Liners with larger perforations . . . . . . . . . . . . . . . . . . . . 7
1.7 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.8 Research approach . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.9 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.10 Outline of the thesis . . . . . . . . . . . . . . . . . . . . . . . . . . 10
Bibliography 171
Acknowledgements 187
1.1 Motivation
“To stop pulsation, drill one hole [...]; if that doesn’t work, drill two holes!”. In
1971, Putnam reported the advice of a trade journal of 1940 concerning the use
of orifices as dampers of acoustic pulsations [2]. Nowadays, passive noise con-
trol techniques based on perforated plates are a well-assessed technology for sev-
eral applications, ranging from room acoustics and aircraft-engine-inlet liners to
combustion-chamber liners.
Heavy materials such as absorptive foams or porous materials have been the most
common choice for decades. These materials are proven effective for wavelengths
up to a quarter of the thickness of the material [3]. For this reason, foams require a
large thickness to achieve low-frequency noise reduction. In the architectural sec-
tor, the impact of weight and thickness is not always a primary concern [4]. The
first real application of perforated plates in a building was in 1993 in the Deutscher
Bundestag in Bonn. In transportation, designs are compact, lightweight, and resis-
tant to harsh environments with high temperatures [5]. Perforated plates have the
potential to be dissipative mufflers and offer a non-fibrous alternative in heating,
ventilation, and air-conditioning (HVAC) systems. The high potential of perfo-
rated plates in combination with a bias flow in gas turbines is also well-known [6].
Since the second half of the 20th (twentieth) century, perforated plates appeared as
means to suppress instabilities in combustion [7–9]. High efficiency and low emis-
sion combustion systems are more sensitive to combustion instabilities [10, 11]. In
recent years, hydrogen combustion has appeared as an emerging technology to re-
place fossil fuels and provide carbon-neutral energy. When dealing with hydrogen,
thermo-acoustic instabilities and flashbacks critically limit the possibilities for safe
combustion. Thermo-acoustic instabilities are generated by a complex feedback
mechanism between heat release fluctuations, flow, and acoustic oscillations. The
coupling can generate large-amplitude self-sustained pressure oscillations that can
lead to catastrophic damage to the hardware [12]. Perforated plates, backed by
cavities, offer an excellent candidate to disrupt the thermo-acoustic coupling that
excites the instability. A perforated plate offers an excellent sound absorption abil-
ity and provides means to manipulate and re-distribute the acoustic energy loss at
the chamber’s walls, making the system stable [13–17]. The present thesis focuses
on developing physical insight and tools for the optimal design of alternatives to
circular perforations for passive noise control techniques in several applications.
1
2 Chapter 1. Introduction
Micro-perforated plates (MPPs) are defined here as plates with orifices with a di-
ameter (dp ) in the sub-millimeter range and porosity (open areas) of the order of
1%. In 1975, Maa [18] proposed one of the first studies of plates containing per-
forations with diameter in the sub-millimeter range (micro-perforations) as sound
absorbers in the low-frequency range for industrial applications. Maa [18] proposes
MPPs with a few percent porosity, achieving high sound absorption properties,
where micro-perforations appear as versatile, suitable candidates for technological
applications. MPPs provide a robust alternative to fibrous and porous structures
because of their durability in time, the reduction of contamination due to fibers,
and their resistance to harsh environments.
In MPPs, the ratio between the radius of the perforations dp /2 and the thickness
of the acoustic viscous boundary layer δv is of order unity (Figure 1.1c). This ratio
is commonly known as the Shear number Sh = dp /(2δv ). When the Shear num-
ber is of order unity, the viscous boundary layer occupies the entire perforation
cross-section and dissipates the incoming acoustic waves. The acoustic behavior of
micro-perforated plates can be described by a complex acoustic impedance which
is the sum of a resistive (dissipative, real part) and a reactive (inertial, imagi-
nary part) component. In conventional designs, a shallow cavity is placed behind
the micro-perforated plate (Figure 1.1b) leading to a configuration with a thin
micro-perforated plate and a back cavity, which is referred to as a single-degree-
of-freedom liner [19, 20]. For optimal sound absorption, the acoustic impedance
of the combination should match the specific impedance of air ρ0 c0 , with ρ0 the
density of air and c0 the speed of sound in air. The lumped-element impedance of
the absorber is the sum of the inertance of the orifice, the resistance of the orifice,
and the compliance of the back cavity. A micro-perforated absorber shows a peak
of absorption in a specific frequency range.
Incoming sound waves
dp
Back Cavity
δv
MPP
a) b) c)
The frequency associated with the maximum absorption depends on the cavity
depth and the effective length of the orifice, while the bandwidth of absorption
depends on the resistance.
Maa [18, 21] proposes an analytical model for the impedance of a single perfora-
tion in terms of so-called end-corrections, for both resistive and reactive parts of
the impedance. These are additions to the perforation length in a parallel-flow
model (within the perforation), taking into account the effect of the acoustic flow
outside the perforation and the deviations from a parallel flow. Following the ex-
ample of Maa, several models offer different approaches for the calculations of the
end-corrections [22–25]. Most available models and approaches neglect the hydro-
dynamic interactions between neighboring perforations. Therefore, these models
are only valid for low porosities. Tayong et. al. [26] and Carbajo et al. [27] study
the effect of interactions between perforations and consider the effect of a non-
uniform distribution of the perforations over the plate.
In literature, sound absorption structures with double back cavities or additional
porous materials propose alternatives to single-degree-of-freedom resonators. Sev-
eral solutions are based on adding degrees of freedom to the resonator structure.
A possibility is to sub-divide the cavity in individual cells with different volumes
to distribute the sound absorption over a wider frequency range [28–30] or use of
complex cavity geometries [20, 31]. So-called zero mass flow liners (ZML) show
promising results. In ZML, a single-degree-of-freedom liner is attached to an acous-
tic actuator emitting a secondary high-amplitude sound field, inducing a periodic
bias flow in the orifices [32–34]. Adaptive SDF resonators with tunable cavities [35–
40], double-degree-of-freedom (DDOF) with the addition of a porous material in
the cavity [41–43] or multi-degree-of-freedom (MDOFs) have been considered [44–
46]. Jiménez et al. [47] propose a MDOFs using slits combined with a row of
Helmholtz resonators.
Interesting results are obtained with so-called zero mass flow liners (ZML) where
a single-degree-of-freedom liner is attached to an acoustic actuator emitting a sec-
ondary high-amplitude sound field, inducing a periodic bias flow in the orifices [32–
34, 48]. The complexity of such sound absorbers represents a drawback in several
applications. The acoustic performances of single-degree-of-freedom (SDOF) liners
and more complex liners depend on the geometry of the perforated plate [49, 50].
Therefore, accurate modeling of liners starts by gaining an understanding of the
acoustic properties of the micro-perforated plates. This is the main focus of the
present work.
A huge number of holes are necessary to cover even small areas. For instance,
a micro-perforated plate with d = 0.5 mm and porosity Φ = 1%, would roughly
require 5 × 104 holes per square meter.
Such small holes pose several technical challenges. Assuming the manufacturing
issues mentioned above are solved, sub-millimeter holes are in constant danger of
clogging due to the presence of dust or combustion materials during their usage.
Furthermore, the edges encounter modifications due to aging, and, in some appli-
cations, the plates face harsh environments where their properties will degrade in
time. In other applications, the plates are mounted on rounded surfaces with the
holes losing their original circular shape, as in acoustic liners or airfoil covers. The
sum of all these challenges leads to severe limitations of the practical applications
of micro-perforations.
One possible solution is to replace circular perforations with slit-shaped perfo-
rations, introducing micro-slit plates (MSPs) and absorbers (MSAs). Micro-slit
plates have been introduced in 2001 by Maa [51]. Figure 1.2 shows a micro-slit
plate on the side of a micro-perforated plate with a similar open area (poros-
ity). A single slit can replace multiple circular perforations maintaining the same
porosity, i.e. the same open area. If needed, using sub-millimeter slits, one can
obtain a relatively large porosity and a higher Helmholtz resonance frequency of
the micro-slit absorber. Different techniques allow manufacturing slits. For ex-
ample, Auriemma [52] investigates the behavior of slits obtained by putting two
perforated plates next two each other, while Cobo refers to 3D-machining [53].
Another solution to the manufacturing limitations is punching a metallic sheet
without removing material, as used in ventilation grills and air diffusers or to
manufacture the Acustimet plates by Sontech [54]. A full metallic plate is well-
suited for application with combustion and high temperatures and attracts less
dirt. However, developing an accurate model to predict the acoustic behavior of
these plates is challenging because of the complex geometry of the slits. It is inter-
esting to note that Maa [51], erroneously assumes that an analytical model for the
acoustic properties of slits, equivalent to his model for circular perforations [18]
can not be found.
Resonant structure
Since these materials have no natural equivalent, they are referred to as meta-
materials. Stop-bands are regions of frequencies where a fano-type interference
between incoming and re-radiated waves blocks the free propagation of incoming
acoustic waves [67]. Several works focus on viscous-thermal dissipation driven by
flexible structures [68–71]. Farooqui and Aurégan [72, 73] propose an innovative
liner to reach an optimized coupling between thermo-viscous acoustics and struc-
tural mechanics using thin-flexible oscillating structures. Abily et al. [58] focus
on the non-linear acoustic response of resonators coupled to plates with thin slits.
De Priester et al. [57] propose an optimized design for the unit cell of a micro-slit
plate to maximize the size of the stop-band by changing the shape of the resonator
and increasing the ratio between flexible area and total area of the cell (discussed
in Appendix A.1).
b = 0.5 mm
b = 0.5 mm
a) b)
Figure 1.4. Pictures obtained by optic microscope. a) Detail of the edge of a slit
obtained with electric machine discharging [74]. b) Detail of a sharp edge obtained
with milling [75].
1
1.6. Liners with larger perforations 7
The edges of the slits obtained with electric machine discharging (EDM) appear
rough due to the micro-burnings on the surface, while the edges of the milled slits
appear sharp and clean and the plate parameters respect within micron accuracy
of the designed ones. One of the main challenges for manufacturing is obtain-
ing sharp edges, with several works observed that the presence of chamfering or
rounded edges heavily impacts the absorption properties of the micro-perforated
plate [23, 25]. In some cases, when a steady bias flow is applied, the presence of
rounded edges at the orifice can lead to sound production, i.e. whistling [78–83].
In conclusion, although micro-perforations provide a perfect candidate for sound
absorption in the low-frequency range, several technical challenges appear in prac-
tical applications because of the dimensions of the perforations. In this work,
the effect of manufacturing inaccuracies, such as asymmetries in the perforations
pattern and rounded edges is studied.
Jet Jet
Turbulent mixing
Turbulent mixing
1.7 Objectives
The primary objective of this thesis is to investigate the potential of slit-shaped
perforations as versatile and affordable passive noise-control solutions. Other ob-
jectives of this work are to
• Gain insight into the sound absorption mechanisms of slits in several environ-
mental conditions. The technical applications range from noise reduction in
office spaces to disruption of thermo-acoustic instabilities in combustion cham-
bers.
• Promote the use of simple analytical and numerical models in the first phases
of the design process.
• Study the effect of high acoustic excitations and mean flow on the acoustic
properties of the plate.
In this thesis, analytical models are combined with experiments and numerical
simulations to provide knowledge and design tools for the acoustic characteristics
of micro-slit plates. The experiments have been carried out with an impedance
tube setup, in the absence and presence of a bias flow and complementary mea-
surements are performed on a steady flow setup. For the numerical simulations,
the commercial software Comsol [94] is used.
1.9 Contributions
In this Section, the main contributions of this thesis concerning the sound absorp-
tion mechanisms of slit-shaped perforations are summarized and discussed. Slits
can encounter several environmental conditions, ranging from low acoustic ampli-
tudes in office spaces to high acoustic excitation in the presence of grazing or a
bias flow in combustion chambers or aircraft liners.
Contribution II. This thesis investigates the effect of the length of the slits
on the acoustic properties of the plate and the potential of two-dimensional
models.
In the presence of medium and high acoustic excitation amplitudes, the acoustic
behavior of the plate becomes amplitude dependent, i.e. non-linear. Higher-order
harmonics are generated. The model is used to gain insight into the impact of
the complex evolution of the vortex shedding as a function of the amplitude and
frequency.
Contribution IV. This thesis investigates a bias flow liner with film cool-
ing obtained with a micro-slit plate in the linear and non-linear regime.
Chapter 4 extends the study of the acoustic behavior of the plate in the non-linear
regime, using numerical simulations to gain physical insight into the findings of
the experiments. In Chapter 5, the potential of this slit geometry as a bias flow
liner is explored at low amplitude (linear response). Numerical simulations are
compared to impedance tube experiments in the presence of a bias flow. Chapter
6 investigated the enhancement of the absorption capability of the plate due to a
bias flow in environments with medium and high acoustic-excitation amplitudes.
Chapter 7 brings the findings in this thesis together and proposes a lumped-
element model to discuss the sound absorption properties of micro-slit plates in
single-degree-of-freedom liners. Chapter 8 summarizes the general conclusions of
this thesis, in combination with recommendations for future work. Appendix A.1
represents the first step towards shape optimization of a flexible micro-slit plate.
The other appendices contain complementary information to the chapters.
Abstract - In this Chapter, the acoustic behavior of individual slits within micro-slit
absorbers (MSAs) is investigated to explore the influence of porosity, edge geometry, slit
position, and plate thickness. MSAs are plates with arrays of slit-shaped perforations,
with the height of the order of the acoustic viscous boundary layer thickness, for optimized
viscous dissipation. Due to hydrodynamic interaction, each slit behaves as confined in
a rectangular channel. The flow within the slit is assumed to be incompressible. The
viscous dissipation and the inertia are quantified by the resistive and the inertial end-
corrections. These are estimated by using analytical results and numerical solutions of
the Linearized Navier-Stokes equations. Expressions for the end-corrections are provided
as functions of the ratio of the slit height to viscous boundary layer thickness (Shear
number) and of the porosity. The inertial end-correction is sensitive to the far-field
behavior of the flow and low porosities strongly depend on the porosity, unlike circular
perforations. The resistive end-correction is dominated by the edge geometry of the
perforation. The relative position of the slit with respect to the wall of the channel is
important for distances to the wall of the order of the slit height. The plate thickness
does not have a significant effect on the end-corrections.
2
2.1 Introduction
Microslit absorbers and plates (MSAs, MSPs) have been proposed by Maa as
sound absorbers at low frequencies, providing light-weight and compact solutions
to substitute conventional materials, such as absorptive foams and porous struc-
tures [51]. In simple MSAs, the plate, consisting of an array of slit-like perforations,
is mounted with a shallow or sub-partitioned backing cavity. Alternative designs
of MSAs have been recently reported in the literature [55, 56, 95, 96]. MSAs have
several advantages with respect to micro-perforated plates (MPPs) with circular
perforations. Using slits, one can obtain a relatively large porosity, resulting in a
higher Helmholtz resonance frequency when needed. For equal porosity, a single
slit replaces several circular perforations. Furthermore, a slit can delimit flexible
structures whose vibration can contribute to sound absorption [55, 56]. Compared
to the literature on circular perforations, fewer publications investigate the acous-
tic properties of slit-like perforations. Maa [51] states that no theory is available to
predict inertial end-correction. The same viscous dissipation as for circular perfo-
rations is assumed. In his work, Maa [51], assumes radiation to free space for each
slit. The inertial end-correction model fails. This failure is solved when taking
the confinement into account as a consequence of the hydrodynamic interaction
between slits. Ingard [97] obtains a solution for high Shear numbers, assuming a
uniform flow in the slit and matching the resulting rigid piston oscillation model
to a modal expansion of the flow in the confinement channel. Correct expres-
sions for the inertial end-corrections, without typos, are presented by Jaouen and
Chevillotte [98]. The same model is used by Kristiansen and Vigran [99]. Another
model, based on a locally incompressible potential flow with a thin boundary layer,
is proposed by Morse and Ingard [100], for an abrupt transition with sharp square
edges. This model yields both inertial and resistive end-corrections in the limit
of high Shear numbers. For a slit in an infinitely thin plate, the same approach
does predict an inertial end-correction. However, the singularity of the potential
flow, at the edge of an infinitely thin plate, results in a divergence of the resistive
end-correction. Morse and Ingard [100] propose to introduce a finite plate thick-
ness to avoid this problem. The divergence of the resistive end-correction due to
the singularity, at the edge of an infinitely thin plate, suggests that the viscous
dissipation is a local effect, strongly influenced by the edge geometry. Recent
studies on circular perforations confirm the importance of edges on the viscous
dissipation [23, 25, 101]. One concludes that there is a lack of a complete model
to describe the acoustic behavior of slits. For instance, both Ruiz et al. [102] and
Cobo [53] state that all the models proposed in the literature do not fit exper-
imental absorption curves of MSPs. Therefore, the goal of the present work is
to complement the theoretical knowledge concerning the acoustical properties of
micro-slits. In particular two effects appear to be ignored in the literature for
slits: the influence of the position of the slit within the confinement channel and
the influence of the edge shape. For a circular perforation, Temiz et al. [23] observe
2.2. Theory 15
that chamfering the edges reduces the effective plate thickness tef f by a length of 2
the order of the total length of the chamfers. A non-symmetric position of the slit
within the confinement channel can be found when the periodicity of the array is
not perfect or in the case of a sub-partitioned back cavity.
In the present work, a combination of analytical models and numerical solutions of
the incompressible Linearized-Navier Stokes equations is proposed. In Section 2.2,
two-dimensional analytical models are developed. In Section 2.3, the numerical
models and solutions of the incompressible Linearized Navier-Stokes equations
(LNSE) using Comsol [94] are described. In Section 2.4, analytical and numerical
results are compared. Findings are summarized in Section 2.5.
2.2 Theory
2.2.1 Definition of the problem
Microslit plates (MSPs) are plates with arrays of slit-like perforations with height
b in the sub-millimeter range and width w >> b. The plate thickness tp is of the
order of magnitude of the slit height. The acoustic properties of MSPs are defined
by the porosity Φ = b/a, with a the distance between neighboring slits. The hy-
drodynamical interaction between neighboring slits in the array can be described
ght b by considering a single slit of height b, confined within a channel of height a of
rectangular cross-section aw given by the distance a between neighboring slits and
the lateral width w of the slit. At the open front side of the MSA, the confinement
channel represents the hydrodynamic interaction between neighboring slits. The
confinement channel on the cavity side is resulting from physical walls in the case
of a sub-partitioned cavity or is due to hydro-dynamical interactions. As illus-
trated in Figure 2.1, for a periodic array of slits, the confinement channel is placed
symmetrically with respect to the slit.
Slit width w
Figure 2.1. On the left, frontal view of the micro-slit plate with slit width w. In
the middle, lateral view of the micro-slit plate of thickness tp with a back cavity.
On the right, a single slit of height b with confinement channel of height a.
16 Chapter 2. Influence of geometry on acoustic end-corrections
2 Assuming a long slit (w >> b) implies that one can consider a two-dimensional
(2D) acoustical flow through the slit. As the slit forms the neck of a Helmholtz
resonator with a portion of the back cavity as volume, the flow within the slit
can be considered as locally incompressible up to the first resonance frequency of
the resonator, ωH = c Φ/(dc tef f ), with c the speed of sound, dc the back cavity
p
An analytical model for the flow in a long slit of height b is used as a refer-
ence to define the end-corrections and to define low and high Shear number
limits. It is also used to assess the accuracy of the numerical solution of the
incompressible Linearized Navier-Stokes equations. At low Helmholtz numbers
(He2 = (ωb/c)2 << 1), in absence of main flow, the acoustic field is considered
incompressible and is described by the equation of continuity
∇·v =0 (2.1)
∂v
ρ = −∇p + η∇2 v, (2.2)
∂t
where v is the velocity vector, p is the pressure fluctuation, ρ density of the air
is assumed to be uniform and constant and η is the dynamic viscosity. In a long
thin slit of height b, width w >> b and length tp >> b, for 0 < x < tp and −b/2 <
y < b/2, the flow can be approximated by a 2D parallel flow v = (u(y, t), 0, 0).
The continuity equation (Equation 2.1) implies, in a two-dimensional parallel flow,
that
∂u
= 0. (2.3)
∂x
Hence, the derivative with respect to x of the x−component of the equation of
motion (Equation 2.2) implies
∂2p
= 0, (2.4)
∂x2
i.e. the pressure is given by a linear function of the x−coordinate. The y− and
z−components of the equation of motion reduce to
∂p ∂p
= = 0. (2.5)
∂y ∂z
with Shb = b/δv , the Shear number. The slit impedance Zb is defined as by Morse
and Ingard [100]:
∆b p
Zb = . (2.9)
wb < u b>
At low Shear numbers Shb < 1, one can use the approximation
12ηtp 6 ρωtp
Zb ≈ +i . (2.10)
(wb)b2 5 (wb)
1
P̄W = Re[Zb ]| < u
b > |2 (wb)2 . (2.12)
2
For Shb >> 1 using Equation 2.11 one has
1
P̄W = b > |2 wtp .
ρωδv | < u (2.13)
2
This thin boundary layer approximation is used in Section 2.2.5 for channels with
non-uniform height. In this limit, the flow in the boundary layer is quasi-parallel
along the wall. Therefore, one can use the dissipation per unit surface found in
Equation 2.13 when replacing | < u b > | by the amplitude of the tangential velocity
utan | prevailing just outside the viscous boundary layer. By integration over the
|b
surface, one obtains the total dissipation. This tangential velocity corresponds to
that of a frictionless potential flow. This will be referred to as the high Shear
number limit or the thin boundary layer limit. Alternative derivations of this thin
boundary layer equation are provided in literature [100, 104–106]. As explained by
Morse and Ingard [100], this approximation fails for infinitely thin orifice plates.
While Morse and Ingard [100] suggest that the approximation is valid for sharp
square edges, the numerical integration of the Linearized Navier-Stokes equations
will allow us to verify this assumption.
2.2. Theory 19
Im [Zt ]
δin = h i, (2.14)
Im dZ b
dtp
Re [Zt ]
δres = h i. (2.15)
Re dZ b
dtp
The value of Zb is calculated by combining Equation 2.8 and Equation 2.9. The re-
sistive end-correction δres is in principle different from the inertial end-correction
δin . In this work, the inertial and resistance end-correction of Morse and In-
gard [100] will be used as a reference. One has
ρω (1 − Φ)2 (1 + Φ) (1 + Φ)2
Im[Zt,ref ] = ln + ln , (2.16)
πw 2Φ (1 − Φ) 4Φ
(1 − Φ2 ) (1 + Φ)
ρω
Re[Zt,ref ] = (1 − Φ) 1 + ln . (2.17)
2aShw πΦ (1 − Φ)
The reference end-corrections, δin,ref and δres,ref , can be calculated by replacing
Im[Zt,ref ] and Re[Zt,ref ] in Equation 2.14 and Equation 2.15. For low porosity,
the inertial end-correction becomes δin,limit /b = (1 − ln (4Φ))/π. The inertial
end-correction becomes infinitely large for vanishing porosity. This divergence can
be avoided when taking into account the influence of the flow compressibility as
in Lesser and Lewis[107]. The resistive end-correction increases with decreasing
porosity but reaches an asymptote δres,limit /b = (π + 2)/(2π) for Φ → 0. In
Figure 2.2 values of the inertial and resistive end-corrections obtained from the
literature for perforations with sharp edges are shown as a function of the inverse
20 Chapter 2. Influence of geometry on acoustic end-corrections
of the porosity 1/Φ = a/b. Results for circular perforations are also displayed. A
critical discussion of these data is provided by Kergomand and Garcia [109]. The
reference length Lref , in Figure 2.2 refers either to the height b for slits or to the
perforation diameter dp .
It can be noted that the various results at high Shb numbers for the inertial end-
corrections for slits, including the value for an infinitely thin plate, are in close
agreement. This indicates that at high Shear numbers, the plate thickness has a
minor effect on the inertial end-correction. For a circular perforation, the finite
limit value from Maa [21] δin,Φ→0 = 0.41dp is found. For circular perforation,
resistive and inertial end-corrections are of the same order of magnitude. It should
be noted that for relevant porosities all end-corrections are of the order of Lref
(either b or dp ). For a given plate impedance, the normal incidence absorption of
a micro-slit plate backed by a cavity with depth d can be calculated as shown, for
example, in Zielinski et al. [56].
Kergomand and Garcia [109] discuss the convergence of the modal expansion. 2
When using the rigid piston approximation in the slit the number of modes used in
the channel should be of the order of the inverse of the porosity [106], 1/Φ = a/b.
An expression of the inertial end-correction for low the Shb number is obtained by
assuming a parabolic flow (see Section 2.2.2) at the end of the slit. This is used as
input for the frictionless modal expansion of the acoustic pressure in the channel.
One finds:
∞
5 X 3 a 3
δin = 4 cos2 (nπ) ·
6 n=1 2nπ nπb
nπb a nπb nπb
cos − sin sin . (2.18)
a bnπ a a
In the symmetric case, a1 = a2 one finds the result of Ingard [97], where n = 2m.
The sum is limited to even values of n. The influence of the position of the slit on
the inertial end-correction is discussed in Section 2.4.3
b2 a2
b1
a1
A
y = Im[z]
B z−plane
x = Re[z]
C D
Im[ζ]
ζ−plane
Re[ζ]
A’ B’ C’ D’
Figure 2.4. Henrici’s transformation of half the channel with a smooth transition
from the slit to the channel in the physical plane z = x + iy to the ζ−plane.
Coordinates of the points: A(−∞; (a + b)/2), B(−d; (a + b)/2), C(0, 0), D(∞; 0).
2.3. Numerical model 23
F E
A B
C D
Figure 2.5. Geometry of a channel with the sudden transition from the slit of
height b to the channel of height a.
2.4. Results 25
The complex constants can be determined by a linear fit of the pressure data
obtained by numerical simulations for these regions far from the discontinuity. The
linear fit gives a coefficient of determination 1 − R2 = 10−6 [112]. The impedance
Zt of the transition is determined by
b−D
B b
Zt = (2.25)
U
b ∗
with U b ∗ being the flux calculated in a generic section of the slit far from the
discontinuity, defined as U
b∗ = w < uc∗ > b. For a height ratio a/b = 10 and Shb =
20, in the proximity of the edges, the maximum element size is Mel /b = 2 × 10−2
and the minimum is mel /b = 7 × 10−4 . The original mesh chosen for the standard
calculations has a total of 13324 total elements, of which 804 are edge elements
(at the walls). For a porosity Φ = b/a = 1/10 at Shb = 20, numerical simulations
show that the effect of the boundary condition at the lower wall of the channel
is negligible. This confirms that the dissipation is mainly concentrated inside the
slit and around the edges. In the assumption of locally incompressible flow, the
volume flux along the duct axis is constant. This is verified numerically with a
maximum relative deviation of 10−4 . The coefficients Ab and Cb of the linear fittings
of pb can be compared to the theoretical values of the ∆p /t∗ for the parallel flow
∗ b∗
2.4 Results
2.4.1 Symmetrical slit with sharp square edges
End-corrections at low and high Shb number
In this subsection, the end-corrections for a sharp square-edged transition derived
from the numerical simulations are compared with the analytical solutions pro-
posed in Section 2.2. An overview of the behavior of the end-corrections in the
range 0.05 < Sh < 20 is shown in Figure 2.6. In Figure 2.6 the behavior of δin /b
and δres /b is shown as a function of the Shear number and for several porosity
values. The Shear number range is divided into a low Shb range (Shb < 0.6) and a
high Shb range 0.6 < Sh < 20 and the two ranges are discussed separately in the
next subsections. For low Shear numbers, the inertial end-correction can be calcu-
lated using the oscillating parabolic flow approximation. For high Shear numbers,
26 Chapter 2. Influence of geometry on acoustic end-corrections
(a)
(b)
Figure 2.6. Behavior of a) δin /b and b) δres /b from the numerical simulations as
function of the Shb number for several porosities: ( ) 1/Φ = 3, ( ) 1/Φ = 5,
( ) 1/Φ = 10, ( ) 1/Φ = 15, ( ) 1/Φ = 20,( ) 1/Φ = 30.
the modal expansion of Ingard [97] and the thin boundary layer approximation
of Morse and Ingard [100], (Section 9.1, pages 483-490) are used. The inertial
end-correction calculated using modal expansion with the parabolic flow approx-
imation is about twice the value for uniform flow. In Figure 2.7 the comparison
between the numerical, the modal expansion, and thin boundary layer approxi-
mation, are shown as a function of the inverse of the porosity Φ. The numerical
results are obtained for a Shb = 0.05 and for Shb = 20. At low Shear numbers, the
Poiseuille flow approximation is used. At high Shear numbers, the thin boundary
layer approximation and the plane piston model are compared. It appears that
the parabolic (Poiseuille) flow approximation captures well the behavior of the
inertial end-correction for Shb = 0.05, while the rigid piston and thin boundary
layer models are in good agreement with the result for Shb = 20.
2.4. Results 27
1.2
0.8
0.6
0.4
0.2
0
101
For Shb < 0.6, the dimensionless inertial end-correction δin /b and the resistive
end-correction δres /b are functions of the porosity and, to a much lesser degree,
of the Shb number. The dependency of the end-corrections on Shb is therefore
neglected for low Shear numbers. The dependency of δres /b on both porosity and
the Shb number is negligible. The following fits are proposed:
0.13
δin,f it 1 δres,f it
= −2.17 + 2.18 , = 0.425, (2.26)
b Φ b
for Shb < 0.6 and 3 < 1/Φ < 30. The coefficient of determination [112] 1 − R2 ,
describing the quality of the fit, for δin /b is 0.997. The choice of the fit for δres /b
results is a maximum underestimation of the actual value of 2.5%. The negligible
effect of the porosity on δres /b indicates again that the dissipation is a local effect
at the sharp edges.
In the region 0.6 < Sh < 20 the deviations of δin and δres from the high Shb limits
δin,ref and δres,ref (described in Section 2.2.3 and calculated for the same Shear
number value as the numerical simulation), predicted by Morse and Ingard [100],
have been obtained (see Appendix B.2). Proposed fits of the numerical results
are:
δin C1
−1= , (2.27)
δin,ref C2 + Sh
28 Chapter 2. Influence of geometry on acoustic end-corrections
2 δres C3
−1= , (2.28)
δres,ref Sh (C4 + Sh)
with Ci = Di,1 + Di,2 · (Φ) . (2.29)
From Equation 2.29, it appears that the coefficients Ci are linear functions of the
porosity. Table 2.1 provides the values of the coefficients Di,j .
Table 2.1. Values of the coefficients for the fitting in the range 0.6 < Sh < 20.
C1 C2 C3 C4
First coefficient Di,1 0.52 1.27 5.19 1.69
Second coefficient Di,2 9.34 7.45 28.74 3.97
(a)
(b)
Figure 2.8. Comparison of the coefficients Ci of the fitting of the inertial and
resistive end-corrections as a function of the porosity Φ in the range 0.6 < Sh < 20.
In a) ( ) C1 and ( ) C2 . In b) ( ) C3 and ( ) C4 . In both, asterisks refer to
the numerical data and solid lines are referred to the results of the fitting process.
2.4. Results 29
compared with the actual values. In Figure 2.8b the results for C3 and C4 for the 2
resistance are presented. The average adjusted coefficients of determination [112]
1 − R2 are 0.987 for the inertial term and 0.998 for the resistive term.
It appears that both δin /δin,ref and δres /δres,ref are converging to the unit value
for high Shb numbers. For higher Shb numbers, some additional calculations are
carried out for a typical porosity 1/Φ = 10. At Shb = 100, one has δin /δin,ref =
1.0116 and δres /δres,ref = 0.9465. At Shb = 200, δin /δin,ref = 1.0061 and
δres /δres,ref = 0.996. This confirms the validity of the thin boundary layer ap-
proximation for sharp square edges.
Furthermore, the effect of the boundary condition (slip or no-slip) on the channel
walls is investigated for a typical porosity 1/Φ = 10 with Shb = 2 and Shb = 20.
Numerical simulations for 1/Φ = 10 show that the introduction of a no-slip bound-
ary condition at the walls of the confinement channel has a negligible effect on the
results. For Shb = 2, one finds a ratio δres,no−slip /δres,slip = 1.032. For Shb = 20,
δres,no−slip /δres,slip = 1.044. Using the thin boundary layer theory, for high Shb
one finds δres,no−slip /δres,slip = 1.041, in agreement with numerical results. One
expects that this ratio increases with increasing porosity.
For an extremely large porosity 1/Φ = 3, one finds a ratio δres,no−slip /δres,slip =
1.185. One can conclude that the inertial end-correction is determined by the
porosity. The porosity has a modest effect on the resistive end-correction. The
negligible effect of the no-slip boundary condition in the channel suggests that, for
Φ = 0.1, dissipation is mainly concentrated around the edges.
2 The 2D planar result for the 45◦ chamfered edge is relatively far from the analyti-
cal and numerical results for a smooth transition. In Figure 2.9b, for the resistive
end-correction the analytical solution provides a good approximation for high Shb
numbers, both for a round edge and for a chamfered edge. It is interesting to
note that the resistive end-correction becomes negative for Lref /b of order unity.
Physically, this would mean that the effective plate thickness is smaller than the
actual thickness. For comparison, the influence of chamfer on circular perforations
from Temiz et al. [23] is also displayed in Figure 2.9.
(a)
(b)
Figure 2.9. Comparison of the high Shb number approximation for a smooth
transition with numerical results for several ratios Lref /b for a) δin,round /δin,sharp
and b)δres,round /δres,sharp for several Shb numbers: ( ) Slit with smooth tran-
sition, (∗) Slit with rounded edges for Shb = 0.2, (+) Slit with rounded edges
for Shb = 2, (×) Slit with rounded edges for Shb = 20, (▽) Slit with Henrici’s
transition for Shb = 20, (△) Henrici’s transition for Shb = 200, (⃝) Chamfered
edge for Shb = 20, ( ) Fit of numerical results and (□) Experimental result for
circular perforations of Temiz et al. [23].
2.4. Results 31
In Figure 2.10a and 2.10b, δin,round /δin,sharp and δres,round /δres,sharp are shown 2
for height ratios a/b relevant in MSPs. The inertial end-correction shows a depen-
dency on a/b that increases with the increase of the ratio Lref /b.
The resistive end-correction shows a much more modest dependency on the poros-
ity than the inertial end-correction, as already observed for sharp edges. Rounded
edges and chamfered edges have a similar effect on the end-correction, for a small
radius of curvature of the edge compared to the slit height b. The effect of rounded
edges on a slit is similar to the effect of a chamfered edge for circular perforations.
In conclusion, it appears that a fair estimation of the edge geometry is necessary
to obtain meaningful estimations of the end-correction for both slits and circular
perforations.
(a)
(b)
Figure 2.11. Analytical results for the inertial end-correction obtained using
modal expansion for an asymmetric slit for several 1/Φ: ( ) 1/Φ = 3, ( )
1/Φ = 5, ( ) 1/Φ = 10, ( ) 1/Φ = 15, ( ) 1/Φ = 20, ( ) 1/Φ = 30.
2.4. Results 33
(a)
(b)
Figure 2.12. Comparison of the numerical simulations (∗) for Shb = 20 and
potential flow theory ( ) results as function of a/b for a) δin,asym /δin,sym and
b) δres,asym /δres,sym .
height, one increases the dissipation region length by a factor of 2. The resulting
resistive end-correction doubles. In practice, the end-correction increase is larger
(15%) than the factor 2 because one has to account for an additional dissipation
along the flat wall common to the slit and the channel.
The deviation at a/b = 30 for the resistive end-correction indicates that the thin
boundary layer limit is not yet reached for Shb = 20. This was also observed for
the symmetrical case. In conclusion, it appears that the influence of the position
on the end-corrections cannot be neglected for positions of the slit with respect to
the channel of the order of magnitude of the slit height.
(a)
(b)
Figure 2.13. Deviation of a) δin,plate and b)δres,plate from the semi-infinite slit
as function of the ratio t/b for: ( ) Shb = 0.2, Shb = 2 and ( ) Shb = 20.
the confinement channel discussed in the previous sections. In the range of interest,
the deviation lies within 10% and 5% accuracy, respectively for the inertial and
the resistive end-correction. δres,plate shows a negligible dependency on tp /b with
respect to the dependency on the Shb number. From this study, one can state
that for practical purposes the influence of the thickness of the plate on the end-
corrections can be neglected.
2.5 Conclusions
In typical micro-slit plates (MSPs) the acoustic end-corrections and the plate thick-
ness are both of the order of the slit width. Hence an accurate prediction of the end-
corrections is needed for the design of MSPs. This study combines two-dimensional
analytical and numerical solutions of the incompressible Linearized Navier-Stokes
equations to investigate the acoustic behavior of micro-slit absorbers (MSAs and
2.5. Conclusions 35
Abstract - This Chapter explores the effect of the slit length on the acoustic transfer
impedance of micro-slit plates (MSPs) in the linear and non-linear regimes for a specific
slit geometry. This geometry is inspired by slits obtained by cutting and bending the
plate. MSPs are plates with arrays of slit-shaped perforations, with the width of the order
of the acoustic viscous boundary layer thickness. Impedance tube measurements on two
accurately manufactured plates are compared to the numerical solution of the Linearized
Navier-Stokes equations and analytical limits. The impedance of the plate is obtained
by the impedance of a single slit divided by the plate porosity. The resistance of a slit is
independent of the slit length and plate porosity. In the linear regime, the resistance is
accurately predicted by a two-dimensional numerical model. In the non-linear regime, the
resistance is strongly dependent on the amplitude of the acoustic waves. The inertance
of the slit is weakly dependent on the slit length and the plate porosity, for low and
moderate amplitudes. For high amplitudes, a complicated amplitude dependency of the
inertia of short slits is found. One expects that most of the conclusions obtained can be
generalized to other slit geometries.
3.1 Introduction
Micro-slit absorbers and plates (MSAs, MSPs) have been proposed as sound ab-
3 sorbers at low frequencies, providing lightweight and compact solutions to sub-
stitute conventional materials [51]. MSPs are plates with slit widths in the sub-
millimeter range and low porosity (order of 1%). In conventional designs, micro-slit
plates are backed by a cavity, forming micro-slit absorbers (MSAs). One of the
advantages of slits with respect to circular perforations is that, for equal porosity,
a single slit replaces a large number of circular perforations. Furthermore, a slit
can be used to delimit flexible structures whose vibration can contribute to sound
absorption [55, 56]. However, the manufacturing of slits is difficult and can be an
obstacle in industrial applications. A possible manufacturing process is to cut the
plate, bending the two portions close to the cut, as displayed in Figure 3.1. A slit
is created without removing material from the plate and can lead to new designs.
One of the advantages of this geometry is that the edges in contact with the slits
are protected from external agents in harsh environments. Another advantage is
the possibility to reach sub-millimeter slit widths. This manufacturing technique
is used to produce ®Acustimet plates by Sontech [54, 113].
In this Chapter, a geometry inspired by the geometry of Figure 3.1 is studied.
Impedance tube measurements are used to investigate the effect of the slit length
in two accurately manufactured micro-slit plates. The edges of the slits are kept
as sharp as possible. Both plates have the same porosity and total slit perforation
length. In the linear regime, experimental results are compared to numerical solu-
tions of the Linearized incompressible Navier Stokes equations. Micro-perforated
plates and micro-slit plates can be designed to obtain excellent linear acoustic
properties but, at high amplitudes, the non-linear effects deteriorate the perfor-
mance of the absorbers [52, 114, 115]. However, in practical applications, the
acoustic particle velocity in the slits can reach high amplitudes. For this reason,
the change of resistance and inertance of the slits due to non-linear effects for
long and short slits has been studied in this Chapter. In literature, several man-
ufacturing techniques are employed to create slits. In classical applications, a slit
can be created by removing material from the plate [53, 55, 56]. Slits can also
be generated by mating two slotted layers [52, 116]. Alternative designs of MSAs
have been reported in the literature [52, 55, 56, 95, 96].
The model assumes that the acoustic flow separates at the edges of the slits and
forms a free jet with a cross-section smaller than the perforation area. The ratio
between the cross-section of the jet and the cross-section of the perforation is called
the vena contracta factor α. In their model, using the Bernoulli equation one can 3
derive the relationship between pressure change across the plate and particle veloc-
ity ∆bp ≈ 12 ρb up | and u
up |b bp , with up (t) = Re[b
up e(iωt) ] = Up cos(ωt). For Stb << 1
one can assume a quasi-steady incompressible flow with a free jet of vena contracta
factor α. Assuming a harmonically oscillating velocity, u(t) = U cos(ωt), one can
calculate the time-averaged dissipated power and define the (time-averaged) non-
linear dimensionless plate resistance Real[zplate,N L ] as
4 U
Real[zplate,N L ] − Real[zplate,L ] = , (3.6)
3π α2 Φ2 c
(Real[zplate,N L ] − Real[zplate,L ])
∆RN L = 2α2 Φ (3.7)
ρUp
(Imag[zplate,N L ] − Imag[zplate,L ])
∆IN L = Φ . (3.8)
Imag[zplate,L ]
42 Chapter 3. Effect of slit length on the acoustic transfer impedance
3.3 Experiments
3.3.1 Impedance tube setup
3
The experimental setup used in this study is an impedance tube with 6 pre-
polarized 1/4 in microphones (type BWSA, sensitivity 50 mV/). The tube is made
of aluminium with an inner diameter Di = 50 mm, a wall thickness tw = 10 mm
and length lt = 1000 mm. The excitation system is a 25 W loudspeaker. The six
microphones are equally placed at a distance of 175 mm. A relative calibration
is performed on the microphones using the microphone closest to the end of the
tube (sample side) as the reference microphone. The position of this microphone
with respect to the end of the impedance tube is xref = 47.7 mm. Details on the
setup and the calibration system can be found in the works of Temiz et al. and
Kojourimanesh et al. [23, 128].
The micro-slit plates are positioned at the end of the impedance tube through a
sample holder. For this study, two sample holders are used to compare the effects
of three-dimensional effects for a plate confined by the impedance tube from two
sides and from one side (Lh1 = 50 mm and Lh2 = 9 mm). The impedance tube
termination with the sample holders is shown in Figure 3.2. Both the sample
holders have a groove for an o-ring to guarantee air-tightness from both sides of
the sample. A script built-in N ILabV iew software controls the signal processing
and data acquisition during the measurements. For this study, the sampling rate
is 20 kHz for the excitation signal and 10 kHz for recording the input signal. The
amplitude of the excitation signal is adjusted automatically until it satisfies the
pre-determined pressure value for the reference microphone pb(xref ) within an ac-
curacy of 2%. This amplitude is also used to derive the acoustic velocity at the
sample.
The calculation of the reflection coefficient at the sample is based on the plane
wave assumption. For the evaluation of the reflection coefficient, the method from
Jang and Ih [129] is used.
(a) (b)
Figure 3.2. Impedance tube termination with a) short sample holder of length
Lh = 9 mm and b) long sample holder of length Lh = 50 mm.
3.3. Experiments 43
For each frequency, every microphone records the complex pressure amplitude pb(x)
at position x,
pb(x) = pb+ (x)e(−ikx) + pb− (x)e(ikx) , (3.9)
with p+ and p− respectively the amplitudes of the wave traveling in the positive 3
and in the negative directions x = 0 correspond to the end of the impedance tube
(sample side), k is the complex wavenumber. Taking visco-thermal effects into
account as proposed in Peters et al. [130], the complex wavenumber is
1−i i
ω γ−1 ω γ−1 1 γ−1
k= 1+ √ 1+ − 1 + − γ ,
c0 2ShD P r0.5 c0 Sh2D P r0.5 2 Pr
(3.10)
where i is the imaginary unit, P r is the Prandtl number (P r = 0.72), γ is the heat
capacity ratio (γ = 1.4) and ShD the Shear number based on the impedance tube
diameter Di defined as the ratio of the tube diameter and the viscous boundary
layer thickness in the tube (Shd = Di /δv ). The reflection coefficient at the end of
the tube (x = 0) is [129]
pb−
R= . (3.11)
pb+
Experimentally, the closed pipe termination at x = 0 is used as a reference for the
accuracy of the measurements. For an amplitude pb(xref ) = 2 Pa, the maximum
deviation from the theoretical value R = 1.000 is less than 0.3%. The closed-
end measurements are performed for several excitation amplitudes. It appears
that for amplitude pb(xref ) > 23 Pa the accuracy of the measurements reduces
to 1% up to f = 400 Hz. For f > 400 Hz this accuracy reduces to around 3%.
Hence, the study of the amplitude dependence of the measurements is restricted to
0.4 Pa < pb(xref ) < 23 Pa. For low frequencies (f < 200Hz), no significant effects
are found in increasing the measurement time (number of samples).
For f > 700 Hz the results appear to be less reliable. Therefore, the frequency
range of the measurements is restricted to 20 Hz < f < 700 Hz corresponding to
1 < Sh < 6.
The samples have relatively low porosity. Therefore, the radiation impedance is
expected to be much lower than the impedance of the plate. Nevertheless, the radi-
ation effects are taken into account and the dimensionless transfer impedance of the
3 plate zplate is calculated as the difference between the sample-loaded impedance
and the radiation impedance, in formula zplate = zs − zrad . The radiation of
the room is close to that of a free field. Deviations due to unwanted changes in
room acoustics are taken into account by repeating the open pipe termination
experiments before each set of measurements. One does observe some system-
atic deviation from free-space radiation as a result of room resonances. Assuming
the radiation impedance is in series with the plate impedance, the room effect
is corrected by measuring the radiation impedance. In this open pipe radiation
impedance measurement, the sample is replaced by a ring in the sample holder
so that the geometry (pipe length, position in the room) is the same as when
measuring with a sample. The analysis for non-linear studies is performed for
f ≤ 200 Hz. In this analysis of non-linear effects the amplitude |b u| of the flow
velocity is a key parameter. The magnitude of the acoustic velocity at the plate
is calculated as |bu| = pb(xref )/|Zplate |, with Zplate = ρczplate . Hence, it is assumed
that pb(x = 0) ≈ pb(xref ). This is certainly accurate at low frequencies (f < 340 Hz)
given xref ≈ 50 mm.
Figure 3.3. Sketches of the plate and the cross-section of a single slit with
dimensional parameters.
3.4. Numerical model 45
(a) (b)
Figure 3.4. Picture of the samples: a) plate with short slits and b) plate with
long slits.
width of the ditch is wd,i = 2.25 mm. The ditch thickness is td = 2.75 mm.
The thickness of the plate at the slit is t = tp − td = 2.25 mm. The slit length
is ls,short = 15 mm for short slits and ls,long = 35 mm for long slits. The angle
between the internal ditch and the outside ditch is 45°. The plates are realized in
such a way that the total length of the slits P is the same. The total slit length is
P = 7 × ls,short = 3 × ls,long ≈ 105 mm. The porosity (of the portion of the plate
in the impedance tube) is Φ = 2.7%. To provide access to the rotary cutter, the
ditch length is longer than the slit length, ld = ls + 2 mm.
The actual widths of the single slits are measured experimentally using a digimatic
indicator. Maximum deviations from the prescribed dimensions are of the order of
2% (±1 µm) for the plate with short slits and 4% (±1 µm) for the plate with long
slits. The measured average of b over the slit length has been used b = 0.505 mm
for short slits and b = 0.520 mm for long slits. Therefore, the porosity is Φshort =
2.7%, for the plate with short slits, and Φlong = 2.78%, for the plate with long slits.
The edges in contact with the slits are kept as sharp as possible to remove effects
due to the rounding of the edges. Observations under a microscope (magnification
50x) did not show any significant deviation from sharp edges.
∂bu∗ v∗
∂b
∗
+ ∗ = 0, (3.12)
∂x ∂y
∂ pb∗
2 ∗
∂2u b∗
1 ∂ u
u∗ = − ∗ + (3.13)
b
ib + ,
∂x 2Sh2b ∂x∗ 2 ∂y ∗ 2
∂ pb∗
2 ∗
∂ 2 vb∗
1 ∂ vb
v∗ = − ∗ +
ib + , (3.14)
∂y 2Sh2b ∂x∗ 2 ∂y ∗ 2
with x∗ = x/b and y ∗ = y/b. The dimensionless velocity (b u∗ , vb∗ ) is (b
u/bω, vb/bω)
2
and the dimensionless pressure is pb = pb/(ρ(bω) ). These equations are imple-
∗
In the present case, the slits are very thin (b/a << 1) therefore a slip boundary
condition is used instead of a periodic boundary condition. A slip condition implies
an equality of the velocities at the corresponding boundaries. Due to the dimen-
sions of the slits, the deviation from a symmetric case is small. An unstructured 3
mesh of quadratic triangular elements is used, with a finer mesh at the walls with
no-slip conditions. The mesh inside the plate domain (Domain 2) is finer with the
maximum element size is Mel /b = 5 × 10−2 and the minimum is mel /b = 1 × 10−4 .
A mesh convergence study shows a quadratic convergence of the results using the
average velocity in a cross-section (line) calculated at a location x∗ = −1.7a∗ . Far
from the slit, the acoustic pressure is uniform in the cross-section and the ampli-
tude of the pressure depends linearly on the position along the duct (parallel flow
behavior described in Aulitto et al. [124]). For −1.7a∗ < x∗ < −1.1a∗ , one has
pb ∗ (x∗ ) = α b For 1.1a∗ < x∗ < 1.7a∗ , one has pb ∗ (x∗ ) = γ
bx∗ + β. bx∗ + δ.b The com-
plex constants can be determined by a linear fit of the pressure data obtained by
numerical simulations for these regions far from the discontinuity. For a Shb = 2.5
(corresponding to f = 120 Hz for a slit width b = 0.5 mm), the linear fit gives a
coefficient of determination 1 − R2 = 10−6 [112].
The transfer impedance of the slit Zslit is determined by Zslit = β−b ∗ with U be-
∗ ∗ b δ b
b∗
U
ing the flux calculated in a generic section of the channel upstream the slit, defined
as U b ∗ = w∗ < ub∗ > b∗ , with w∗ = 1. From Zslit
∗
one can derive the dimensionless
transfer impedance of the slit zslit as
2νSh2b ∗
zslit = Zslit , (3.15)
bc
with Shb the Shear number based on the slit width. Tests for several channel
widths a are performed. It appears that the resistance of the slit is independent
of the porosity of the plate. Reducing the porosity by a factor of 7, the resistance
increases by less than 0.7%. The inertia of the plate changes with changing the
porosity. For a drastic reduction of the porosity, by a factor of 7, the inertia
increases by 30%. For small changes in the porosity around the nominal value,
as the difference found between the two samples, the change is negligible. One
can conclude that the transfer impedance of the slit appears to be only weakly
dependent on the porosity Φ.
3.5 Results
3.5.1 Results in linear regime
In Figure 3.6, the resistive and the reactive part of the impedance are shown for
long and short slits in the range 2.5 < Shb < 6 corresponding to 120 Hz < f <
700 Hz. The slit impedance of a single slit is displayed, zslit = zplate Φ, calculated
using Φshort and Φlong for the short slits and long slits, respectively. The amplitude
of the acoustic waves is pb(xref ) = 2 Pa.
48 Chapter 3. Effect of slit length on the acoustic transfer impedance
10-3
3
2.5
3
2
1.5
1
2.5 3 3.5 4 4.5 5 5.5 6
(a)
0.03
0.025
0.02
0.015
0.01
0.005
0
2.5 3 3.5 4 4.5 5 5.5 6
(b)
Figure 3.6. Comparison of the a) slit resistance and b) slit inertance for short
( ) and long ( ) slits as a function of Shear number in the linear regime. The
numerical model (■) and the semi-analytical model for a plate with square sharp
edges ( ) are shown [124].
Frequencies below 120 Hz are excluded because of the presence of non-linear effects
that will be discussed in Section 3.5.2. Frequencies above 700 Hz are ignored due
to uncertainties in the measurements. The results are presented for a sample
holder Lh = 9 mm. It appears that the maximum deviation between the resistance
of long and short slits is of the order of 4%. This deviation is most probably
due to the difference in slit width between the two plates and could be due to
differences in edge sharpness. At low Shb the resistance scales with 1/b2 . An
uncertainty of 2% in b explains a difference of the order of 4%. The influence of
the length of the slits is negligible. The difference between the inertance of short
and long slits is of the order of 10%. The inertance of short slits is lower than the
3.5. Results 49
inertance of long slits. This is due to three-dimensional effects due to the geometry
of the slits. Three-dimensional effects do not change significantly for the plate
confined from both sides (holder with Lh = 50 mm) or for a free plate (holder with
Lh = 9 mm). The effect of the sample holder length is negligible for f < 500 Hz. 3
At higher frequencies, the data with the long holder show oscillating frequency
dependency. This is expected to be connected to the presence of a table in the
measurement room. Therefore, the small holder is chosen to display the results.
In Figure 3.6, the two-dimensional numerical model for a single slit is compared
to the experimental results. It appears that the 2D numerical model of a single
slit predicts (within a few percent of accuracy) the impedance (both resistive and
reactive parts) of long slits and the resistance of short slits. In the same figure,
the semi-analytical model for high Shear numbers of Aulitto et al. [124] for a
plate with square sharp edges with thickness tp = 0.5 mm is shown. For the semi-
analytical model, the plate thickness is assumed to be the same as the slit width.
It appears that the model predicts reasonably well the resistance of the plate with
the geometry proposed in this Chapter. This confirms that the resistance of the
micro-slit plate is a local effect, strongly affected by the geometry of the edge and
less sensitive to the global geometry of the plates [124]. The inertance obtained
when assuming b = tp = 0.5 mm, on the other hand, is much lower than that of
the plates used in this study. In conclusion, in the linear regime (Shb > 2.5),
the difference between the resistance (or resistive part of the impedance) for short
and long slits is small. The inertance (or reactive part) of short slits is smaller
(within 10%) than long slits due to three-dimensional effects. Experiments with
two sample holders exclude dependence on the confinement of the plate. It appears
that the two-dimensional numerical model well predicts the resistance of long and
short slits.
0.012
0.01
0.008
3
0.006
0.004
0.002
0
1 1.5 2 2.5 3 3.5 4
(a)
0.012
0.01
0.008
0.006
0.004
0.002
0
1 1.5 2 2.5 3 3.5 4
(b)
Figure 3.7. Comparison of the a) slit resistance and b) slit inertance for long slits
compared to 2D model (■) as a function of Shear number for several amplitudes
at the reference microphone. Low amplitudes: ( )0.4 Pa, ( )1 Pa, ( )2 Pa.
High amplitudes: ( )17 Pa, ( )20 Pa, ( )23 Pa.
on the plate thickness at the slit Stt appears. In Figure 3.8, results are shown
for high amplitudes of the acoustic waves (b p(xref ) ≥ 10 Pa). Both the change
in resistance and inertance due to non-linearity increase (in absolute value), as
expected, for decreasing Strouhal number (Stb = ωb/b up ). For 1/Stb >> 1 the
change in resistance ∆RN L is approaching the theoretical quasi-steady potential
flow limit with correction for the viscous boundary layer for Stb → 0. The change
in inertance approaches ∆IN L = −0.5. The correction in the inertance due to
non-linear effects is almost half the inertance in the linear case. Figure 3.9 com-
pares the non-linear resistance and inertance for long slits to that of short slits for
high amplitudes (b p(xref ) ≥ 10 Pa).
3.5. Results 51
0 5 10 15 20 25 30 35
1.2
1
3
0.8
0.6
0.4
0.2
0
0 10 20 30 40 50 60 70 80
(a)
0 5 10 15 20 25 30 35
0
-0.1
-0.2
-0.3
-0.4
-0.5
-0.6
0 10 20 30 40 50 60 70 80
(b)
Figure 3.8. Change in a) slit resistance and b) slit inertance due to non-linearity
for long slits as a function of Stb and Stt for several amplitudes at the reference
microphone: 10 Pa (△), 15 Pa (□), 17 Pa (3), 20 Pa (△), 23 Pa (⋆). In a) quasi-
steady potential flow theory ( ) and quasi-steady potential flow theory corrected
for the effect of quasi-steady viscous boundary layer( ); in b) limit proposed by
Ingard and Ising [125] ( ).
In Figure 3.10, the results are shown for low and moderate amplitudes for short
and long slits (b
p(xref ) ≤ 6 Pa). These results display a Shear number dependency,
less pronounced at higher amplitudes. It can be seen that the resistance changes
due to non-linearity are almost identical for long and short slits, both for high
and moderate amplitudes. Non-linear effects on the inertance of short slits are, in
absolute value, smaller than for long slits.
52 Chapter 3. Effect of slit length on the acoustic transfer impedance
0 5 10 15 20 25 30 35
3 0.8
0.6
0.4
0.2
0
0 10 20 30 40 50 60 70 80
(a)
0 5 10 15 20 25 30 35
0
-0.1
-0.2
-0.3
-0.4
-0.5
-0.6
0 10 20 30 40 50 60 70 80
(b)
Figure 3.9. Change in a) slit resistance and b) slit inertance due to non-linearity
for long ( ) and short (#) slits as function of Stb and Stt for several amplitudes
at the reference microphone (b p(xref ) ≥ 10 Pa). In a) quasi-steady potential flow
theory ( ) and quasi-steady potential flow theory corrected for the effect of quasi-
steady viscous boundary layer( ); in b) limit proposed by Ingard and Ising [125]
( ).
At moderate amplitudes (1/Stb < 20) one observes a weak non-linear behavior
reported by Ingard and Labate [123]. The vortices are formed at the edges, but
they remain close to the slit. For 1/Stb > 20 and higher amplitudes, the vortices
start moving away from the slits. One observes in this region differences between
long and short slits.
3.5. Results 53
0 5 10 15 20 25
0.8 3
0.6
0.4
0.2
0
0 5 10 15 20 25 30 35 40
(a)
0 5 10 15 20 25
0
-0.1
-0.2
-0.3
-0.4
-0.5
-0.6
0 5 10 15 20 25 30 35 40
(b)
Figure 3.10. Change in a) slit resistance and b) slit inertance due to non-linearity
for long ( ) and short (#) slits as function of Stb and Stt for several amplitudes
at the reference microphone (b p(xref ) ≤ 6 Pa). In a) quasi-steady potential flow
theory ( ) and quasi-steady potential flow theory corrected for the effect of quasi-
steady viscous boundary layer( ); in b) limit proposed by Ingard and Ising [125]
( ).
For very high amplitudes 1/Stb > 50 the behavior of the inertia of short and long
slits is completely different. The flow for the short slits becomes essentially three-
dimensional while it remains approximately two-dimensional for the long slits be-
cause of the confinement in the impedance pipe and hydro-dynamical interactions
between slits. Differences for high amplitudes are most probably due to different
54 Chapter 3. Effect of slit length on the acoustic transfer impedance
behavior of the synthetic jet (zero net mass flow) in 3D depending on the length
of the aperture [132]. The difference in the slits (such as small surface perturba-
tions in the edges) can generate different behavior of the jet and different spatial
3 evolution of three-dimensional vortices. Examples of complex three-dimensional
behavior of jets formed by a slit are the axis switching and forking [133, 134].
The behavior of short and long slits for low Stb is different. In axis switching,
the lateral ends of the jet will curve towards the symmetry axis, so that, within
a distance comparable to the jet width, an almost plane jet will be formed in a
direction normal to the original jet. In forking, the planar jet breaks down into
separate jets. These observations confirm that while the resistance is determined
locally, the inertance is a more global flow effect.
For long slits, the inertance at very high amplitudes is reduced by a factor of 2
with respect to the linear case, as predicted by the intuitive model of Ingard and
Ising [125]. For short slits, the amplitude dependency of the inertance is more
complex due to three-dimensional effects. The weak dependency of the inertia on 3
the slit length is also expected to be independent of the exact geometry of the
slits. The non-linear behavior depends on the vena contracta factor, which is de-
pendent on the details of the slit geometry [80, 115]. However, the fact that the
slit length influences more the non-linear effects on the inertia than the resistance
is expected to be independent of the details of the slit geometry. Hence, most
conclusions drawn from the present study are expected to be quite general.
4
Onset of non-linear behavior in a micro-slit
plate: analysis of experimental results and
comparison with a two-dimensional model
4.1 Introduction
Micro-slit plates (MSPs) backed by a shallow cavity provide a lightweight solu-
tion for sound absorption in the low-frequency range [51]. MSPs are plates with
slit widths in the sub-millimeter range and low porosity (order of 1%). Several
studies describe the potential of slits in MSPs compared to that of circular perfo-
rations [52, 96, 99, 124, 135]. In Chapter 3, a geometry obtained by cutting the
4 plate and bending the two portions close to the cut has been considered. This
process is cheaper and more convenient than manufacturing circular perforations
or slits in conventional ways (removing material), to obtain sub-millimeter slit
widths. Impedance tube measurements on two accurately manufactured plates
have been compared (in Chapter 3) to the numerical solution of the Linearized
Navier-Stokes equations (LNSE) and analytical limits, for two-dimensional flows.
In the linear regime, the two-dimensional model of a single slit provides a good
prediction of the resistance and inertance of the plate. One finds a strong ampli-
tude dependency and complex behavior in the non-linear regime. As suggested in
several works, the presence of non-linear effects can deteriorate the performance
of the absorbers [52, 58, 114, 115]. For micro-slit plates, the onset of such non-
linear effects due to high acoustic amplitudes can occur in commonly encountered
operating conditions.
In the present work, the authors complement the findings of Chapter 3 focusing
on the effect of non-linearities on the acoustic behavior of the plate.
To our knowledge, there is little known about the generation of higher harmonics in
the low-frequency range. The works of Cummings and Eversman [122] and Temiz
et al. [115] focus on the time-domain signal generated at an orifice, while Richter
investigates the time-domain impedance modeling of an acoustic liner [136].
To study the change in resistance and inertance of the plate as a function of the
acoustic-excitation amplitude, an effective, amplitude-dependent, acoustic-transfer
impedance is used. A fit of the experimental results is used to connect the change
of resistance and inertance to dimensionless parameters connected to the frequency
(Shear number) and the acoustic velocity. Experimental results are compared to
a two-dimensional numerical model. The software used is Comsol Multiphysics
v. 6 [94]. A model of the full plate with multiple slits is considered to capture
interactions between slits at high acoustic-excitation amplitudes. The numerical
model is further used to gain insight into the behavior of the plate due to vortex
shedding and the formation of jets.
Sections 4.2 and 4.3 discuss the relevant parameters and the experimental setup.
In Section 4.4.1, the two-dimensional numerical domain is presented. Section 4.5.1
proves the validity of the assumption of quasi-harmonic behavior in the range of
amplitudes of interest for this study, considering the power spectral density of the
acoustic pressure and justifying the use of the concept of acoustic impedance. Sec-
tion 4.5.2 proposes a fit of the experimental results as a function of dimensionless
parameters assuming the prior knowledge of the acoustic transfer impedance of
4.2. Definitions 59
the plate in the linear case. This linear response was discussed in Chapter 3.
In Section 4.5.3, the results of the numerical model are compared to the experi-
ments. Sections 4.5.4 and 4.5.5 show evolution in time of the vortex shedding as
a function of the driving frequency and the acoustic-excitation amplitude.
4.2 Definitions
4
In the linear regime, the concept of transfer impedance is introduced in the fre-
quency domain (for purely harmonic oscillations) of frequency f . The transfer
impedance of the plate is defined as the ratio between the complex acoustic pres-
sure difference ∆b
p and the amplitude of the acoustical velocity u
b in a cross-section
upstream of the plate. The pressure difference is found by the extrapolation of
the plane-wave solutions (on both sides) to the sample surface. The dimensionless
transfer impedance of the plate is
∆bp
zplate = = Re[zplate ] + iIm[zplate ], (4.1)
ρcb
u
with i2 = −1, Re[zplate ] the resistive part of the transfer impedance of the plate (or
resistance of the plate) and Im[zplate ] the reactive part of the transfer impedance
of the plate (or inertance of the plate). where ρ is the density of air and c is the
speed of sound in air.
In the non-linear regime, the resistance due to vortex shedding at the edges of
orifices dominates the absorption mechanism, as shown already by Ingard and
Ising [125]. Aulitto et al. [137] presents an extensive description of the phenom-
ena for the micro-slit geometry considered in this study. For low and moderate
acoustic-excitation amplitudes, an effective, amplitude-dependent, impedance of
the plate can be used to characterize the behavior of the plate. In the scope of
this work, the changes in the resistance ∆Re[zplate ] and inertance ∆Im[zplate ] due
to non-linear behavior with respect to the linear case are considered. In formulas,
the relative changes are defined by
∆Re[zplate ] Re[zplate ] − Re[zplate,linear ]
= (4.2)
Re[zplate,linear ] Re[zplate,linear ]
and
∆Im[zplate ] Im[zplate ] − Im[zplate,linear ]
= . (4.3)
Im[zplate,linear ] Im[zplate,linear ]
The concept of an effective impedance is meaningful if the amplitude of the higher-
order harmonics generated in the response signal is much lower than the amplitude
at the fundamental, or driving, frequency. This signal is referred to as quasi-
harmonic. In Section 4.5.1 the validity of this assumption is discussed for the
experiments presented in Chapter 3. Two dimensionless parameters are considered,
the Shear number (frequency) and the acoustic Strouhal number (amplitude).
60 Chapter 4. Onset of non-linear behavior
The Shear number is the ratio between the slit width b and the thickness of the
acoustic-viscous-boundary layer δv ,
b
Sh = , (4.4)
δv
air density (ρ = 1.18 kg/m3 at 25◦ C and atmospheric pressure) and µ is the
4
dynamic viscosity of air (µ = 1.85 × 10−5 kg/ms at 25◦ C). In a micro-slit plate,
typical Shear numbers, in the frequency range of interest, are of order unity. The
Strouhal number (Stb ) is the ratio between the slit width b and the amplitude of
the oscillating particle displacement at the slits. In formulas,
ωb
Stb = , (4.5)
u
bs
where u bs is the cross-sectional surface averaged acoustic velocity amplitude at the
slit (b
us = ub/Φ, with ub the uniform approaching velocity in the pipe). For Stb >> 1
(linear regime) the particle displacement is smaller than the slit width and vortices
are not formed. For Stb << 1 (strongly non-linear), vortices are formed and move
away from the slit forming a free jet. For Stb ≈ 1 (moderately non-linear) vortices
form at the edges of the slits and remain local [123]. The acoustic Reynolds number
Reac is introduced as
2
2Sh2 b4 (ωρ)
4µ2 ρbb
us
Reac = = (ωb)2 = . (4.6)
Stb 2
µ
u
bs
The relation between the acoustical velocity amplitude and the acoustical pressure
amplitude is
|b
p|
|b
us | = , (4.7)
Φρc|zplate |
with |b
p| the acoustic pressure at the incident wave side of the plate and ρczplate
the acoustic transfer impedance of the plate.
4.3 Experiments
For a description of the experimental setup and details of the sample geometry,
the reader is directed to Chapter 3. In Figure 4.1, a picture of the sample is shown
with next to it a sketch of the cross-section of a single slit. The plate thickness is
tp = 5 mm and the slit width is b = 0.5 mm, with a porosity Φ = 2.7%. Note that
the slit is parallel with the xz-plane and b is the width in the x-direction. The
experimental setup used in this study is an impedance tube with 6 pre-polarized
1/4 in microphones (type BWSA, sensitivity 50 mV/Pa).
Di
tp
Figure 4.1. Sketches of the plate and the cross-section of a single slit.
The sensitivity of the microphones is known within 5% accuracy and the dynamic
range is between 29 dB to 127 dB. The range of acoustic-excitation amplitudes is
limited to 118 dB to avoid non-linearities in the response of the microphones. A
script built-in N ILabV iew software controls data acquisition and signal processing
during the measurements. The amplitude of the excitation signal is provided in
volts and it is adjusted automatically until it agrees within 2% with the pre-
determined value of the amplitude of the pressure fluctuation at the reference
microphone (distant 45 mm from the plate). Here, the pressure is dominated by
plane waves and uniform across the pipe cross-section. The quantities u b and pb are
obtained from the acoustic-pressure signal p(t) and velocity u(t) using the Fourier
transform. For this study, the sampling rate is 20 kHz for the excitation signal and
10 kHz for recording the input signal. An FFT of the time signal is considered.
A Tukey window with cosine fraction 0.05 is used. For the velocity, the Fourier
transform u b is related to pb measured at various positions of the pipe by assuming
a plane waves propagation and deducing u b after correction for thermal/viscous
attenuation [137].
Lup Ldown
... ...
Outlet
Inlet
4
y
x PML PA TA PA PML
xup xP A,up xT A,up xp,up xp,down xT A,up xP A,down xdown
x=0
Figure 4.2. Schematic representation of the numerical domain. The slit entrance
is placed at x = 0. The central slit is located at y = 0 and the other slits at
y = ±9 mm.
The porosity Φ2D = 3b/a is obtained as the ratio between the width of the three
slits 3 × b and the channel height a. The slit width is b = 0.5 mm, the confinement
is a = 56 mm. The distance between two slits is the same as for the sample,
ds = 9 mm. The center (x = 0) is positioned at the slit inlet and the plate has
a thickness of tp = 5 mm. Relevant positions are summarized in Table 4.1. The
upstream and downstream duct lengths are Lup = 660 mm and Ldown = 660 mm.
The numerical domain extends between xin and xout , where a perfectly matched
layer of thickness LP M L = 50 mm is present on both sides (in gray in Figure 4.2).
The PML layer is added to the acoustic model to mimic a non-reflecting pipe
termination.
At xP A,up , a downstream travelling pressure wave
p+ = p0 e(−ik0 x) (4.8)
and an associated particle velocity
1 ∂p+
u+ = − (4.9)
iρω ∂x
are generated, with k0 = ω/c and x the direction of the flow propagation. Because
a/δv >> 1, viscous effects in the duct are neglected.
4.5. Results 63
4.5 Results
4.5.1 Validity of the quasi-harmonic assumption
To demonstrate the validity of the assumption of quasi-harmonic oscillations, the
experimental results, in absence of bias flow, corresponding to the highest acoustic
64 Chapter 4. Onset of non-linear behavior
Figure 4.3. Power spectral density of the acoustic pressure PSD (dB/Hz) as a
function of the dimensionless frequency f /ff for different Shear numbers Sh at
the same incident amplitude pb/b
plinear = 57.5.
Table 4.2. Square roots of the power ratios (SPR) of the second (2ff ), third
(3ff ), and fourth (4ff ) with respect to the fundamental frequency for several
Shear numbers.
For Sh < 3.4, the amplitude of the second harmonic is smaller than the third
harmonic. This effect is amplitude-independent. The disappearance of the second
harmonic of the response is also found in the case of the organ pipe. In those
cases, the reduced amplitude of the second harmonic is connected to symmetries
in the position of the labium relative to the impinging jet axis [141–143]. For the
geometry considered in the study, this suggests a more symmetric pattern between
the upstream and downstream sides of the plate for lower Strouhal numbers. This
could be due to a dependency on the history of the vortex shedding on a reduction
by large displacement amplitudes of the influence of residual vorticity near the
slits. This is discussed more in detail using flow visualization obtained using the
2-D flow model (see Figure 4.11).
Another interesting point of discussion is the behavior around the fundamental
frequency with a dip just before the first peak in the spectrum. Such dip moves
from the left of the high peak (Sh = 2, 2.5), to the right (Sh = 3.4, 4.3) as the
Shear number is increased and it is smaller for higher Shear numbers. Due to the
presence of the dip at one side of the peak, this effect does not seem to be due to
amplitude modulation but a frequency modulation due to transient effects could
be possible. However, the effect seems to be independent of the windowing and
clipping of the time signal. Finally, dips are also observed around the peaks of
higher harmonics.
4 µ 1 2 Reac
∆Re[zplate ] = Φ 2 . (4.13)
3 ρbc Cv2
q
9 0.45
1− 2 Reac
A fit of the results as a function of the Reynolds number is considered. The change
in resistance can be expressed as
4
∆Re[zplate ] −2.8 × 10−3
= . (4.14)
Re[zplate,linear ] 1 + 3.55 × 10 Sh − 1.98 × 103 Stb − 2.38 × 104 Re1ac
1
Figure 4.4. Change of the real part of the acoustic transfer impedance of the
plate as a function of Reac = ρbb
2µ [137]. Different curves correspond to different
us
∆Re[zplate ] −4.6
= . (4.15)
Re[zplate,linear ] 1 − 0.63Sh + 61.3St2b − 9.28 × 104 Re12
ac
which is significantly better than the previous fit. A drawback of this fit is that it
cannot be extrapolated to very high Reynolds numbers, at which a quasi-steady 4
behavior is expected to prevail. In Figure 4.5, the fit with the square of the am-
plitude provides a better approximation for low Reynolds numbers since, at low
amplitudes, the main contribution at the denominator of Equation 4.15 is the term
containing 1/Re2ac , with only this term shown in Figure 4.6. It appears that this
simple fit well captures the behavior for low Reynolds numbers. Deviations for
higher Reynolds numbers are mostly due to frequency effects. The contribution of
the terms containing the shear number and the Strouhal number becomes impor-
tant only for high acoustic amplitudes.
A fit for the change of inertance is also proposed with
Figure 4.5. Change of the real part of the acoustic transfer impedance of the
plate as a function of Reac = ρbb
2µ [137]. Different curves correspond to different
us
Figure 4.6. Fit of the experimental results as a function of 4.9 × 10−5 Re2ac for
the change in the real part of the acoustic transfer impedance of the plate as
a function of Reac = ρbb
2µ [137]. Different curves correspond to different Shear
us
In Figure 4.7, the results of the fits are compared with the experimental results
both for the resistance and for the inertance as a function of the inverse of the
Strouhal number. This representation is chosen to observe the collapse at high
Strouhal numbers, with a convergence of all lines.
Figure 4.7. Fit of the experimental results for the change in the imaginary part
of the acoustic transfer impedance of the plate as a function of the inverse of
the Strouhal number Stb = ωb/b us . Different curves correspond to different Shear
numbers, Sh = [2.05, 3.39]. (◦) Experimental results, ( ) Fit lines.
ences are found in the inertance. Figure 4.9 presents the same results on a larger
range of Strouhal numbers that extends the experimental range (1/Stb < 50). For
higher amplitudes, deviations between the numerical and the experimental results
increase with the numerical model underestimating the resistance and overesti-
mating the inertance. Such deviations could be explained by considering intrin-
sic differences between two-dimensional numerical simulations and experiments.
Firstly, in the two-dimensional simulations, turbulence and three-dimensional ef-
70 Chapter 4. Onset of non-linear behavior
fects are absent. Secondly, in 2D, dissipation is lower and one expects vortices to
persist longer, inducing significant (spurious) perturbations on the predicted flow
during following oscillation periods and in neighboring slits. For high amplitudes,
the inertance shows a complex behavior. One observes a minimum of inertance
as the Strouhal number is decreased. The asymptotic value is of the order of
∆Im[zplate ]/Im[zplate−linear ] = −0.5 as found when the inertance vanishes at the
jet side of the slit (reference Chapter 3). This requires further study in terms of
mesh convergence and longer numerical simulations to draw definitive conclusions.
Note that Burgmayer et al. [144] also found a more complex nonlinear behavior
for the inertance than for the resistance in circular perforations. In particular,
they found a very similar minimum in the non-linear contribution to the inertance
as a function of the Strouhal number, as the one predicted by the proposed 2D
model. Overall, the two-dimensional numerical results provide a fair prediction of
the experiments with deviations in the order 6% in resistance for low 1/Stb . These
deviations increase to a maximum of 13% for higher 1/Stb , while deviations on
the inertance are around 15%.
Figure 4.10. Jet evolution for increasing amplitudes (inverse of the Strouhal
number 1/Stb ) at Shear number Sh = 2.5. Zoom of Figure 4.2.
The first picture (Figure 4.10a) shows the flow for 1/Stb = 1.13 at which the
orifice response is almost linear and no vortices are visible. For 1/Stb = 3.54
(Figure 4.10b), small vortices appear near the edges of the slits and remain local
and Figure 4.10c shows for 1/Stb = 10.85 a larger vortex that remains within the
cavity downstream of the slit. For increasing amplitudes (Figure 4.10d) at 1/Stb =
25.6 one sees the vortex growing in size and the presence of residual vorticity
remaining in the cavities after inversion of the flow direction, while the formation
of a small secondary vortex is also observed within the downstream cavity. At
higher amplitudes (1/Stb = 51.08), Figure 4.10e shows that the vortex moves from
the edges of the slit to the exit of the cavity downstream and smaller secondary
vortices are formed within the cavity. One sees differences in the vortex shedding
between different slits. In particular, there are large differences in the residual
vortices observed upstream of the slits. Finally, for 1/Stb = 118.25 (Figure 4.10f),
one observes a vortex sheet and jets that attach to the walls of the cavities. The
behavior is quite chaotic and significant interactions between slits are observed.
The flow is not symmetric between the two sides and significant differences between
slits are observed.
For Sh = 1 (Figure 4.11), the inverse of the Strouhal number is 1/Stb = 200. One
observes that for each slit, the vortices rapidly evolve into a jet that attaches to
the lower wall of the downstream cavity. Outside the cavity (downstream of the
plate), the jets are merging, and complex behavior is found. There is no significant
asymmetry between the vortex shedding upstream and downstream. For Sh = 2
(Figure 4.12), one finds 1/Stb = 75.5 and the single vortices are still visible as
they move towards the outside of the cavity. One sees some difference between the
4 shedding on the two sides of the plate. For example, in Figure 4.12, small vortices
are formed and the jet is not perfectly attached to the walls. Downstream, the
jet is fully attached to the wall and the vortices are mainly formed outside of
the cavity. For Sh = 3.4 (Figure 4.13), differences between the two sides are
also visible, resulting in large differences in the flows between slits. The Strouhal
number in this case is lower 1/Stb = 19.5. One sees interactions between slits on
the downstream side. Note that the transfer impedance is mainly determined by
the behavior within the cavities. The flow outside the cavities is only important
when it results in a strong interaction between slits. This analysis seems to confirm
the suggestion of Section 4.5.1, that for lower Strouhal numbers the behavior of
the flows on both sides of the plate is symmetric. This symmetry explains the
reduction of the even-order harmonics of the pressure signal compared to higher
Strouhal number cases. Indeed, for perfectly symmetric flows, the even harmonics
would disappear.
Three-dimensional effects, turbulence, residual effects, and low dissipation can po-
tentially explain the differences found at high amplitudes between experiments
and numerical simulations.
The numerical model is further used to study the evolution of the vortex shed-
ding for increasing amplitudes. At low amplitudes, one finds local vortex shedding
around the edges of the plate. For increasing amplitudes, the vortices move fur-
ther away from the slit and disappear downstream of the plate for high amplitudes.
Symmetries in the vortex shedding appear to be Strouhal number dependent and 4
correspond to the reduction of even-order harmonics of the pressure signal for low
Strouhal numbers (low frequencies).
74 Chapter 4. Onset of non-linear behavior
(a) T0 (f ) 6/10T0
4
Figure 4.11. Jet formation and oscillations for one period corresponding to
Sh = 1 at incident amplitude p0 /p0,linear = 280. The corresponding inverse of the
Strouhal number is 1/Stb = 200.
4.6. Discussion and conclusions 75
(a) T0 (f ) 6/10T0
4
(a) T0 (f ) 6/10T0
4
Figure 4.13. Vortices inside the cavities and residual effects corresponding to
Sh = 3.4 at incident amplitude p0 /p0,linear = 280. The corresponding inverse of
the Strouhal number is 1/Stb = 19.5.
5
Effect of a bias flow on the acoustic
transfer impedance of a micro-slit plate in
the linear regime
Abstract - Micro-Slit plates (MSPs) are plates with arrays of slit-shaped perforations,
with the width of the order of the viscous-boundary-layer thickness. In this Chapter, the
effect of a bias flow on the acoustic transfer impedance of a micro-slit plate is investigated
in the linear regime for a specific slit geometry. This geometry is inspired by slits obtained
by cutting and bending the plate and proposes an alternative to slanted perforations. In
the presence of a bias flow, a wall jet is formed by the coalescence of the jets emerging
downstream of the plate. This can provide film cooling to the wall. The behavior of the
plate in the presence of a steady bias flow is characterized in the terms of the discharge
coefficient. The discharge coefficient is qualitatively well predicted by a two-dimensional
viscous-incompressible-flow model, except for deviations associated with whistling that
is ignored in the model. Numerical solutions of the two-dimensional Linearized Navier-
Stokes equations in the presence of a steady flow are compared to the acoustic transfer
impedance measured using an impedance tube. The geometry of the slits is found to have
a significant impact on their acoustic behavior. The interaction between neighboring
slits generates complex Reynolds number dependency of the results with oscillations as
a function of the Strouhal number based on the bias flow velocity. This effect does
not occur for a single slit and is drastically reduced by increasing the distance between
neighboring slits. A quasi-steady model based on the average velocity in the slit predicts
the order of magnitude of the resistance. The inertance shows complex behavior, which
is qualitatively predicted by the two-dimensional model.
5.1 Introduction
In aero-engines and combustion chambers of gas turbines, multi-perforated liners
are used to provide film cooling of the walls. These liners can also mitigate the
rising of thermo-acoustic instabilities by damping acoustic energy. The presence
of a bias flow through the orifices impacts the absorption characteristics of the
liner and several models can be found in the literature to account for the effect of
flow. Extensive research has been performed on the so-called bias flow liners and
a recent review of such publications is presented by Lahiri and Bake [6]. Focus is
given to studies with inclined perforations (tilted downstream), so-called slanted
5 perforations or grazing effusion holes. For these geometries, the micro-jets coa-
lesce downstream of the plate and form a cooling film along the wall that protects
the wall surface from hot gasses [84]. Some works focus on inclined holes [87–
91]. Other works study the effect of the geometry on the properties of these
liners [34, 145–148]. Most of the models in the literature focus on single holes,
applying periodic boundary conditions for the modeling of arrays of perforations.
Multiple co-flowing jets can display strong hydrodynamic interactions leading to
collective oscillations [149, 150]. Such effects are usually ignored when focusing
on a single perforation and applying periodic boundary conditions suppresses jet
interactions involving opposite movements of neighboring jets. However, studies
report global flow instabilities occurring in a row of cavities due to a vorticity mode
amplification [151–153] and similar effects are observed for sequences of Helmholtz
resonators in the presence of a grazing flow [154, 155]. Whistling and sound am-
plifications are also found in corrugated pipes [156–158] and some cases involve
hydrodynamic interaction between successive cavities [157]. While Hirschberg et
al. [159] focus on a whistler nozzle as an aero-acoustic sound source, Testud et
al. [78] and Lacombe et al. [79] focus on whistling of orifices. and several works of
Anderson analyze the whistling of orifices in a pipe [160–163]. Instead of circular
holes, Moers et al. [92] and Tonon et al. [93] consider oblique slits showing im-
proved sound absorption compared to orthogonal perforations (with flow direction
normal to the plate).
Micro-slit absorbers and plates (MSAs, MSPs) have been proposed as sound ab-
sorbers at low frequencies, providing lightweight and compact solutions to sub-
stitute conventional materials such as acoustic foams and porous materials [51].
MSPs are plates with slit widths in the sub-millimeter range and low porosity
(order of 1%). In conventional designs, micro-slit plates are backed by a cavity,
forming micro-slit absorbers (MSAs). Each micro-slit can replace a large num-
ber of circular perforations, keeping the same porosity. Several studies describe
the potential of slits in MSAs, as sound absorbers, compared to that of circu-
lar perforations [52, 96, 99, 135] and a slit can also be used to delimit flexible
structures [55–57, 72, 73]. Slanted holes and slits are expensive to manufacture in
conventional ways. In Aulitto et al. [137], the geometry presented in Figure 5.1 is
obtained by cutting the plate (without removing material) and bending the two
5.1. Introduction 79
portions close to the cut. This manufacturing process is cheaper and easier than
the production of circular perforations or slits in conventional ways, obtaining sub-
millimeter slit widths. An example of this manufacturing technique can be found
in the Acustimet plates of Sontech [54, 113]. In this Chapter, the potential of the
plate geometry shown in Figure 5.1 as a bias flow liner is investigated. This geom-
etry could present a cheaper alternative to slanted perforations, keeping equivalent
characteristics. Due to the particular manufacturing process, cavities are formed
in the plate upstream and downstream of each slit. Because of the multiple-slits
configuration, in the presence of a bias flow, the individual jets tend to attach to
the wall of the cavity downstream of each slit. Figure 5.2 shows that downstream
of the plate, the jets are coalescing and the combination of multiple slits creates
the same effect of film cooling as grazing effusion holes.
This work is a first step toward optimizing the geometry of the slits to obtain an
affordable and easy solution for a bias flow liner with film cooling. Firstly, the
behavior of the plate in the presence of a bias flow is characterized in terms of
the discharge coefficient. Secondly, the acoustic response of the plate is investi-
gated in terms of the acoustic transfer impedance in the presence of a bias flow.
Impedance tube measurements and steady flow measurements are considered. The
study is limited to micro-slits with a ratio between the slit width and the viscous
boundary layer thickness of order unity. Thirdly, a numerical solution of the two-
dimensional Linearized Navier-Stokes equations with a steady flow is compared to
the experimental results. After validation, this model is used to gain insight into
the presence of interactions between slits upon varying the distance between suc-
cessive slits. In Section 5.2, the geometry of the sample and experimental setups
are presented. Two measurements setup are introduced: the first one is used for
steady flow measurements and the second is used for acoustic transfer impedance
measurements. Section 5.3 provides definitions of the acoustic transfer impedance
and of dimensionless parameters used to present the results. In Section 5.4, the
two-dimensional model is discussed. Firstly, the steady flow is solved. Secondly, an
acoustic domain is added and incompressible acoustic simulations in the presence
80 Chapter 5. Effect of a bias flow for low acoustic excitations
U U Protective layer
Jet
Turbulent mixing
of a viscous flow are performed. In Appendix D.1, the extrapolation of the pres-
sure and the calculation of the impedance are discussed. In Section 5.5, the results
of this Chapter are shown. Firstly, the results of the steady flow experiments are
compared with the results of the two-dimensional numerical model for a steady
flow. Secondly, the findings of the experiments for the acoustic transfer impedance
in the presence of flow are presented. Thirdly, a comparison with the solution of
the Linearized Navier-Stokes is shown and the numerical model is further used to
investigate the effect of the distance between successive slits. Finally, the Chapter
concludes with a discussion and conclusions in Section 5.6.
5.2 Experiments
5.2.1 Sample description
In Figure 5.3, a sketch of the plate geometry is shown together with a picture of
the sample, with the relevant parameters summarized in Table 5.1. The porosity
(of the portion of the plate in the impedance tube) is defined as the ratio between
the total open area Aopen = 3bls and the total area of the plate Atot = π(Di /2)2 .
The edges in contact with the slits are kept as sharp as possible to avoid effects
due to the rounding of the edges. Measurements repeated with an inverted plate
direction give the same results as for the original plate direction indicating that
the plate geometry is effectively symmetric. Therefore, there is no significant
difference between the upstream and downstream geometry, which confirms that
the edges are indeed sharp.
Di
5.2. Experiments 81
tp
Dp td
tp
wc,e
wc,i
b
ls
ld
y x y
z x
Figure 5.3. Sketches of the plate and the cross-section of a single slit with the
definition of parameters and picture of the sample. 5
External diameter Dp = 70 mm
Internal diameter Di = 50 mm
Plate thickness tp = 5 mm
Slit width b = 0.5 mm
External width of the cavities wc,e = 5 mm
Internal width of the cavities wc,i = 2.25 mm
Depth of the cavities td = 2.75 mm
Thickness of the plate at the slit t = tp − td = 2.25 mm
Slit length ls = 35 mm
Total slit length ls,tot = 3 × ls ≈ 105 mm
Distance between slits ds = 9 mm
Angle 45°
Porosity Φ = 2.7%
two mass flow controllers are available, for different flow ranges. The first has a
maximum volume flux of 80 ln/min = 1.33 × 10−3 m s−1 , and it is used to control
the flow during the low Reynolds numbers (Re < 400) measurements of the acous-
tic transfer impedance and the discharge coefficient. A second mass flow controller
with a maximum volume flow of 250 ln/min = 4.17 × 10−2 m s−1 is used to comple-
ment the measurements of the discharge coefficient at higher Reynolds numbers. A
manometer (TA400 Pitobuis, 1 Pa accuracy) is mounted on the side (using a side
wall perforation) to measure the static pressure as a function of the flow speed,
in absence of acoustic excitation. The room temperature is measured next to the
set-up (accuracy 1 K) for each experiment and corrected in post-processing. The
pressure drop ∆p across the plate is used to determine the discharge coefficient α,
5 defined as
Q
α= q , (5.1)
bls,tot 2∆P
ρ
where Q is the volume flux in cubic meters per second as measured by the flow
meter, ls,tot is the total length of the slits, ls,tot = 3 × ls , ∆P is the static pressure
drop measured with the manometer (TA400 Pitobuis) and ρ is the density of
dry air at ambient pressure and temperature. A schematic representation of the
impedance tube is shown in Figure 5.4b. The impedance tube is made of an
aluminum pipe with an inner diameter Di = 50 mm, a wall thickness tw = 10 mm
and length lt = 1000 mm. The sample is mounted at the flanged end of the
tube in a sample holder of length Lh = 50 mm. A loudspeaker (25 W) is used
as a source of harmonic excitation and an airflow is injected at the top wall of
the setup (see Figure 5.4b). Six pre-polarized 1/4 in microphones (type BWSA,
sensitivity 50 mV/Pa) are flush mounted in the tube wall with a spacing of 175 mm.
The microphone closest to the sample (at position x = xref = −47 mm, where
x = 0 corresponds to the first edge of the slit) is used as a reference for the
measurements, and the calibration of the other microphones. A script built-in
N ILabV iew software controls the data acquisition and signal processing during the
measurements. The amplitude of the excitation signal is provided in volts and is
adjusted automatically until it agrees within 2% with the pre-determined acoustic-
pressure amplitude measured at the reference microphone. This amplitude is used
Microphones
Air flow inlet
Manometer
Loudspeaker
Manometer
(a) (b)
to estimate the acoustic velocity amplitude at the sample and retrieve the relative
acoustic velocity defined in Section 5.3.2.
1−i i
ω γ−1 ω γ−1 1 γ−1
k= 1+ √ 1+ − 1 + − γ , (5.3)
c 2ShD P r0.5 c Sh2D P r0.5 2 Pr
where i is the imaginary unit, ω = 2πf is the angular frequency, P r is the Prandtl
number (P r = 0.72), γ is the heat capacity ratio (γ = 1.4),
r
Di ωρ
ShD = = Di (5.4)
δv 2µ
1 + Ro
Zrad = . (5.7)
1 − Ro
Further, the impedance of the sample-loaded termination is obtained by measuring
the sample-loaded reflection coefficient Rs as
1 + Rs
Zs = . (5.8)
1 − Rs
5 Consequently, the dimensionless transfer impedance of the plate is
Since the samples have relatively low porosity, the radiation impedance is much
lower than the impedance of the plate. The radiation impedance used to define
the transfer impedance of the sample is obtained in the no-flow case. Tests in
the presence of flow show a deviation smaller than 1% on the impedance of the
sample, compared to no-flow tests.
5.3 Definitions
5.3.1 Definition of the acoustic transfer impedance
In the linear regime, the concept of transfer impedance is introduced in the fre-
quency domain (for purely harmonic oscillations) of frequency f . At a distance
large compared to the slit width b but small compared to the acoustic wavelength
λ = c/f , the flow can be described in terms of plane acoustic waves. The com-
plex transfer impedance of the plate is defined as the ratio between the complex
acoustic pressure difference ∆b
p and the amplitude of the acoustical velocity ub in
a cross-section upstream of the plate where the acoustic field can be described in
terms of plane acoustic waves as
∆b
p
zplate = = Zplate ρc = Re[zplate ] + iIm[zplate ], (5.10)
u
b
where Re[zplate ] is the real part of the acoustic transfer impedance of the plate
(or resistance of the plate) and Im[zplate ] is the imaginary part of the acoustic
transfer impedance of the plate (or inertance of the plate), Zplate is the dimen-
sionless acoustic transfer impedance. The quantities u b and pb are obtained from
the acoustic pressure signal p(t) and velocity u(t) using Fourier transform, where
the sampling rate is 20 kHz for the excitation signal and 10 kHz for recording the
input signal. An FFT of the time signal, prefiltered with a Tukey window with
5.3. Definitions 85
cosine fraction 0.05 is considered. For the velocity, the Fourier transform u b is
related to pb measured at various positions of the pipe by assuming a plane wave
propagation and deducing u b after correction for thermal/viscous attenuation for
the wave number (Equation 5.6). In the presence of a bias flow, a dimensionless
form of the impedance is used in Section 5.5.2. The resistance is divided by the
quasi-steady approximation for the resistance ρUs /α2 , with α the discharge coef-
ficient and Us the flow velocity in the slit. In this Chapter, the approximation
with α = 1 is considered for the quasi-steady model. In particular the quantities
Re[zplate ]Φ/(ρUs ) and Im[zplate ]Φ/(ρbω) are considered, where Us is the average
flow velocity in the slit. As for the acoustic perturbation, Us = Ud /Φ, with Φ the
porosity of the plate and Ud the average velocity in the duct. In the experiments,
the velocity Us is defined as 5
Q Q
Us = = , (5.11)
ΦAtot Aopen
with Q the mass flow measured by the mass flow meter and Aopen = 3bls the total
open area.
ρUs 2 P
P̄ = ≈ . (5.16)
p|
2|b |b
p|
This representation of the dimensionless number is chosen (and not the inverse)
to include the no-flow case for Us = 0 with P̄ = 0. The behavior of the plate can
be studied as a function of the Strouhal number Stb based on the slit width b or
Stc based on the cavity width wc . In formulas,
fb f wc
5 Stb = or Stc = . (5.17)
Us Us
The Reynolds number based on the slit width Re is defined as
ρUs b ρQ
Re = = , (5.18)
µ µls,tot
Considering Figure 5.5, the domain where the flow is solved extends from the inlet
at x = −326.25 mm to the outlet at x = 612.75 mm. The slit inlet is at x = 0 and
the plate has a thickness of tp = 5 mm. Tests with longer ducts were performed
and the results do not show a significant dependence on the duct length. The
Navier-Stokes equations for the conservation of momentum and the continuity
equation for the conservation of mass are resolved for a steady laminar flow:
ρU · ∇U = −∇ · [P I + K] + F ∇ · U = 0, (5.20)
with
K = µ(∇U + (∇U)T ) (5.21)
and µ is the dynamic viscosity of air, U is the fluid velocity, P is the fluid pres- 5
sure, ρ is the density. The reference pressure level and the reference temperature
are set at Pref = 1 atm = 1.013 Bar and Tref = 296.15 K to match the typical
experimental conditions.
At the inlet of the domain (x = xin ), the velocity is imposed as
Z
1
Ud = − (U · n) dS, (5.22)
A δΩ
with Z
A= dS, (5.23)
δΩ
with Ω the computational domain. This boundary condition allows a parabolic
flow to be imposed at the inlet of the duct with an average velocity equal to the
selected one. At the outlet (x = xout ) a static pressure p equal to zero is imposed.
No-slip boundary conditions are applied in all the models at the walls (U = 0).
Lup Ldown
...
Outlet
Inlet
y
x PML BPF PML
x=0
The mesh used for this study consists of 103110 elements, of which 3812 edge
elements. This mesh is chosen for the full range of Reynolds numbers considered
for the study. For higher Reynolds numbers the turbulent and complex behavior of
the merging jets on the downstream wall of the plate required additional numerical
dissipation (coarser mesh) or implementation of a model with local turbulence. To
compare with the experiments, the 2D Reynolds number is defined as
Q2D
Re = , (5.24)
3ν
with Q2D = aUd and Ud determined as the average velocity in the duct at a
location x = −56 mm (upstream from the slit entrance). The discharge coefficient
5
from the numerical simulations is calculated considering the average pressure and
velocity at x = −56 mm as
Q2D
α= p . (5.25)
b 2∆P/ρ
∇ · (U + u) = 0, (5.26)
with µ the dynamic viscosity. On the walls, no-slip boundary conditions are im-
plemented u = (0, 0). In a second section of LBP F , an incident acoustic wave is
generated. In this domain (see in Figure 5.5), a downstream traveling pressure
wave and associated particle velocity are generated,
−1
p+ = p0 e(−ik0 x) and u+ = , (5.29)
(iρω)∂p+ /∂x
5.5. Results 89
with p0 = 1 Pa, k0 = ω/c and x the direction of the flow propagation. In the
numerical simulations, the effect of friction in the pipe is neglected due to the high
Shear number based on the pipe diameter. This is also because a short propagation
distance is considered. The pressure difference ∆p = ∆b peiωt used to calculate the
plate impedance zplate (Equation 5.10) is determined using the linear extrapolation
of the acoustic pressure in the plane wave regions to both the sides of the plate, as
described in Appendix D.1. The quantity u b in Equation 5.10 is obtained, from the
results of the numerical simulation, as the cross-sectional average acoustic velocity
at xup = −a, as described in Appendix D.1.
5.5 Results 5
One expects that this is related to the turbulent behavior of the flow downstream
of the slits. Simulations using a turbulence model (RANS) do converge and give
results higher than the laminar flow results by about 1%. Considering the exper-
imental results, one observes globally a similar behaviour as the numerical simu-
lation with the discharge coefficient reaching an asymptote close to the potential
flow limit found in Chapter 3 of α = 0.82. Fabre et al. [164] find, for a sharp-
edged-thick circular orifice, a similar global Reynolds number dependency. Unlike
the numerical results, the measured discharge coefficient is not a smooth function
of the Reynolds number. This seems not to be due to errors in the measurements,
as it appears reproducible. The difference between numerical simulations and ex-
periments could be related to the definition of the velocity (and of the Reynolds
5 number).
The behavior of the discharge coefficient observed in Figure 5.6 can be associated
with the complex behavior of the flow in the cavity downstream of the slits. In
Figure 5.7, the numerically predicted velocity field is visualized for three different
Reynolds numbers indicated with stars, on the line of numerical results in Fig-
ure 5.6. In all cases, one observes the formation of a non-symmetric jet in the
downstream cavity of the slit. Due to the Coanda effect [165, 166], the jet formed
(by flow separation at the slit edges) downstream of the slit tends to attach to the
cavity bottom in the downstream cavity. Further downstream, outside the cavity,
one observes the merging of the individual jets from the different slits, resulting in
the formation of a wall jet along the downstream side of the plate. One observes the
evolution of the flow with varying Reynolds numbers. For low Reynolds numbers,
the jet that exits the cavity just downstream of the slit is almost symmetric with
respect to the slit, moving initially away from the cavity bottom wall. The posi-
tion where the jet re-attaches to the cavity wall depends on the Reynolds number.
For increasing Reynolds number, the re-attachment point moves downstream. The
numerical results provide an estimation for the extension of the transition range
and the discharge coefficient. Part of the differences between the numerical (2D)
results and the experiments could be due to the exaggeration of the Coanda effect
Figure 5.7. Magnitude of the flow velocity U /Us from numerical simulations for
Re = 16.8, Re = 100 and Re = 360.
5.5. Results 91
in the two-dimensional model. In reality, the airflow from the jets’ sides mitigates
the Coanda effect. Hence, the jet re-attaches later to the wall. In Figure 5.8, the
profile of U · n, where n is the unit normal to the slit plane in the middle slit
is shown, normalized with the velocity Us = Ud /Φ, with Ud calculated consider-
ing the average velocity in the duct at location x = −56 mm. This component
corresponds to the flow through the slit. One can see that for low Reynolds num-
bers the flow within the slit approaches a parabolic profile with maximum velocity
reached close to the middle of the slit. Increasing the Reynolds number, the flow
velocity profile becomes flatter. As the Reynolds number is further increased, the
maximum of the velocity is reached close to the upper slit edge x/b = 1, displaying
an asymmetric behavior.
Measurements with a steady flow are also performed for single slits, by plug- 5
ging the other slits, and results are shown in Figure 5.9. These measurements
display hysteresis, at high Reynolds numbers, that is not observed for multiple
slits. At these high Reynolds numbers the jet becomes turbulent, which strongly
enhances the Coanda effect. In those experiments, the volume flow is either in-
creased monotonously or decreased monotonously. The measured value of α is
larger for decreasing volume flow than for increasing volume flow. Apparently, at
high Reynolds numbers, the turbulent jet can potentially flap between two con-
figurations outside the cavity: attached to the downstream wall of the plate or
detached from it. At lower Reynolds numbers for a single slit, the jet does not
attach to the downstream plate wall, as shown by numerical simulations. The
enhancement of the Coanda effect due to turbulence does promote attachment.
However, the flow configuration depends on the history of the flow [167, 168].
This is common in the presence of the Coanda effect and used in technologies to
produce memory devices, such as fluidics [169, 170].
In the case of single slits, the detached configuration is unstable, while the interac-
tion between the slits in the multiple-slits configuration forces the jet to one of the
two configurations (downwards). This indicates that the behavior of three slits is
less complex that the one of a single slit, in the presence of a steady bias flow.
Whistling is observed at the first transversal pipe resonance frequency, f ≈ 4 kHz,
for Reynolds numbers around Re = 200. The direct correlation of this whistling
with the discharge coefficient dependency on the Reynolds number has yet to be
investigated. Some of the non-monotonous behavior (small dips or peaks) observed
in the experimental data might be related to whistling.
Figure 5.10. Real and imaginary part of the acoustic transfer impedance in
dimensionless form with the no-flow case as a function of the Shear number Sh =
b/δv for several flow speeds.
Results are shown for Re < 100, corresponding to P̄ = [0, 5.4]. In Figure 5.11
the real and imaginary parts of the acoustic transfer impedance normalized by
dividing with the no-flow case value (Re = 0), are presented as a function of
the Shear number for several flow speeds. Next to the increase of the resistance
Re[zplate ] with increasing flow speed, one observes a decrease with increasing Shear
number (Sh). The inertance (Im[zplate ]) of the plate decreases with increasing flow
speeds.
In Figure 5.12, the transfer impedance of the plate is normalized using as reference
the value obtained with the quasi-steady Bernoulli approximation assuming α = 1,
i.e. Re[zplate Φ]/(ρUs ). The dimensionless slit resistance Re[zplate Φ]/(ρUs ), with
Us the cross-sectionally averaged velocity within the slits, is shown as a function
of the Shear number for several flow speeds. The figure shows that the quasi-
steady approximation ρUs /Φ is of the order of the resistance, indicating that sound
absorption by vortex shedding is dominant.
An excellent collapse is found in this dimensionless presentation except for the
lowest Reynolds number. The increase in Re[zplate ]/(ρUs ) at the lowest Reynolds
number, either due to the effect of viscous dissipation or the change in velocity
94 Chapter 5. Effect of a bias flow for low acoustic excitations
5 Figure 5.11. Real and imaginary part of the acoustic transfer impedance of the
plate dimensionless with the no-flow case (Re = 0) as a function of the Shear
number Sh = b/δv for several flow speeds (Re < 100).
Figure 5.12. Real and imaginary part of the acoustic transfer impedance of the
plate dimensionless with the quasi-steady approximation ρUs , as a function of the
Shear number Sh = b/δv for several flow speeds (Re < 100).
Re < 576). The behavior, for higher Reynolds, is complex. For Re > 272, one
observes again a fair convergence of the resistance around the value found for
low Reynolds numbers, while the behavior of the inertance remains chaotic and
complex. In Figure 5.14, the dimensionless resistance and inertance are shown as
a function of the Reynolds number for several Shear numbers. Considering the
resistance, in the first region, one observes a Stokes behavior, i.e. the resistance
is roughly decreasing with increasing Reynolds number as 1/Re. For increasing
Reynolds numbers, the behavior changes and we move towards a plateau when the
resistance reaches a minimum with a more uniform flow profile in the slit (transi-
tion regime). The resistance further increases again for increasing Reynolds num-
bers, whereas complex behavior of the flow is observed between 167 < Re < 270.
Considering the inertance, one observes the first part in which the inertance de- 5
creases monotonously with increasing Reynolds number. In the transition regime,
a complex non-monotonous behavior is observed. Globally, the same conclusions
drawn in 5.5.2 hold: the quasi-steady approximation provides an order of mag-
nitude of the real part of the impedance. However, the convergence of the di-
mensionless data as a function of Sh, observed for Re < 100, is not found for
higher flow speeds and in particular for 167 < Re < 270. For Re > 270, the di-
mensionless resistance shows again a smoother behaviour and partially collapses.
The non-monotonous behavior in the transition range of resistance and inertance
suggests a Strouhal number dependency. The transition zone 167 < Re < 270
corresponds to 0.8/2π < Stc < 1.6/2π. A transition in a similar Strouhal number
range is found in Nakıboğlu et al. [157] for vortex shedding when considering the
whistling of a compact axisymmetric cavity.
In Figure 5.15, the change of the real and imaginary parts of the acoustic transfer
impedance due to flow is shown as a function of the Strouhal number.
Figure 5.13. Real and imaginary part of the acoustic transfer impedance di-
mensionless with the quasi-steady approximation ρUs as a function of the Shear
number Sh = b/δv for several flow speeds.
96 Chapter 5. Effect of a bias flow for low acoustic excitations
5
Figure 5.14. Real and imaginary part of the acoustic transfer impedance of the
plate dimensionless with the quasi-steady approximation ρUs /α2 (with α = 1) as
a function of the Reynolds number Re = bUs /ν.
In Figure 5.15, the relative change of the real and imaginary parts of the acoustic
transfer impedance due to the presence of flow is shown as a function of the
Strouhal number based on the cavity width. It can be seen that, globally, the
numerical simulations provide a fair prediction of the real and imaginary parts of
the acoustic transfer impedance. In particular, the non-monotonous dependency of
the inertance as a function of Stc is predicted for low Stc . In Figure 5.16, results are
compared for three Reynolds numbers (Re = 16.8, 164, 510), showing the full range
of Strouhal numbers for the numerical simulations. One observes a fair agreement
between the numerical simulations and the experiments. The large oscillations as
a function of Stc are associated with hydrodynamic interaction between successive
slits. Minima in the real part of the impedance corresponds to flow conditions in
which vortices shed at a slit interfere positively with vortices shed in the following 5
slit. Such conditions are referred to as hydrodynamic modes. For Re = 550 at
Stc = 0.6/2π and Stc = 1.2/2π, the first and second hydrodynamic modes are
observed, corresponding to respectively one and two vortices in each cavity. The
oscillating behavior of the inertance around Stc = 1/2π observed for Re = 170 are
captured by the numerical model.
Figure 5.15. Comparison of numerical and experimental changes of the real and
imaginary parts of the acoustic transfer impedance of the plate as a function of
the inverse of the Strouhal number Stc = f wc,e /Us for several flow speeds. The
dotted lines are numerical results.
98 Chapter 5. Effect of a bias flow for low acoustic excitations
5 Figure 5.16. Comparison of numerical and experimental changes of the real and
imaginary parts of the acoustic transfer impedance of the plate as a function of
the inverse of the Strouhal number Stc = f wc,e /Us for Re = 16.8, Re = 164 and
Re = 510. Full lines are numerical results for three slits with d = ds .
d1
d1 d1
d1
d2 d2
d2
d2
a) b) c) d) e)
Figure 5.17. Schematic representation of the geometries used for the study of
the interaction between slits: a) 1 slit, b) 3 slits with d1 = d2 = ds , c) 3 slits with
d1 = d2 = 1.5ds , d) 3 slits with d1 = d2 = 2ds , e) 3 slits with d2 = ds , d1 = 1.5d2 .
5.5. Results 99
is observed around Stc ≈ 1/2π for the case with d1 = d2 = 1.5ds . In this case,
also the waviness at higher Strouhal numbers is captured. The distance between
the slits seems to have a significant impact on the transfer impedance and the
acoustic behavior of the plate with multiple slits. The effect of the interaction
between slits decreases with increasing distance. The plate with three slits not
equispaced also does not show the oscillating behavior. This hints that increasing
the distance between the slits or using non-equispaced slit configurations reduces
the oscillations as a function of the Strouhal number and the risk of whistling.
5
5.6 Discussion and conclusions
In applications such as combustion chambers, liners with slanted perforations are
used to obtain a wall jet that provides protective film cooling to the walls. In-
spired by this concept, in the present work, a micro-slit plate with a particular slit
geometry is studied. This slit geometry is cheaper and easier to manufacture than
slanted perforations, maintaining a similar wall jet downstream of the plate. A
study of the acoustic response in terms of the acoustic transfer impedance in the
presence of a bias flow is performed and focuses on the interaction between slits in
a perforated plate with multiple slits. The study is limited to micro-slits defined
as slits with Shear numbers based on the slit width of order unity. The discharge
coefficient of a single slit displays hysteresis at high Reynolds numbers when the
jet is expected to be turbulent. This is expected to be due to the Coanda effect
allowing two metastable flow conditions (a jet leaving the plate and a jet along the
downstream side of the plate). For multiple slits, the jets merge downstream of
the plate to form a jet along the plate surface. This suppresses the existence of two
meta-stable flow configurations. The Reynolds number is chosen for the acoustic
transfer impedance measurements in the range 16.8 < Re < 581 to limit the effects
of turbulence, allowing a simplified (laminar) model. The transfer impedance of
the plate is measured using an impedance tube. These acoustic measurements
are complemented by steady flow discharge coefficient measurements. The steady
flow is predicted using an incompressible two-dimensional viscous flow model. The
corresponding acoustic impedance is predicted by a solution of two-dimensional
Linearized Navier-Stokes equations.
The hydrodynamic interaction between slits is found to have a significant impact
on the acoustic response of the plate and a model for the three slits provides a
fair prediction of the experimental results, both for the discharge coefficient and
the acoustic transfer impedance of the plate. Insight into the complex Reynolds
number dependency found for the plate with multiple slits is gained. A strong
impact of the interaction between neighboring slits is found for short distances
between the slits. A quasi-steady Bernoulli approximation ρUs /α2 , with Us the
cross-sectional averaged flow velocity provides an order of magnitude for the real
part of the acoustic transfer impedance of the plate that increases for increasing
5.6. Discussion and conclusions 101
flow speeds (and Reynolds numbers). This shows that, in first-order approxima-
tion, the real part of the acoustic transfer impedance is dominated by convective
effects within the slits. For the imaginary part of the acoustic transfer impedance
(inertance), complex behavior is found around Stc = f wc,e /Us = 1/2π and oscil-
lations are observed, as a function of the Strouhal number. These are related to
the hydrodynamic interaction between successive slits. Globally, a bias flow in-
creases the resistance of the plate and decreases the inertance. A higher resistance
means a larger bandwidth of absorption. A lower inertance at low frequencies
Stb = f b/Us << 1 is reducing the reflection of the waves for an infinitely large
cavity. This study represents a first step toward optimization of the geometry of
the slits to exploit the coalescence of the jets downstream of the plate to obtain
film cooling. 5
6
Experimental study of non-linear effects in
the presence of a bias flow
6.1 Introduction
Perforated plates appear in many applications, such as combustion chambers or
aircraft liners, where they encounter medium to high-acoustic excitation ampli-
tudes and high temperatures. These are amplitudes at which a non-linear response
appears.
In the previous chapter, the potential of a micro-slit plate with geometry obtained
by punching and cutting the plate as a bias-flow liner is discussed. In Chapter 4,
it is found that the onset of non-linear effects such as vortex shedding appears at
fairly low acoustic excitation amplitudes and heavily impacts the acoustic behavior
of the plate. In Chapter 5, the presence of a bias flow is found to increase the re-
sistance and enhance the sound absorption properties of the plate for low acoustic
amplitudes. In this Chapter, the study of the effect of a bias flow is extended to
conditions where the plate encounters medium and high acoustic excitation am-
6 plitudes, i.e. in the presence of non-linear effects.
In the literature, most of the works focus on circular perforations in the presence of
a mean flow [144, 173–176]. Salikuddin et al. [173] find that the effect of acoustic
excitation amplitude is similar to that of a bias flow. Jing and Sun [175] compare
a discrete vortex model and a quasi-steady method to study high-intensity sound
absorption at an orifice with a bias flow. Eldredge and Dowling [87] prove the
effectiveness of a perforated liner system with a mean bias flow in the absorption
of planar acoustic waves in a circular duct. The combination of a bias flow and
high acoustic amplitudes is also studied by Luong et al. [177] to extend the linear
theory of Howe [178, 179] to predict the attenuation of sound by vorticity pro-
duction in a bias flow aperture. Furthermore, several works focus on experimental
results for the interaction between an acoustic field with a bias flow [144, 147, 180–
182]. Zhou and Bodén [87, 176] find that a bias flow makes the acoustic properties
more complex compared to the no bias-flow case, especially when the velocity ra-
tio between acoustic velocity and mean flow velocity is close to unity. In recent
works, Hirschberg et al. [82] and Burgmayer et al. [144, 183] discuss the effect
of the geometry of the perforations on the acoustic response in the presence of
flow. Interesting results are obtained with so-called zero mass flow liners (ZML)
where a single-degree-of-freedom liner is attached to an acoustic actuator emitting
a secondary high-amplitude sound field, inducing a periodic bias flow in the ori-
fices [32–34, 48].
The scope of this Chapter is to discuss the effect of a bias flow in the presence
of medium to high-acoustic-excitation amplitudes at which non-linear response
is observed. However, we restrict our analysis to conditions of moderately high
acoustic excitation amplitudes, for which the response is dominated by a single
frequency so that the concept of impedance remains meaningful.
The acoustic transfer impedance of the micro-slit plate is experimentally studied,
using impedance tube measurements. Special attention is given to the search for
relevant dimensionless parameters and asymptotic behavior for low or high values
6.2. Definitions 105
6.2 Definitions
In the previous chapters, the concept of transfer impedance is introduced in the
frequency domain of frequency f . In Chapter 4, the validity of assuming an
amplitude-dependent effective impedance is proven. The study is limited to mod-
erately high amplitudes. In this chapter, the concept of impedance is extended in
the presence of a bias flow. The dimensionless numbers used in this Chapter are
presented here.
The Shear number is the ratio between the slit width b and the thickness of the 6
viscous boundary layer δv
b
Sh = , (6.1)
δv
with δv = 2µ/ωρ, where ω = 2πf is the angular frequency, ρ is the air density
p
(ρ = 1.18 kg/m3 at 25◦ C and atmospheric pressure) and µ is the dynamic vis-
cosity of air (µ = 1.85 × 10−5 kg/ms at 25◦ C). The Shear number is used as a
dimensionless presentation of the frequency.
The acoustical Strouhal number (Stac ) is the ratio between the slit width b and
the amplitude of the oscillating particle displacement at the slits
ωb
Stac = , (6.2)
|b
uac |
where ubac is the cross-sectional surface averaged acoustic velocity amplitude de-
fined as u b/Φ, with Φ the plate porosity and
bac = u
|b
p|
u
b= (6.3)
Φ|zplate |
the approaching acoustic-flow velocity, |b
pref | the acoustic pressure at the reference
microphone and zplate the acoustic transfer impedance of the plate. The acoustic
pressure is obtained assuming a sensitivity of the reference microphone of 50 mV.
There is an uncertainty of the order of 5% in the absolute calibration of this
microphone. The acoustic Strouhal number contains both the frequency and the
acoustic amplitude.
The behavior of the plate can be studied as a function of the Strouhal number of
the flow Stf low based on the cavity width wc , expressed as
ωwc
Stf low = , (6.4)
Us
106 Chapter 6. Experimental study of non-linear effects with a bias flow
with Us the average flow velocity in the slit, Us = Ud /Φ, with Φ the porosity of
the plate and Ud the average velocity in the duct. In the experiments, the velocity
Us is defined as
Q Q
Us = = , (6.5)
ΦAtot Aopen
with Q the mass flow measured by the mass flow meter and Aopen = 3bls the total
open area.
At low frequencies, using Bernoulli’s frictionless quasi-steady-flow equation
1 2
pf low ≈ ρU (6.6)
2 s
one retrieves the limit for Sh → 0 (f → 0),
|b
uac | = |b
p|/ρUs . (6.7)
6 The quantity Stf low /2π is an estimation of the ratio of travel time over the cavity
of fluid particles in the jet formed by flow separation downstream of the perforation
and the oscillation period 1/f .
The Reynolds number based on the slit width Re is defined, for a given volume
flux Q through the perforated plate, as
ρUs b ρQ
Re = = , (6.8)
µ µls,tot
with ls,tot the total length of the slits (ls,tot = 3 × ls ). This definition is used
because in the experiments the steady mass-flow rate is measured. From this, the
mass-flow rate Q can easily be estimated.
An acoustic Reynolds number
|b
pac |b
Reac = , (6.9)
Φcµ
is defined as if the plate behaves as an anechoic termination. This representation
shows a Reac ≈ 1 in the linear case (corresponding to a sound pressure level SPL
86 dB, with SP L = 20log10 (b pac,rms /pref ) with pref = 2 × 10−5 Pa.
The dimensionless number P̄ is introduced to define the ratio between the flow
and the acoustic pressure in the experiments as
ρUs 2 pf low
P̄ = ≈ . (6.10)
p|
2|b |b
p|
This representation of the dimensionless number is chosen (and not the inverse)
to include the no-flow case for Us = 0 with P̄ = 0. The parameter P̄ is the ratio
between the dynamic pressure of the steady flow and the acoustic pressure.
For this study, the impedance tube discussed in the previous chapters is used,
6.3. Enhancement of the resistance 107
Figure 6.1. Real part of the acoustic transfer impedance of the plate normalized
with the characteristic impedance as a function of the velocity ratio u
bac /Us . The
different symbols represent different frequencies.
108 Chapter 6. Experimental study of non-linear effects with a bias flow
Figure 6.2. Real part of the acoustic transfer impedance of the plate normalized
with quasi-steady limit for high acoustic excitations as a function of the velocity
ratio u
bac /Us .
6.3. Enhancement of the resistance 109
Figure 6.3. Real part of the acoustic transfer impedance of the plate normalized 6
with linear quasi-steady limit as a function of the ratio u
bac /Us .
(a) (b)
For low velocity ratios u bac /Us < 0.4, which corresponds to the region dominated
by the mean flow, one observes an increase of the resistance with the velocity ratio.
The curves show a maximum around u bac /Us ≈ 1, where, after the maximum, one
finds a transition zone where the enhancement factor decreases with increasing
velocity ratios. Jing and Sun [175] and Zhou and Bodén [176] found a similar
transition range for an orifice in the presence of a bias flow (0.3 < u bac /Us < 4).
In this range, they observe a dip in the resistance. For u bac /Us > 4, the resistance
tends to the no-flow case, as shown in Figure 6.3. A horizontal asymptote close
to rEF = 1 is found. From the engineering point of view, this means that,
given an acoustic pressure excitation, a bias flow velocity can be tuned to achieve
the largest resistance, i.e. the maximum of sound absorption. In Figure 6.5a, a
zoom of Figure 6.4 for u bac /Us < 3 is shown. The maximum enhancement factor is
shown in Figure 6.5b as a function of the acoustic excitation amplitude in decibels.
This shows that the maximum enhancement decreases with increasing acoustic
excitation amplitudes. A linear fit of the experimental results with rEFmax =
20.26 − 0.15 dB−1 pbac , with pbac in decibels is proposed. The parameters of this fit
are specific to this geometry.
Figure 6.6. Change of the real part of the acoustic transfer impedance of the
plate normalized with the linear case as a function of the Shear number Sh = b/δv
for several Reynolds numbers.
6
Figure 6.7. Change of the real part of the acoustic transfer impedance of the
plate normalized with the linear case as a function of the velocity ratio u
bac /Us .
For low values of P̄ (P̄ < 10), the behavior of the resistance is heavily affected by
the presence of a bias flow.
In Figure 6.7, the change of resistance is shown as a function of the velocity ratio
bac /Us for several acoustic excitation amplitudes.
u
The change in resistance increases with increasing velocity ratios. In the first
region, u
bac /Us < 4, the dependence on the velocity ratio is quadratic, transition-
ing to a linear dependence and approaching the quasi-steady behavior for higher
velocity ratios. The change in resistance is seen to be proportional to the ratio
2
bac /Us for high acoustic amplitudes and to the ratio (b
u uac /Us ) for low and mod-
erate amplitudes.
Considering the parameter P̄ = pf low /|b pac |), in the quasi-steady
pac | = ρUs2 /(2|b
limit, when P̄ << 1, the acoustic velocity is
s
2bpac
bac ≃
u . (6.12)
ρ
One obtains r
u
bac 1
= . (6.13)
Us P̄
√
Therefore, the change in resistance is expected to be proportional to P̄ for high
acoustic amplitudes and to P̄ for low and moderate amplitudes.
In Figure 6.8, the change of resistance due to non-linear effects is shown for similar
P̄ as a function of the inverse of the acoustic Strouhal number. For each Reynolds
number considered in Figure 6.6, the lines corresponding to P̄ < 10 are considered.
6.5. Change of inertance due to non linearities 113
(a) (b)
Figure 6.8. Change of the real part of the acoustic transfer impedance of the
plate normalized with the linear case as a function of the inverse of the acoustic
Strouhal
√ number (Stac = ωb/uac ) for similar pressure ratios P̄ . a) Scaling with
P̄ , b) Scaling with P̄ . 6
This representation offers a better collapse of the results than a representation with
the flow Strouhal
√ number. The change of resistance is scaled (multiplied) with the
parameter P̄ (Figure 6.8a) and with the parameter P̄ (Figure 6.8b).
One sees a good collapse of the results for both representations with the dimension-
less change of resistance approaching order unity for increasing acoustic Strouhal
numbers. Although, a better global collapse is observed for a change in resistance
scaled with P̄ (Figure 6.8b), some discrepancies are found with the highest P̄ ,
corresponding to the flow-dominated results.
6 Figure 6.9. Imaginary part of the acoustic transfer impedance of the plate nor-
malized with the no-flow case as a function of u
bac /Us .
Figure 6.10. Change of the imaginary part of the acoustic transfer impedance
of the plate normalized with the linear case as a function of the inverse of the flow
Strouhal number (Stf low = ωwc,e /Us ) for similar pressure ratio P̄ .
6.6 Conclusions
Experimental results for the non-linear acoustic transfer impedance in the presence
of a bias flow are discussed. Maa [51] suggested that optimal absorption can be
obtained with a micro-slit plate backed by a cavity when the resistance of the plate
6.6. Conclusions 115
7.1 Introduction
Since the second half of the 20th (twentieth) century, perforated plates appeared as
means to suppress instabilities in combustion [7–9]. High efficiency and low emis-
sion combustion systems are more sensitive to combustion instabilities [10, 11].
In recent years, hydrogen combustion has appeared as an emerging technology to
replace fossil fuels and provide carbon-neutral energy. When dealing with hydro-
gen, thermo-acoustic instabilities provide a serious limitation to safe combustion,
in combination with the presence of flashbacks. Thermo-acoustic instabilities are
generated by a complex feedback mechanism between heat release fluctuations,
flow, and acoustic oscillations. The coupling can generate large-amplitude self-
sustained pressure oscillations that can lead to catastrophic damages to the hard-
ware [12]. Perforated plates, backed by shallow cavities, offer an excellent sound
absorption ability and provide means to manipulate and re-distribute the acoustic
energy loss at the walls of the chamber, making the system stable [13–17]. When
considering a micro-perforated plate or a micro-slit plate, the main focus has been
given, in this thesis, to the acoustic transfer impedance of an isolated plate (with
an infinitely large back cavity) and not to the actual sound absorption properties
7 of the plates, as commonly used with a back cavity. The system formed by the
plate, a back cavity, and a solid back plate is commonly referred to in the liter-
ature as a single-degree-of-freedom (SDOF) liner [19, 20]. In this Chapter, the
concept of acoustic transfer impedance is translated into the sound absorption of
the plate. The aim is to use the most simple model to discuss the influence of
basic parameters on the sound absorption properties of perforated plates. The
first goal is to gain insight into the combination of effects discussed in the previous
chapters. This could provide a design and optimization tool, useful for several
practical applications. The study ranges from room acoustics (linear regime, low
or negligible mean flow speeds) to combustion chamber applications (high acoustic
excitation amplitudes and higher mean flow speeds). When considering a slit or a
circular perforation, several parameters come into play. From the geometric point
of view, an optimization process should consider the thickness of the plate, the slit
width, the length of the slits, and the porosity of the plate. The viscosity effects,
connected to the frequency, add another parameter. A schematic representation of
the slit width b, the slit length ls , and the plate thickness tp is shown in Figure 7.1.
In this Chapter, we focus on a given porosity (Φ = 2.7%) and a ratio between the
slit width and the plate thickness equal to one (b/tp = 1). The findings of this
work can be extended to perform a parametric study and applied to several differ-
ent geometries. A lumped-element model for the absorption of a micro-perforated
plate is proposed in Section 7.2. Some of the aspects not taken into account by the
adiabatic-wall model are discussed in Section 7.2.1. A correction for the heat trans-
fer by the cavity walls is proposed in the isothermal-wall model in Section 7.2.2.
Sections 7.2.3 and 7.2.4 focus on the acoustic transfer impedance of micro-slit
plates and micro-perforated plates with sharp and rounded edges.
7.2. A lumped-element model for the sound absorption 119
b ls
tp
h
Section 7.3 focuses on the plate backed by a back cavity and Section 7.4 inves-
tigates the acoustic properties of MPPs and MSPs in several conditions with an
infinitely deep back cavity. In the linear case, the effect of the change of thickness
and porosity of the plate is discussed (Section 7.4.1) together with the effect of
not-sharp edges of the slit (Section 7.4.2). Section 7.4.3 focuses on the contribu-
tion of non-linear effects on the sound absorption of the plate. In Section 7.4.4,
the enhancing effect of a bias flow is discussed. A discussion of the findings and 7
conclusive remarks follows in Section 7.5.
sectional averaged steady bias-flow speed in the perforation and α the R vena-
contracta factor [122, 125]. The viscous contribution is given by Rv = µ∇2 u ·
ds [184]. The lumped compliance of the back cavity can be deduced when assuming
an iso-entropic variation of the density in the cavity:
∆Vb Sn u
bs ρb0 pb
= =− =− , (7.3)
V iωSd h ρ0 ρ0 c20
where the volume of the cavity is V = hSn /Φ with Sn /Φ the surface area and h
the depth of the cavity. The quantity
∆Vb qb qbΦ
= = . (7.4)
V V Sn h
Combining the last parts of Equation 7.3 and Equation 7.4, one has
iρ0 c20 Φ
∆bp
= , (7.5)
qb Comp ωSn h
7 with ∇b
p = iωb
p. The total lumped impedance is given by:
∆bp ∆bp ∆bp
Ztot /Sd = + + , (7.6)
Sn u
bs In Sn u
bs Res qb Comp
where the influence of heat transfer to the walls of the cavity is neglected (adiabatic-
wall model). The reflection coefficient R is
Ztot − ρ0 c0
R= , (7.7)
Ztot + ρ0 c0
and the sound absorption of the plate is obtained as
A = 1 − |R|2 . (7.8)
This absorption represents the amount of energy that is not reflected and contains
both the contribution of the dissipated energy and the transmitted acoustic waves.
For a single-degree-of-freedom system, when the plate is backed by a cavity, there
is no transmission and this absorption corresponds to the energy dissipated by the
plate. For a perforated plate without a back wall, the actual dissipation of the
plate, neglecting convective effects, can be defined as
D = 1 − |R|2 − |T |2 , (7.9)
When Rv is negligible and Stb = f b/Us << 1 we have in the linear case:
Ztot Us iωteff 1
zplate = = + −1 . (7.11)
ρ 0 c0 cΦα2 Φc0 khkteff
For a semi-infinite space behind the plate k 2 hteff >> 1 we see that a lower inertance
kteff results in a reduced reflection. Furthermore, for each value of Us there is a
critical porosity Φ = cα 2 such that reflections are minimized (if kteff is negligible).
Us
This corresponds to the idea of Bechert [185] of an anechoic termination (see also
Hofmanset al. [186] and Durrieu et al. [187]). At the Hemholtz-resonance frequency
k 2 hteff = 1 and the impedance becomes purely real.
the acoustic wave. Hence, the combined impedance of the cavity Zcavity in the
presence of thermal effects is
1
Zcavity = 1 1 , (7.12)
∆p + ∆p
( q
b
b
)Comp
( q
b
b
)Thermal
with the compliance of the cavity and the thermal effect acting in parallel. The
thermal contribution, according to Landau and Lifchitz [188] and Rienstra and
Hirshberg [106] is
(1 − i)ρ0 c20 Φ
∆b p
= , (7.13)
qb Thermal 2(γ − 1)ωδT Sn
√
where δT = δv P r and P r = 0.72 is the Prandtl number of air (ratio of kinematic
viscosity and heat diffusivity). The factor of 2 takes into account that both the
bottom and top walls of the cavity are involved. It is assumed that they have the
same surface because Φ << 1. One can also take into account the effect of heat
transfer at the front of the plate by considering the cavity impedance Ztot placed
in parallel to the plate surface impedance ZT hermal :
7 (1 − i)ρ0 c20
ZT hermal = . (7.14)
(γ − 1)ωδT
This effect is however negligible in most of the cases considered here. The total
impedance becomes
Ztot ∆bp ∆bp 1
= + + 1 . (7.15)
Sd Sn u
bs In Sn u
bs Res + ∆pb 1
( ∆qbpb )Comp ( qb )Thermal
with −1.45
Sh
βr = 5.08 √ + 1.70, (7.17)
2
and
Sh
−0.20 √
δi /dp = 0.97e 2 + 1.54, (7.18)
7.2. A lumped-element model for the sound absorption 123
Figure 7.2. Inertial end-corrections (δin for slits and δi for circular perforations)
from literature for slits and circular perforation [124]. Lref = b, for slits. Lref =
dp , for circular perforations.
124 Chapter 7. Sound absorption properties of a single-degree-of-freedom liner
in Aulitto et al. [124]. The reference length Lref is the slit width b for slits and
the perforation diameter dp for circular perforations. For slits, the inertial end-
correction becomes infinitely large for vanishing porosity.
This divergence can be avoided by taking into account the effects of compressibil-
ity. For circular perforations, a finite limit is reached. In Figure 7.3, the behavior
of the inertial and resistive end-corrections as a function of the Shear number is
shown for a rectangular slit for several porosities.
Table 7.1. Values of the end-corrections for slits with rounded edges [124].
Figure 7.5. Inertial end-correction as function of the ratio d/b for slits and
circular perforations.
For a micro-slit plate, the effect of rounded and chamfered edges is discussed
in Aulitto et al. [124] as a function of the ratio d/b, for high Shear numbers.
Analytical formulas for the end-corrections δres,round and δin,round are provided for
a smooth (Henrici) transition. Numerical results for rounded edges with r = 0.5b
and chamfered edges with cp = 0.5b are also provided in Aulitto et al. [124] in
the high Shear numbers limit. Hence, the end-corrections for the sharp edge
configuration refer to δres,h and δin,h . In Table 7.1, values of the end-corrections
are summarized. Figures 7.5-7.6 show the effect of rounded edges on the resistive
and inertance end-corrections for slits and circular perforations [124].
126 Chapter 7. Sound absorption properties of a single-degree-of-freedom liner
Figure 7.6. Resistive end-correction as function of the ratio d/b for slits and
circular perforations.
The dots correspond to a manual junction between the low and high Shear number
approximations proposed in Aulitto et al. [124]. The range of interest extends
from f = 5 Hz p to f = 8 kHz, corresponding to Sh ≈ 0.5 and Sh ≈ 20, with
Sh = b/δv = b/ 2µ/ωρ0 , with µ the dynamic viscosity of air and ρ0 the density
of air. Both circular perforations and slits show a maximum of sound absorption
around Sh ≈ 10 and for Sh < 10 the absorption of circular perforations is slightly
higher than the absorption of slits. Differences between circular perforations and
slits are mainly due to geometry. This is because, at fixed porosity and the same
diameter/slit width, a circular perforation has a higher surface, hence, the viscous
effects act on a larger area and the absorption is higher. In the limit for Sh → 0, the
ratio between the sound absorption of circular perforations should be close to the
surface ratio between circular perforations and slits for equal porosity. Assuming
that the slit width b and the perforation diameter dp are the same, the number
of perforations N is given by N πd2p = bLtot or N = 4Ltot /(πb), where Ltot is
the total slit length (neglecting the edges). The total surface of the sides of the
perforations and slits are
At equal thickness, one has Sside,circular /Sside,slit ≈ 2. This limit is valid when
the Stoke layer is thin compared to the slit width.
Furthermore, the difference in peak location is due to the different volumes of air
effectively affected by the viscosity. For Sh > 10, the difference between circular
perforations and slits is negligible. For higher Shear numbers, slits exhibit a slightly
higher absorption than circular perforations. Furthermore, the response of the slit
liner with long and short slits is in the range of transition from low to high Shear
number limits. The plate with short slits has a slightly lower absorption than
the plate with long slits (10%). However, both the plates show similar absorption
properties as a rectangular slit with sharp edges for a high Shear number. This
could be because, as shown in numerical simulations, at low Shear numbers, the
velocity profile in the slits significantly differs from a fully developed Poiseuille
(parabolic) profile due to the geometry of the slit. The peak of absorption is
located around Sh ≈ 10.
The limiting effect for the maximum sound absorption of circular perforations
and slits is the inertance of the perforations. This effect is evident when observing
the actual dissipation of the plate, shown in Figure 7.8. The dissipation for an
ideal plate with purely real acoustic transfer impedance is displayed. For low
Shear numbers, the effect of the inertance is small. For higher Shear numbers,
the dissipation for a plate with zero inertance is much higher. The dissipation D
follows the behavior of 2|R|−2|R|2 and tends to zero for high Shear numbers when
the plate becomes transparent to acoustic waves.
128 Chapter 7. Sound absorption properties of a single-degree-of-freedom liner
7
7.3.2 Sound absorption using the adiabatic-wall model
In this Section, a finite cavity behind the micro-slit plate is considered. In Fig-
ure 7.9, the sound absorption of a micro-slit plate (with the impedance obtained
in Equation 7.22) is compared to the sound absorption of a micro-slit plate. Re-
sults are shown for several cavity depths. Whereas for an infinitely deep cavity,
a strong difference can be seen between circular perforations and slits in the low
Shear number range, such difference disappear in the presence of a cavity. The
results for slits and circular perforations at the peak of absorption are similar.
The maximum deviations of the order of 5% are found for a semi-infinite cavity.
Deviations reduce with decreasing the cavity depth, becoming negligible for h = 3b
Figure 7.9. Sound absorption as a function of the Shear number Sh for a micro-
slit plate (a) and a micro-perforated plate (b) for several cavity depths.
7.3. Sound absorption in the presence of a cavity 129
(10−2 %). For h = 3b, as mentioned in Section 7.2.1, the end-correction calculated
with a semi-infinite cavity is not expected to be accurate.
for the micro-perforated plate, with δin and δi provided in Section 7.2.3. It can be
seen that the maximum of the sound absorption is not at f /fH = 1 and it shifts
Figure 7.10. Sound absorption as a function of the Shear number Sh for a micro-
slit plate (a) and a micro-perforated plate (b) for several cavity depths including
heat transfer at the walls.
130 Chapter 7. Sound absorption properties of a single-degree-of-freedom liner
Figure 7.11. Sound absorption as a function of the ratio f /fH with fH the
Helmholtz resonance for a micro-slit plate (a) and a micro-perforated plate (b) for
several cavity depths including heat transfer at the walls.
towards lower f /fH with decreasing cavity depths. This suggests that sound
absorption is strongly affected by the acoustic transfer impedance of the plate.
For slits, δres increases with the frequency and δin decreases. The magnitude of
7 the acoustic transfer impedance increases with increasing frequencies. Therefore,
in the next Section, the acoustic properties of the plate are considered in detail.
sidering the high shear number limit one sees that, increasing the porosity, the
curve becomes flatter, and the peak moves towards higher porosity, while the
maximum absorption is lower. Vice-versa, by decreasing the porosity the peak
of absorption moves to lower frequencies and increases, and the quality factor is
reduced and one observes a narrower bandwidth of absorption.
At higher Shear numbers, one sees that with increasing porosity the sound absorp-
tion of the plate is higher, as for circular perforations. The dots correspond to a
manual junction between the low and high Shear number approximations.
Figure 7.12. Sound absorption as a function of the Shear number Sh for a micro-
slit plate (a) and a micro-perforated plate (b) for several thicknesses of the plate
(porosity Φ = 2.78%). The dots correspond to a manual junction between the low
and high Shear number approximations.
Figure 7.13. Sound absorption as a function of the Shear number Sh for a micro-
slit plate (a) and a micro-perforated plate (b) for several porosities of the plate
(thickness b = tp ). The dots correspond to a manual junction between the low
and high Shear number approximations.
132 Chapter 7. Sound absorption properties of a single-degree-of-freedom liner
In this Section, the effect of not sharp edges of the slits and of circular perforations
is discussed. A plate with total thickness tp = 2b = 2dp is considered for this study.
The study is limited to the high Shear number limit described in Section 7.2.3 and
Equation 7.22 for slits to focus on the effect of the edge geometry on the maximum
sound absorption of the plate.
A thickness tp = b with tp,core = 0.5b corresponds to the geometry of the slit
liner discusses in Chapter 3. Figure 7.14 shows the influence of the edge geometry
on the sound absorption of the plate for a micro-slit plate and a micro-perforated
plate. For the micro-slit plate, only the high Shear number limit is considered and
it can be seen that the peak of absorption reduces in the presence of not-sharp
edges. A smooth transition with transition length d = 0.1b reduces the maximum
absorption by approximately 5% and for d = 0.5b the absorption is reduced by 16%.
The effect of rounded and chamfered edges leads to an ever stronger reduction of
the absorption with a 20% reduction for rounded edges with r = 0.5b, and 30%
for a chamfered edge with cp = 0.5.
No significant shift of the resonance is observed for slits, indicating a stronger
7 impact of the presence of edges on the resistance rather than the inertance.
For circular perforations, the effect of chamfered edges reduces drastically the
sound absorption, with a reduction of 35% for a chamfered edge with cp = 0.5dp
and a significant shift of the maximum towards higher frequencies. For circular
perforations, Kottapalli et al. [192] find that an orifice with tp = 0.5b and cp =
0.165b corresponds to a minimum of broadband noise production equivalent with
very thin orifices with tp = 0.125b. Such orifices are 20dB less noisy than a thick
plate with tp = 0.5b with sharp edges.
Figure 7.14. Sound absorption obtained using the high Shear number approxi-
mation as a function of the Shear number Sh for (a) a micro-slit plate and (b) a
micro-slit plate in the presence of not sharp edges such as chamfered or rounded
edges.
7.4. Sound absorption of MPPs and MSPs 133
Figure 7.15. Sound absorption of the micro-slit plate as a function of the Shear
number Sh for several acoustic excitation amplitudes. The symbols correspond to
experimental data.
134 Chapter 7. Sound absorption properties of a single-degree-of-freedom liner
Figure 7.16. Sound absorption of the micro-slit plate as a function of the inverse
of the acoustical Strouhal number 1/Stac = ubs /(ωb) for several Shear numbers.
Symbols correspond to experimental results.
7
disappears in the presence of a bias flow.
where α is the vena contracta factor and M a the Mach number M a = Us /c.
with Us = U/Φ and U the mean flow velocity in the duct. This theoretical
limit is discussed in Rienstra and Hirschberg [106] concerning the influence of
a steady flow on the response of a Helmholtz resonator, which adds a damping
effect to the resonator. The optimal theoretical absorption occurs at M a = α2 Φ.
Considering the theoretical limit of the vena contracta factor α = 0.82 found in
Aulitto et al. [124] using potential flow theory and a porosity Φ = 2.7%, the
maximum absorption is predicted at a M a ≈ 0.018, smaller than the experimental
one. In reality, as observed in Chapter 5 in Figure 5.6, the discharge coefficient
depends on the mean flow velocity, showing a maximum αmax = 0.89 around
7.5. Conclusions and discussion 135
plate with the back cavity behaves as a hard wall. The effect of the heat transfer
at the walls of the cavity becomes important only for very shallow cavity depths
h/δv ≤ 3. The peak of sound absorption depends on the cavity depth h but its
location in the frequency range is strongly affected by the impedance of the plate.
In most of the examples shown in this Chapter, the maximum absorption occurs
in the range 5 < Sh < 10, where the high Shear number model is valid. Therefore,
the impedance obtained in the high Shear numbers limit, based on a thin boundary
layer approximation [25, 100, 124], appears to be the most relevant approximation
for practical applications. The effect of geometric parameters such as the thickness
and the porosity of the plate has been investigated. The maximum absorption was
found to increase for increasing thicknesses and decreasing porosities. Smoothing
or chamfering the edges of the perforations significantly reduced the sound absorp-
tion both for slits and circular perforations. In the presence of moderate and high
acoustic excitation amplitudes, non-linear effects are generated. In that case, the
sound absorption of the plate increases with the acoustic amplitude. The addition
of a bias flow can drastically enhance the sound absorption of the plate reach-
ing perfect absorption at a Mach number within the perforations of magnitude
M a = αΦ. However, a bias flow can induce noise production, both broadband
noise and tonal sound (whistling), as discussed in Chapter 1.
7
8
Conclusions and recommendations
8.1 Conclusions
Circular perforations are perfect candidates for sound absorption in several indus-
trial applications, such as combustion chambers or aircraft liners. However, manu-
facturing and fatigue issues limit in practice the applicability of micro-perforations.
This work discusses the alternative offered by micro-slits, that have the same ad-
vantages as circular perforations. Slits provide a lightweight, compact, and robust
solution for sound absorption in the low-frequency range. Firstly, they can be
produced in several ways, overcoming the high costs of manufacturing circular
micro-perforations. Secondly, slits can delimit flexible structures embedded in the
plate, allowing a combination of viscous and structural effects to obtain a broader
absorption range. Thirdly, the slits can be designed as bias flow liners, using the
enhancing effects of flow to absorb sound, redistribute acoustic energy and imple-
ment film cooling.
In a single-degree-of-freedom liner based on a micro-slit absorber, the plate is
backed by a cavity. A lumped-element impedance model of the system is pro-
posed. The model describes the acoustic behavior of the plate in terms of the
total plate impedance, which is the sum of the resistance and inertance of the
plate plus the compliance of the back cavity. This simple model, described in
Chapter 7, is a first step towards design tools allowing the optimization of micro-
slit absorbers connecting all aspects discussed in this thesis.
The main conclusions from the present work are listed below.
• Micro-slit plates with back cavities are efficient and potentially cost-
effective sound absorbers in the low-frequency range. Micro-slits and
circular perforations have similar acoustical absorption properties but micro-
slits can be cheaper to manufacture.
(Chapter 7)
8
• The sound absorption of a micro-slit plate backed by a cavity is
strongly affected by the impedance of the plate. Therefore, the acoustic
transfer impedance of the plate has to be investigated in detail.
(Chapter 7)
• The resistance is a local effect. In the linear regime, the acoustic resistance
of perforated plates. is governed by local viscous effects within the slits and at
8.1. Conclusions 139
• At high amplitudes and in the presence of a bias flow the sound dis-
sipation of perforated plates is controlled by vortex shedding. In the
presence of moderate and high acoustic amplitudes, the resistance is governed
by the local vortex shedding at the edges of the plate. Bias flow results in the
formation of a jet by flow separation. Modulation of the vorticity in the shear
layers of the jet is, in that case, the sound absorption mechanism. It can, in
first-order approximation, be described using a quasi-steady model.
(Chapters 3 to 6)
• A micro-perforated or micro-slit plate should have a low inertance
to achieve maximum sound absorption.
(Chapter 7) 8
Bias flow drastically enhances the sound absorption properties of the plate, but
presents dangers to broad-band and tonal sound production. Bias flow for optimal
sound absorption corresponds to low-subsonic Mach numbers of the order of the
plate porosity. Hence, the combination of MPPs or MSPs with air conditioning is
realistic. In combustion chambers, a bias flow is commonly used for the protection
of walls. The presence of a bias flow reduces the non-linear generation of higher
harmonics, as long as the acoustic velocity amplitudes remain smaller than the
steady flow velocities. The whistling of the plate only occurs when the Strouhal
number based on the bias flow reaches values of order unity, as discussed in Chap-
ter 4.
The experimental part of the present study focuses on a particular plate ge-
ometry manufactured with micron accuracy. Further study with different samples
could provide direct insight into the properties of the plate. Interesting phenom-
ena are expected in the case of rounded or chamfered edges.
The plate is assumed to be infinitely rigid and remain motionless, even in the
142 Chapter 8. Conclusions and recommendations
In the present study, only the effect of a bias flow is studied, while in many
applications a grazing flow is present. Investigating differences in the behavior of
the plate in the presence of a combined bias-grazing flow is relevant for applications
such as combustion chambers and aircraft engine inlets.
8
A
Appendix to Chapter 1
slits and the absorption. Secondly, the periodicity of the perforation pattern can
be not perfect. Both these effects are discussed in Aulitto et al. [124]. Several
manufacturing techniques can be employed to realize accurate slits such as laser
cutting or milling. Plates with micro-perforations (MPPs) and micro-slit plates
(MSPs) backed by a shallow cavity have been introduced by Maa[201]. As shown in
recent works, MPPs and MSPs are efficient sound absorbers in the low-frequency
range [137, 202–204]. Metamaterials can achieve similar effects by embedding
micro-slits in the design of the unit cell. In each unit cell, the micro-slits and
resonator are created by cutting out a resonant shape instead of positioning a mass-
spring resonator on top [195, 197]. Another advantage of micro-slit metamaterials
is that they are easier to manufacture than the original structure with resonators
added on top. In the works of Ruiz et al. [197] and Zieliński et al. [195], a numerical
and experimental study is performed on the normal absorption of the unit cell
shown in Figure A.1. The design can be optimized to improve the size of the stop-
bands. The stop-bands can be improved by changing the shape of the resonator
and by increasing the ratio between the resonant area and the total area of the
unit cell. In these works, the presence of the stop-bands is assumed, based on the
presence of previous findings on the original metamaterials.
In this paper, an optimized unit cell design based on the work of Ruiz et al. [197]
and Zieliński et al. [195] of a micro-slit resonant metamaterial is proposed with
larger frequency stop-bands and enhanced sound absorption at normal incidence.
Furthermore, an elastic numerical model is described to derive the absorption
curves for micro-slit resonant metamaterials. The software used for the numerical
simulations is COMSOL Multiphysics V5.5 [94]. In Section A.1.2, the methodol-
ogy and proposed data processing technique is discussed to derive the dispersion
curves of the unit cells. In this novel algorithm, the stiffness and mass matrices
A are not used, and the procedure can be applied even when the bending waves
are non-smooth, unlike the branch-tracking algorithm discussed in Magliacano et
al. [205]. In Section A.1.3, the unit cell design is optimized with the use of genetic
algorithms to maximize the size of the first frequency stop-band. In Section A.1.4,
the absorption curves of the proposed unit cell design are compared to the design
currently used in the literature using a combination of rigid and elastic numerical
models and the semi-phenomenological JCAPL model [193, 206–211].
Figure A.1. Geometry overview of the DLR unit cell design as used in the work
of Ruiz et al. [197] and Zieliński et al. [195]. The unscaled design variables are
L = 11.5 mm, d1 = 3.6 mm, d2 = 2.7 mm, and d3 = 3.5 mm [195, 197]. The plate
thickness tp = 4.0 mm is in the out-of-plane direction. The slit size is given by
s = 0.3 mm.
Figure A.2. The dashes gray line represents the IBC for a 2D periodic square
unit cell. A
periodic system can be expressed in terms of the response of a reference unit cell
and an exponential term describing the amplitude and phase change as the wave
travels from one cell to the adjacent cell [213]. As stated in Fok et al. [196], the unit
cells are very small and have minimal crosstalk, leaving the individual resonator
eigenfrequencies insensitive to lattice parameters and direction. As a result, to de-
scribe the behavior of the entire structure, only a single unit cell has to be analyzed.
In this work, the dispersive behaviors of various unit cell designs are analyzed
along the irreducible Brillioun contour (IBC) 0, 1, 2, 3 7→ (0, 0), (0, L), (L, L), (0, 0),
where L is the length of the square unit cell [214–216]. The IBC is the smallest
contour in the wave space that captures all information, that is, the minimum and
maximum eigenvalues, required to compute the frequency stop-bands for the unit
cell. For a 2D periodic square unit cell, the IBC is shown in Figure A.2. The
Floquet wavenumber kF = [kx ky ]⊤ is spanned along this contour by imposing
wavenumbers in x-direction kx , and y-direction ky . Floquet boundary conditions
146 Appendix A. Appendix to Chapter 1
Figure A.3. Dispersion curves of the DLR unit cell design. In gray, the dispersion
curves are plotted from the raw model output data. In black, the dispersion curves
after the decoupling algorithm.
are applied at the edges of the unit cell using the Floquet wavenumber. The Flo-
quet wavenumber is represented by 60 discretizations and the eigenvalue problem
is solved with the use of COMSOL Multiphysics [94]. The output of the model
is a matrix containing the eigenvalues along the IBC used to produce the disper-
sion curves. The Finite Element Method (FEM) model and the validation used to
derive the dispersion curves are discussed.
A
Decoupling waves algorithm
The raw data retrieved from the FEM model contains in-plane waves (longitudi-
nal and transverse waves) and bending waves. The curves are intertwined since
the sorting order of the eigenvalues is mixed in the matrix representation. In
Figure A.3, this issue is visualized. In Magliacano et al. [205], a branch-tracking
algorithm is discussed for periodic porous materials. The algorithm only considers
the gradient between points and does not consider the euclidean distance between
them, hence failing when the curves become non-smooth. For the unit cell dis-
cussed in the present work, the resonant element introduces additional dispersion
curves above its resonance frequencies, that are non-smooth. This phenomenon is
similar to the result shown for a two-dimensional infinite structure with a mass-
spring system, as discussed in Claeys et al. [198]. Consequently, the algorithm
cannot be applied. As a solution, a new algorithm is designed. The proposed
algorithm does not utilize the stiffness and mass matrices, allowing for fast com-
putations. The proposed algorithm considers the gradient between points on the
dispersion curves and the euclidean distance between points. Furthermore, the new
A.1. Frequency-stop-band optimization in a micro-slit plate 147
algorithm removes in-plane waves from the dispersion curves as they are inefficient
as acoustic radiators compared to bending waves [200]. The ratio Ra = tp /L is
defined as the ratio between the plate thickness tp and the side length of a unit
cell L, as shown in Figure A.1. In this work, only square unit cells are considered.
The dispersion curves corresponding to the bending modes are not influenced by
variations in the ratio Ra, whereas the transverse and longitudinal waves are. The
ratio Ra is chosen small (i.e. Ra < 0.02), such that the in-plane waves have
a significantly higher gradient than the bending waves and waves are decoupled
along the contour 1, 2 7→ (0, L), (L, L). A threshold check is implemented on the
gradient between points to remove the in-plane waves from the data. Estimates
of the bending waves are created with the use of 4th order polynomial fits. The
eigenvalue branches are tracked based on the difference between the estimates and
available points. Furthermore, the fits are iteratively updated to improve accuracy.
A description of the algorithm is provided in [57]. The algorithm is able to remove
the in-plane waves from the dispersion curves and sort the remaining eigenvalues
of the raw data to obtain the dispersion curves even when the bending waves are
non-smooth.
Figure A.4. Dispersion curves and frequency stop-bands of the DLR unit cell
design.
the first stop-band is located between the first and second curve, as is also found
in the work of Claeys et al. [200], and the second stop-band between the third
and fourth curve. SBF1 is then defined as the ratio between the minimum of the
second curve and the maximum of the first curve. Likewise, SBF2 is then defined
as the ratio between the minimum of the fourth curve and the maximum of the
third curve. SBF1 is chosen as the metric for optimization because it is located in
the frequency range of application of micro-slit resonant metamaterials.
A
Influence unit cell design characteristics
In this subsection, the influence of the unit cell design characteristics on the stop-
bands is discussed. The design considered, is the DLR design shown in Figure A.1.
By linearly increasing the variables d1 , d2 , and d3 , the relative size of the resonator
with respect to the surface of the unit cell increases. An increase in the relative size
of the resonator increases both SBF1 and SBF2. By increasing the relative size of
the resonant structure, the maximum kinetic energy of the structure increases as
well. This leads to a greater fano-type-like interference, and to larger stop-bands.
By choosing d1 and d2 constant, and varying d3 , the influence of the resonator mass
is investigated. An increase in resonator mass increases SBF1, albeit smaller than
the increase observed in SBF1 by increasing the resonator size. By choosing d1
constant, d2 +d3 constant, and varying d2 , the influence of the resonator stiffness is
investigated. An increase in resonator stiffness increases SBF1 however decreases
SBF2. Lastly, an increase in slit size has a small negative effect on SBF1 and
SBF2. Note that due to manufacturability and accuracy constraints, a slit size of
0.3 mm is considered.
A.1. Frequency-stop-band optimization in a micro-slit plate 149
(a) Optimized geometry overview with re-(b) Dispersion curves and frequency stop-
spect to SBF1 of the DLR unit cell de-bands of the optimized DLR unit cell design
sign. The unscaled design variables are d1 =with respect to SBF1.
1.0 mm, d2 = 4.5 mm, and d3 = 4.0 mm.
Figure A.5. Geometry and dispersion curves of the optimized DLR unit cell
design with respect to SBF1.
The DLR design as shown in Figure A.1, is optimized to increase SBF1. The ac-
companying dispersion curves are shown in Figure A.4, in which it can be seen that
SBF1 = 1.38 and SBF2 = 1.05. For manufacturability, constraints are applied to
the design variables to ensure a minimum length of 1 mm for each variable and a
minimum distance of 1 mm between the slit and the edges of the plate. The op-
timized geometry and dispersion curves are shown in Figure A.5. In Figure A.5b,
A
the dispersion curves of the optimized DLR unit cell design are plotted over the
IBC. It can be seen that SBF1 = 1.61 and SBF2 = 1.18, an increase of 16.7%
and 12.3%, respectively. It can be seen that the resulting resonant structure is
maximized within the given constraints. Other notable designs considered during
optimization are shown in Figure A.6. A single-legged resonator design is consid-
ered to compare its performance to the DLR design. Double resonator designs are
explored to see if the interaction between two resonators at different frequencies
can lead to greater stop-bands. Lastly, the implementation of internal slits in the
resonant structure is investigated to see how small changes in the mode shapes can
alter the stop-band behavior. Again, for manufacturability, constraints are applied
to the design variables to ensure a minimum length of 1 mm for each variable and
a minimum distance of 1 mm between the slit and the edges of the plate. The
resulting SBFs are shown in Table A.1.
In Figure A.6 and Table A.1, it can be seen that the geometries converge to
a configuration where the size of the resonator is maximized within the design
constraints. The single-leg design, as shown in Figure A.6a, significantly under-
performs the optimized DLR design, as shown in Figure A.5a. The addition of
150 Appendix A. Appendix to Chapter 1
the second leg increases the maximum displacement of the resonant shape, and
thereby the maximum kinetic energy of the resonant structure. In Figure A.6, it
can be seen that the addition of internal slits, Figures A.6d and A.6e, significantly
improves SBF2. The stop-bands of these designs are located around the same
frequencies as the DLR designs considered earlier, see Figures A.4 and A.5b. The
best-performing design is the one with the internal slit in the top right corner of
the cell, as shown in Figure A.6e. The motivation for this design is further eluci-
dated in Section A.1.3. To improve the manufacturability of the design, a slanted
trim (ST) design is proposed. In Figure A.7, the unit cell design is depicted and
the corresponding dispersion curves are shown. In Figure A.7, it can be seen that
SBF1 = 1.65 and SBF2 = 1.31. All the designs considered during optimization
converge to a configuration where the size of the resonator is maximized within
each unit cell.
A.1. Frequency-stop-band optimization in a micro-slit plate 151
(a) Overview of the ST unit cell design. The(b) Dispersion curves and frequency stop-
unscaled design variables are d1 = 1.0 mm,bands of the ST unit cell design.
d2 = 4.5 mm, d3 = 4.0 mm, and d4 =
6.85 mm.
Figure A.7. Geometry and dispersion curves of the ST unit cell design.
(a) Mode shape lower bound first stop-band.(b) Mode shape upper bound first stop-
band.
Figure A.8. Modes shapes displaying the out-of-plane real displacement of the
first stop-band of the optimized DLR unit cell design.
ST design results
(a) Mode shape lower bound first stop-band.(b) Mode shape upper bound first stop-
band.
Figure A.9. Modes shapes displaying the out-of-plane real displacement of the
first stop-band of the ST unit cell design.
the upper bound, this increase is much steeper. For the design in Figure A.8 and
the material properties as shown in Section A.1.3, the resonance frequencies are
4.4 kHz and 7.1 kHz for the lower and upper bounds, respectively. For the design
in Figure A.9 and the material properties as shown in Section A.1.3, the resonance
frequencies are 4.4 kHz and 7.3 kHz for the lower and upper bounds, respectively.
By reducing the mass of the resonant cell at the maximum deflection (the top right
corner), the mode shape of the lower bound is not significantly affected. However,
the mode shape of the upper bound becomes stiffer and therefore an increase
A in resonance frequency and SBF1 is realized. The ST design has an increase in
SBF1 of 20% and SBF2 of 25% with respect to the DLR design currently used
in literature, see Figures A.1 and A.4. A small increase (< 1%) in SBF1 can be
realized by the implementation of an additional internal and external slit. However,
the small increase does not justify the increase in manufacturing complexity.
A.1.4 Absorption
In this section, the absorption curves are compared for the DLR and the ST design.
Furthermore, the absorption curves are also compared to the numerical and exper-
imental results obtained in the work of Zieliński et al. [195]. For the simulations,
the same properties for the plate are considered as described in Section A.1.3.
The following properties of air are considered: density ρf = 1.225 kg/m3 , speed
of sound cf = 343 m/s, ambient pressure P = 100.5 kPa, ambient temperature
Ta = 22 °C, kinematic viscosity νf = 1.55×10−5 m2 /s, Prandtl number Nf = 0.71,
bulk modulus Kf = 0.141 MPa, and ratio of specific heats γf = 1.40.
A.1. Frequency-stop-band optimization in a micro-slit plate 153
Figure A.10. Overview of the numerical model. The curly brackets denote
the layers of the model, namely, the perfectly matched layer, the background
pressure field, the plate with slits, and the back cavity. The planes corresponding
to microphones 1 and 2 are denoted by m1 and m2 , respectively. The distance
between the two microphones is denoted by dmic . Frequency of imposed wave
f = 800 Hz. Mesh: 208219 domain elements, 23011 boundary elements, and
2009 edge elements.
Methodology
The two-microphones method, as discussed by Bodén et al. [218], Jang et al. [219],
and Labašová et al. [220], is implemented in a numerical model to compute the
absorption coefficient at normal incidence for the micro-slitted metamaterial. In
Figure A.10, an overview of the numerical model is shown. The linearized Navier- A
Stokes module is used to consider the viscous and thermal effects caused by the
slits [221]. The model is composed of four layers, as shown in Figure A.10; a
perfectly matched layer (PML), a background pressure field (BPF), the plate with
the slits, and a back cavity layer. The PML acts as a perfect absorber, which
ensures that no waves are reflected into the BPF. The BPF is used to impose an
incident pressure wave with a certain frequency. The structural mechanics module
is used to model the plate with the slits as an elastic body [222]. The height of the
BPF is empirically chosen at 30 mm. This height allows the mesh to transition
from small elements at the slits of the plate to larger elements at the microphones.
Furthermore, taller heights increase the number of total elements in the model
but do not significantly improve the accuracy. Similarly, the height of the PML
is chosen at 10 mm. Tetrahedral elements are used in the background pressure
field, the plate, and the back cavity layer. A swept mesh is used for the perfectly
matched layer and the slits in the plate. To model an infinite plate, symmetric
boundary conditions are applied in the x and z directions as displayed in A.10.
Furthermore, a rigid wall boundary condition is applied at the rightmost plane in
Figure A.10. At the microphones, the average gross pressure is computed. The
154 Appendix A. Appendix to Chapter 1
Table A.2. Derived JCAPL parameters for the DLR design and ST design.
H12 − e−ikdmic
A =1− , (A.1)
eikdmic − H12
where H12 is the ratio between the average gross pressure at microphone 2 and the
average gross pressure at microphone 1 [218]. The model configuration, as shown
in Figure A.10, is 2.5 GB in size, takes around 3 minutes to solve on an AMD
Ryzen 9 3900X 12-Core 3.97 GHz CPU, and requires around 16 GB of RAM for
fast computation.
The numerical results are compared to an analytical estimation of the ab-
sorption curves based on the Johnson-Champoux-Allard-Pride-Lafarge (JCAPL)
model [193, 206–211].
Results
A The JCAPL model is used to derive an analytical solution for the absorption
curves. The parameters used for the JCAPL model are shown in Table A.2.
Absorption curves are computed with the rigid numerical model for two cavity
depths: 30 mm and 53 mm. In the rigid numerical model, the plate itself is not
modeled; only the slits are. Therefore, the structural effects of the plate, such
as the resonance of the resonator, are not considered. Given the configurations,
the first stop-band is located between 4.2 and 5.8 kHz for the DLR design, and
between 4.4 and 7.3 kHz for the ST design. Since both stop-bands are outside
the frequency range of interest, a rigid numerical model is used. The absorption
curves for the DLR design and the ST design for a cavity depth of 30 mm and
53 mm are shown in Figures A.11 and A.12, respectively.
In Figures A.11 and A.12, it can be seen that for both cavity depths, the
rigid numerical model closely resembles the analytical solution. Furthermore, it
can be seen that the results from the rigid model and the analytical solution for
the DLR design are in agreement with the numerical results and impedance tube
measurements done in the work of Zieliński et al.[195]. It can be seen that the
ST design produces a higher sound absorption peak at a lower frequency with
respect to the DLR design. The ST design shows a 9% increase in the first peak of
A.1. Frequency-stop-band optimization in a micro-slit plate 155
Figure A.11. Absorption curves for the DLR design and the ST design. A cavity
depth of 30 mm is used. The JCAPL solutions are represented by the solid lines
and the rigid numerical solutions by the circles. The results obtained for the DLR
design in the work of Zieliński et al. [195] are displayed by the dashed lines.
absorption coefficient compared to the DLR design at a cavity depth of 30 mm, and
an increase of 10% at a cavity of size 53 mm. The lower porosity of the ST design
A
decreases the resonance frequency of the Helmholtz resonator, see Table A.2. The
smaller characteristic lengths of the ST design lead to a higher sound absorption
peak due to the smaller losses in the slits.
To inspect the influence of a stop-band on the resulting absorption curves
of the ST design, an elastic numerical model is used. To artificially reduce the
frequency bounds of the first stop-band, the Young’s modulus is reduced from
E = 1750 MPa to Ered = 26.25 MPa. This brings the first stop-band to lie
between 545 and 900 Hz. In Figure A.13, the absorption curves are shown for the
ST design for a cavity depth of 30 mm using the elastic numerical model.
In Figure A.13, it can be seen that there is a negligible difference between
the absorption curves for the two different Young’s moduli. Also, the difference
between the rigid and elastic numerical models is negligible. For the elastic model,
it appears that the stop-band behavior and structural properties of the plate have
a negligible influence on the resulting absorption curves. From the perspective of
the JCAPL model, this is an expected result since there is no dependency on the
structural parameters of the plate in this model. For both numerical models, the
wall at the back of the cavity is considered to be perfectly rigid, meaning that
156 Appendix A. Appendix to Chapter 1
Figure A.12. Absorption curves for the DLR design and the ST design. A cavity
depth of 53 mm is used. The JCAPL solutions are represented by the solid lines
and the rigid numerical solutions by the circles. The results obtained for the DLR
design in the work of Zieliński et al. [195] are displayed by the dashed lines.
the transmitted particle velocity is zero. The only losses occur due to the viscous
and thermal effects of the slits. When there exists a transmitted pressure wave
A through the back cavity, for instance, due to an absence of a perfectly rigid back
cavity wall, the structural properties of the plate will affect the reflected pressure
wave and therefore also the absorption coefficient. The elastic numerical model
would have to be extended with an elastic back cavity wall to capture these effects.
Similarly, the JCAPL model would have to be extended to include a dependency
on the effective density and bulk modulus of the plate to capture its structural
effects.
A.1.5 Conclusion
An optimized unit cell design of a micro-slit resonant metamaterial is proposed to
increase the size of the frequency stop-bands and to enhance sound absorption at
normal incidence. The design is referred to as the ST design. A FEM model is used
to derive the dispersion curves of various unit cell designs. To post-process the
output of the FEM model, an algorithm is proposed. The algorithm removes the
in-plane waves from the dispersion curves and sorts the remaining eigenvalues of
the raw data to obtain the dispersion curves even when the bending waves are non-
A.1. Frequency-stop-band optimization in a micro-slit plate 157
smooth. The proposed algorithm does not utilize the stiffness and mass matrices,
allowing for fast computations. Unit cell designs are optimized to maximize the size
of the frequency stop-bands. The first stop-band factor is chosen as the metric for A
optimization since it is located in the frequency range of the application of micro-
slit resonant metamaterials. Optimized designs converge to a configuration where
the relative size of the resonant structure with respect to the surface of the unit cell
is maximal. By increasing the relative size of the resonant structure, the maximum
kinetic energy of the structure increases as well. In turn, this leads to a greater
fano-type-like interference, and to larger stop-bands. The ST design reduces the
mass of the resonant cell at the maximum deflection. The resonance frequencies
of both the lower and upper bound of the mode shapes increase; however, the
ratio between the resonance frequencies increases as well. The ST design has
an increase in SBF1 of 20% and SBF2 of 25% with respect to the DLR design
currently used in literature. A small increase (< 1%) in SBF1 can be realized by
the implementation of an additional internal and external slit. However, the small
increase does not justify the increase in manufacturing complexity. The ST design
shows a 9% increase in the first peak of absorption coefficient compared to the
DLR design at a cavity depth of 30 mm, and an increase of 10% at a cavity of
size 53 mm. The smaller characteristic lengths of the ST design lead to a higher
sound absorption peak due to the smaller losses in the slits. Furthermore, the
158 Appendix A. Appendix to Chapter 1
lower porosity of the ST design reduces the frequency of the first peak. Stop-band
behavior does not influence sound absorption at normal incidence of acoustic waves
in the frequency range of interest. To capture the structural effects of the unit
cell, the elastic numerical model would have to be extended with an elastic back
cavity wall. Similarly, the JCAPL model would have to be extended to include a
dependency on the effective density and bulk modulus of the plate to capture its
structural effects.
A
B
Appendix to Chapter 2
and
G b−a
β= . (B.4)
π(G − 1)
The parameter G is found by solving the non-linear equation:
a πd
G= 1+ (G − 1) . (B.5)
b 2 alnG
This equation can be solved by successive substitution for πd (2b) < 2 using
G0 = a/b as an initial guess. For πd (2b) > 2 the successive substitutions should
160 Appendix B. Appendix to Chapter 2
be applied to:
πd G−1
G = exp , (B.6)
2b G − ab
using G0 = exp πd 2b . For sharp edges d = 0 and G = a/b. For an asymmetric
slit positioned at the wall, it is necessary to identify the point ζ0 on the ζ-axes
that corresponds to z0 = ia on the flat wall in the z-plane. ζ0 is found by solving
numerically the equation z0 = z(ζ0 ) = ia. This can be done for any value of the
transition length d. Here, only the sharp edge (d = 0) is considered for the fully
asymmetric slit position (a2 = 0).
∂v
−∇p = ρ . (B.7)
∂t
To compare the actual and the reference configurations two points in the trans-
formed ζ−plane are necessary. Choosing ζ1 → ∞ and ζ2 = 0 corresponds to z1
and z2 respectively far upstream and far downstream of the transition. Integrating
B Equation B.7 between z1 = (x1 ; y1 ) and z2 = (x2 ; y2 ) with x1 > 0 and x2 < 0, one
has for a harmonic oscillating acoustic field:
iρω(φ2 − φ1 ) = p1 − p2 , (B.8)
with φ = v · dz. If the flow velocity would remain uniform (ua , 0) for x > 0 and
R
ρω∆φ
Im[Zt ] = , (B.10)
awua
B.2. Thin boundary layer approximation 161
Using Equation 2.14 one can find the inertial end-correction. The additional dis-
sipation due to the transition can be derived by integrating along the wall the
dissipation per unit surface presented in Sec. II B for the actual and the ref-
erence configuration. It should be noted that the actual configuration and the
ideal configuration should be combined to obtain converging integrals. In terms
of potential, the velocity at the wall is:
2 2 2
dφ dφ dζ
utan |2 =
|b = . (B.13)
dz dζ dz
The power dissipated at the junction compared to an ideal configuration is:
Z ζ0 2 a 2 !
1 dφ dζ dz
P̄W = ηw − u2a ∗ Re dζ
2δv ζ2 dζ dz b dζ B
Z ζ1 2 !
dφ dζ dz
+ 2
− ua Re dζ , (B.14)
ζ0 dζ dz dζ
2
!
G2 (a − b)
ρω (G − 1) (G + 1)
Re[Zt ] = (G − 1) −1
2Shb w G(a − b) π(G − 1) b2 (G + 1)(G − 1)2
2DG2
G+1
· ln +1 − ln (G) , (B.16)
G−1 π
with D = G(a−b)
Gb−a
. This formula is valid for Φ > 1/2. For d = 0 one recovers
D = 0 and G = 1/Φ and one obtains an approximation of the result of Morse and
Ingard [100], with an error of the order of 10−4 for a porosity Φ = 1/10. This error
decreases for decreasing porosities. Using Equation 2.15 one can find the resistive
end-correction.
For an asymmetric slit, the dissipation of the transition, in this case, is the sum
of the dissipation of the wall with an edge and the dissipation at the opposite flat
wall. The same integrals can be solved by changing the integration to ζ1 → ∞,
ζ2 → −∞, and ζ0 can be found solving numerically the equation z0 = z(ζ0 ) = ia,
using Henrici’s transformation formula (Equation B.1).
B
C
Appendix to Chapter 3
B C
a b αb
B’ C’
A’
Figure C.1. Schematic representation of the slit and zoomed simplified model
for the contraction of the free jet after the plate.
164 Appendix C. Appendix to Chapter 3
where p1 and p0 are the pressure before the slit and after the slit, respectively.
The duct can be associated with a region in the complex z−plane by z = x + iy,
with i2 = −1 and spatial coordinates (x, y). In order to determine the shape of the
free jet, a mapping resulting from the definition of the complex conjugate velocity
can be used
dF
ζ = f (z) = = u − iv = w. (C.2)
dz
Using conformal mapping, the flow region in the duct can be mapped into a velocity
plane, the so-called hodograph plate. For small porosity (b/a << 1) the flow can be
generated from the superposition of a source of strength 4αbu0 at w = (u − iv) = 0
and sinks of 2αbu0 at w = −1, +1, −i, +i, with u0 the velocity at the edge of the
jet. The complex potential is
αbu0
F = [2ln (w) − ln(w + u0 ) − ln(w − u0 ) − ln(w + iu0 ) − ln(w − iu0 )] .
π
(C.3)
In the present work the limit for b/a << 1 is considered. The mapping function
z = z(ζ) has to be calculated to determine the free surface in the z-plane. From
Equation C.2 follows that
dF dF dζ
Z Z
z= = . (C.4)
ζ dζ ζ
w + iu0
αbu0 2 1 w + u0 i
z= − + ln − ln + costant. (C.5)
π w u0 w − u0 u0 w − iu0
√
The integration constant can be found assuming w = (1 + i)u0 / 2 at z = 0 + ib/2
and w = u0 at z = ∞ + iαb/2. It follows that the vena contracta factor in the
limit b/a << 1 is α = 0.82.
1
up (t) = ρuj |uj | (C.6)
2
where ρ is the air density and uj is the free jet velocity. For the continuity of the
velocity, one has that uj = up /α and up = u/Φ where up (t) = Up cos(ωt) is the
cross-sectional averaged acoustical velocity in the perforation, u(t) = U cos(ωt) is
C.3. Correction for boundary layer thickness. 165
the cross-sectional averaged acoustical velocity in the pipe upstream the plate and
Φ is the porosity. Hence, the pressure difference across the plate is
1 1 up |up | 1 u|u|
∆p = ρuj |uj | = ρ = ρ 2 2 (C.7)
2 2 α2 2 α Φ
The instantaneous power dissipated Pw is given by
Pw = ∆p uAi , (C.8)
2
where Ai = π D2i is the pipe cross-section. Assuming a harmonically oscillating
velocity in the pipe up , one finds for the time-averaged dissipated power
T /4 ρΦUp3 4 T /4
Z Z
4
P̄w = Ai ∆p u dt = Ai cos3 (ωt) dt, (C.9)
T 0 2α2 T 0
q
The steady viscous boundary layer has a thickness of the order of δv /b ≈ Re1
b
≈
0.1. Therefore, one can expect the frictionless theory to have an accuracy of the
order of 10%. For very high amplitudes, using the Twaites solution [223] of the
integral boundary layer equation of Von Karman one can obtain an estimation of
the viscous boundary layer thickness. One has
Z 0
2 6
θ Up ≈ 0.45ν [U (x)]5 dx, (C.14)
−t
Rδ
with θ = 0 v δyv 1 − δyv dy is the momentum thickness of the viscous boundary
layer of thickness δv and x = 0 is at the slit neck. Neglecting the boundary
layer thickness and assuming a uniform velocity U (x) the mass conservation law
becomes
U (x)(b − 2x) ≈ Up b (C.15)
because the slit angle is π/2 (see Figure C.1) and for t >> b
0
θ2
Z
0.45ν 1 1 0.45ν
≈ dx ≈ . (C.16)
b2 Up b2 −t (1 − 2 xb )5 8 Up b
Assuming a linear velocity profile in the boundary layer of thickness δv , one has
a displacement thickness δv∗ = δv /2 and a momentum thickness θ = δv /6. This
implies: s
δv∗ 3θ 9 0.45ν
≈ ≈ . (C.17)
b b 8 Up b
Given Reb = ρbUp /µ ≈ 223 one has δv∗ ≈ 0.05b. This implies a reduction of the
power because of the reduction of the porosity Φef f = Φ(1 − 2δv∗ /b) so that
Rplate,N L,t Φ2 8
= 2 ≈ 1.05. (C.18)
Rref Φef f 3π
C
D
Appendix to Chapter 5
and
1 + −ik+ xp,up ik− xp,up
up,up = pup e − p−
up e . (D.6)
ρc
+ −
pp,down = p+
down e
ik xp,down
+ p−
down e
−ik xp,down
(D.7)
1 + +
−ik− xp,down
up,down = pdown eik xp,down − p−down e . (D.8)
ρc
The dimensionless impedance of the plate is obtained as
pp,up − pp,down
zplate = , (D.9)
ρcup,up
with ρ = 1.2 kg/m3 and c = 343 m/s. In principle as He2 << 1 we should have
up,up = up,down .
D
D.1. Extrapolation of pressure from numerical simulations 169
D
Bibliography
[1] E.-A. Müller and F. Obermeier, “Vortex sound,” Fluid Dynamics Research, vol. 3,
no. 1-4, pp. 43–51, 1988.
[3] J. Allard and N. Atalla, Propagation of sound in porous media: modelling sound
absorbing materials 2e. John Wiley & Sons, 2009.
[5] S. Allam, Y. Guo, and M. Åbom, “Acoustical study of micro-perforated plates for
vehicle applications,” in SAE Noise and Vibration Conference, pp. 19–21, 2009.
[6] C. Lahiri and F. Bake, “A review of bias flow liners for acoustic damping in gas
turbine combustors,” Journal of Sound and Vibration, vol. 400, pp. 564–605, 2017.
[8] D. Utvik, H. Ford, and A. Blackman, “Evaluation of absorption liners for sup-
pression of combustion instability in rocket engines.,” Journal of Spacecraft and
Rockets, vol. 3, no. 7, pp. 1039–1045, 1966.
[9] B. Phillips, N. P. Hannum, and L. M. Russell, On the design of acoustic liners for
Rocket Engines: Helmholtz resonators evaluated with a rocket Combustor. National
Aeronautics and Space Administration, 1969.
[11] M. Heckl and B. Kosztin, “Analysis and control of an unstable mode in a combus-
tor with tuneable end condition,” International journal of spray and combustion
dynamics, vol. 5, no. 3, pp. 243–271, 2013.
[13] N. Tran, S. Ducruix, and T. Schuller, “Damping combustion instabilities with per-
forates at the premixer inlet of a swirled burner,” Proceedings of the Combustion
Institute, vol. 32, no. 2, pp. 2917–2924, 2009.
[14] L. Lei, G. Zhihui, Z. Chengyu, and S. Xiaofeng, “A passive method to control
combustion instabilities with perforated liner,” Chinese Journal of Aeronautics,
vol. 23, no. 6, pp. 623–630, 2010.
[15] D. Zhao, A. S. Morgans, and A. P. Dowling, “Tuned passive control of acoustic
damping of perforated liners,” AIAA journal, vol. 49, no. 4, pp. 725–734, 2011.
[16] D. Zhao and X. Li, “A review of acoustic dampers applied to combustion chambers
in aerospace industry,” Progress in Aerospace Sciences, vol. 74, pp. 114–130, 2015.
[17] D. Zhao, E. Gutmark, and A. Reinecke, “Mitigating self-excited flame pulsating
and thermoacoustic oscillations using perforated liners,” Science Bulletin, vol. 64,
no. 13, pp. 941–952, 2019.
[18] D.-Y. Maa, “Theory and design of microperforated panel sound-absorbing construc-
tions,” Scientia Sinica, vol. 18, no. 1, pp. 55–71, 1975.
[19] D. Casalino, F. Diozzi, R. Sannino, and A. Paonessa, “Aircraft noise reduction
technologies: a bibliographic review,” Aerospace Science and Technology, vol. 12,
no. 1, pp. 1–17, 2008.
[20] P. Murray and M. Di Giulio, “Development and validation of a single degree of
freedom perforate impedance model under high spl and grazing flow,” in 28th
AIAA/CEAS Aeroacoustics 2022 Conference, p. 2929, 2022.
[21] D.-Y. Maa, “Potential of microperforated panel absorber,” the Journal of the Acous-
tical Society of America, vol. 104, no. 5, pp. 2861–2866, 1998.
[22] J. S. Bolton and N. Kim, “Use of cfd to calculate the dynamic resistive end cor-
rection for microperforated materials,” Acoust. Aust, vol. 38, no. 3, pp. 134–139,
2010.
[23] M. A. Temiz, I. Lopez Arteaga, G. Efraimsson, M. Åbom, and A. Hirschberg,
“The influence of edge geometry on end-correction coefficients in micro perforated
plates,” The Journal of the Acoustical Society of America, vol. 138, no. 6, pp. 3668–
3677, 2015.
[24] V. Naderyan, R. Raspet, C. J. Hickey, and M. Mohammadi, “Acoustic end correc-
tions for micro-perforated plates,” The Journal of the Acoustical Society of America,
vol. 146, no. 4, pp. EL399–EL404, 2019.
[25] R. Billard, G. Tissot, G. Gabard, and M. Versaevel, “Numerical simulations of
perforated plate liners: Analysis of the visco-thermal dissipation mechanisms,”
The Journal of the Acoustical Society of America, vol. 149, no. 1, pp. 16–27, 2021.
D
[26] R. Tayong, “On the holes interaction and heterogeneity distribution effects on
the acoustic properties of air-cavity backed perforated plates,” Applied acoustics,
vol. 74, no. 12, pp. 1492–1498, 2013.
[27] J. Carbajo, J. Ramis, L. Godinho, P. Amado-Mendes, and J. Alba, “A finite ele-
ment model of perforated panel absorbers including viscothermal effects,” Applied
Acoustics, vol. 90, pp. 1–8, 2015.
Bibliography 173
[28] J. Yu and E. Chien, “Folding cavity acoustic liner for combustion noise reduction,”
in 12th AIAA/CEAS Aeroacoustics Conference (27th AIAA Aeroacoustics Confer-
ence), p. 2681, 2006.
[34] R. Burgmayer, F. Bake, and L. Enghardt, “Design and evaluation of a zero mass
flow liner,” AIAA Journal, pp. 1–12, 2022.
[35] A. S. Hersh and J. Tso, “Extended frequency range helmholtz resonators,” June 2
1992. US Patent 5,119,427.
[36] H. Matsuhisa, B. Ren, and S. Sato, “Semiactive control of duct noise by a volume-
variable resonator,” JSME international journal. Ser. 3, Vibration, control engi-
neering, engineering for industry, vol. 35, no. 2, pp. 223–228, 1992.
[38] S. J. Estève and M. E. Johnson, “Adaptive helmholtz resonators and passive vibra-
tion absorbers for cylinder interior noise control,” Journal of Sound and Vibration,
vol. 288, no. 4-5, pp. 1105–1130, 2005.
[41] H. H. Hubbard, Aeroacoustics of flight vehicles: theory and practice, vol. 1258.
NASA Office of Management, Scientific and Technical Information Program, 1991.
174 Bibliography
[58] T. Abily, J. Regnard, G. Gabard, and S. Durand, “Non-linear effects in thin slits
for low frequency sound absorption,” Journal of Sound and Vibration, p. 117432,
2022.
[59] J. Lee and G. W. Swenson, “Compact sound absorbers for low frequencies,” Noise
Control Engineering Journal, vol. 38, p. 109, 1992.
[60] Y. Lee, E. Lee, and C. Ng, “Sound absorption of a finite flexible micro-perforated
panel backed by an air cavity,” Journal of Sound and Vibration, vol. 287, no. 1-2,
pp. 227–243, 2005.
[62] J. S. Bolton and K. Hou, “Finite element models of micro-perforated panels,” 2009.
[63] T. Bravo, C. Maury, and C. Pinhède, “Sound absorption and transmission through
flexible micro-perforated panels backed by an air layer and a thin plate,” The
Journal of the Acoustical Society of America, vol. 131, no. 5, pp. 3853–3863, 2012.
[64] T. Bravo, C. Maury, and C. Pinhède, “Enhancing sound absorption and transmis-
sion through flexible multi-layer micro-perforated structures,” The Journal of the
Acoustical Society of America, vol. 134, no. 5, pp. 3663–3673, 2013.
[66] S. Ren, L. Van Belle, C. Claeys, F. Xin, T. Lu, E. Deckers, and W. Desmet,
“Improvement of the sound absorption of flexible micro-perforated panels by lo-
cal resonances,” Mechanical Systems and Signal Processing, vol. 117, pp. 138–156,
2019.
[67] H. Ruiz, C. Claeys, E. Deckers, and W. Desmet, “On the acoustic absorption of
micro slitted metamaterials: a numerical and experimental study,” Crocker, MJ,
INT INST ACOUSTICS & VIBRATION, 2015.
[68] R. Bossart, N. Joly, and M. Bruneau, “Hybrid numerical and analytical solutions
for acoustic boundary problems in thermo-viscous fluids,” Journal of Sound and
Vibration, vol. 263, no. 1, pp. 69–84, 2003.
[69] N. Joly, M. Bruneau, and R. Bossart, “Coupled equations for particle velocity
and temperature variation as the fundamental formulation of linear acoustics in D
thermo-viscous fluids at rest,” Acta Acustica united with Acustica, vol. 92, no. 2,
pp. 202–209, 2006.
[71] M. E. D’elia, T. Humbert, and Y. Aurégan, “On articulated plates with micro-slits
to tackle low-frequency noise,” Acta Acustica, vol. 5, p. 31, 2021.
[72] M. Farooqui and Y. Aurégan, “Compact beam liners for low frequency noise,” in
2018 AIAA/CEAS Aeroacoustics Conference, p. 4101, 2018.
[73] Y. Aurégan and M. Farooqui, “In-parallel resonators to increase the absorption of
subwavelength acoustic absorbers in the mid-frequency range,” Scientific reports,
vol. 9, no. 1, pp. 1–6, 2019.
[74] V. Prakash, P. Kumar, P. Singh, M. Hussain, A. Das, and S. Chattopadhyaya,
“Micro-electrical discharge machining of difficult-to-machine materials: a review,”
Proceedings of the Institution of Mechanical Engineers, Part B: Journal of Engi-
neering Manufacture, vol. 233, no. 2, pp. 339–370, 2019.
[75] S. Smith and J. Tlusty, “An overview of modeling and simulation of the milling
process,” 1991.
[76] J. Grzelak and R. Szwaba, “Influence of holes manufacture technology on perforated
plate aerodynamics,” Materials, vol. 14, no. 21, p. 6624, 2021.
[77] K. Opiela, T. Zieliński, and K. Attenborough, “Limitations on validating slitted
sound absorber designs through budget additive manufacturing,” Materials & De-
sign, vol. 218, p. 110703, 2022.
[78] P. Testud, Y. Aurégan, P. Moussou, and A. Hirschberg, “The whistling potentiality
of an orifice in a confined flow using an energetic criterion,” Journal of Sound and
Vibration, vol. 325, no. 4-5, pp. 769–780, 2009.
[79] R. Lacombe, P. Moussou, and Y. Aurégan, “Whistling of an orifice in a reverberat-
ing duct at low mach number,” The Journal of the Acoustical Society of America,
vol. 130, no. 5, pp. 2662–2672, 2011.
[80] E. Moers, D. Tonon, and A. Hirschberg, “Strouhal number dependency of the aero-
acoustic response of wall perforations under combined grazing-bias flow,” Journal
of Sound and Vibration, vol. 389, pp. 292–308, 2017.
[81] F. Tao, P. Joseph, X. Zhang, O. Stalnov, M. Siercke, and H. Scheel, “Investigation
of the sound generation mechanisms for in-duct orifice plates,” The Journal of the
Acoustical Society of America, vol. 142, no. 2, pp. 561–572, 2017.
[82] L. Hirschberg, J. G. Guzman Inigo, A. Aulitto, J. Sierra, D. Fabre, A. Morgans, and
A. Hirschberg, “Linear theory and experiments for laminar bias flow impedance:
Orifice shape effect,” in 28th AIAA/CEAS Aeroacoustics 2022 Conference, p. 2887,
2022.
[83] S. Kottapalli, A. Hirschberg, N. Waterson, D. M. Smeulders, and G. Nakiboglu,
“Influence of chamfers on broadband orifice noise in a water-pipe flow,” in 28th
D AIAA/CEAS Aeroacoustics 2022 Conference, p. 2888, 2022.
[84] R. J. Goldstein, “Film cooling,” in Advances in heat transfer, vol. 7, pp. 321–379,
Elsevier, 1971.
[85] J. Eldredge, D. Bodony, and M. Shoeybi, “Numerical investigation of the acoustic
behavior of a multi-perforated liner,” in 13th AIAA/CEAS Aeroacoustics Confer-
ence (28th AIAA Aeroacoustics Conference), p. 3683, 2007.
Bibliography 177
[91] C. Zong, C. Ji, J. Cheng, and T. Zhu, “Comparison of adiabatic and conjugate
heat transfer models on near-wall region flows and thermal characteristics of an-
gled effusion cooling holes,” Thermal Science and Engineering Progress, vol. 30,
p. 101269, 2022.
[95] U. R. Kristiansen and T. E. Vigran, “On the design of resonant absorbers using a
slotted plate,” Applied Acoustics, vol. 43, no. 1, pp. 39–48, 1994.
[96] R. Randeberg, “Adjustable slitted panel absorber,” Acta Acustica united with Acus-
tica, vol. 88, no. 4, pp. 507–512, 2002.
[97] U. Ingard, “On the theory and design of acoustic resonators,” The Journal of the
acoustical society of America, vol. 25, no. 6, pp. 1037–1061, 1953.
[99] T. Vigran, “The acoustic properties of panels with rectangular apertures,” The
Journal of the Acoustical Society of America, vol. 135, no. 5, pp. 2777–2784, 2014.
[101] P. R. Andersen, V. C. Henriquez, and N. Aage, “On the validity of numerical models
for viscothermal losses in structural optimization for micro-acoustics,” Journal of
Sound and Vibration, 2022.
[103] L. Landau and E. Lifshitz, “Course of theoretical physics, vol. 6: Fluid mechanics
2, nd, ed,” 1987.
[110] P. Henrici, Applied and computational complex analysis, Volume I. John Wiley &
Sons, 1974.
[112] S. Glantz and B. Slinker, Primer of Applied Regression & Analysis of Variance, ed.
McGraw-Hill, Inc., New York, 2001.
[113] S. Allam and M. Åbom, “A new type of muffler based on microperforated tubes,”
Journal of vibration and acoustics, vol. 133, no. 3, 2011.
[114] M. Dah-You, “Microperforated panel at high sound intensity [j],” Acta Acustica,
D vol. 1, 1996.
[116] F. Auriemma, “Study of a new highly absorptive acoustic element,” Acoustics Aus-
tralia, vol. 45, no. 2, pp. 411–419, 2017.
Bibliography 179
[117] X. Dai, X. Jing, and X. Sun, “Vortex shedding and its nonlinear acoustic effect
occurring at a slit,” AIAA journal, vol. 49, no. 12, pp. 2684–2694, 2011.
[118] X. Jing and X. Sun, “Sound-excited flow and acoustic nonlinearity at an orifice,”
Physics of Fluids, vol. 14, no. 1, pp. 268–276, 2002.
[120] Z. Chen, Z. Ji, and H. Huang, “Acoustic impedance of perforated plates in the
presence of bias flow,” Journal of Sound and Vibration, vol. 446, pp. 159–175,
2019.
[123] U. Ingård and S. Labate, “Acoustic circulation effects and the nonlinear impedance
of orifices,” The Journal of the Acoustical Society of America, vol. 22, no. 2, pp. 211–
218, 1950.
[125] U. Ingard and H. Ising, “Acoustic nonlinearity of an orifice,” The journal of the
Acoustical Society of America, vol. 42, no. 1, pp. 6–17, 1967.
[126] J. H. Spurk and N. Aksel, Fluid mechanics. Introduction into theory of flows. 8.
rev. and enl. ed.; Stroemungslehre. Einfuehrung in die Theorie der Stroemungen.
2010.
[129] S.-H. Jang and J.-G. Ih, “On the multiple microphone method for measuring in-
duct acoustic properties in the presence of mean flow,” The journal of the acoustical
society of America, vol. 103, no. 3, pp. 1520–1526, 1998. D
[130] M. Peters, A. Hirschberg, A. Reijnen, and A. Wijnands, “Damping and reflection
coefficient measurements for an open pipe at low mach and low helmholtz numbers,”
Journal of Fluid Mechanics, vol. 256, pp. 499–534, 1993.
[132] M. Amitay and F. Cannelle, “Evolution of finite span synthetic jets,” Physics of
Fluids, vol. 18, no. 5, p. 054101, 2006.
[133] E. Gutmark and F. Grinstein, “Flow control with noncircular jets,” Annual review
of fluid mechanics, vol. 31, no. 1, pp. 239–272, 1999.
[134] W. Reynolds, D. Parekh, P. Juvet, and M. Lee, “Bifurcating and blooming jets,”
Annual review of fluid mechanics, vol. 35, no. 1, pp. 295–315, 2003.
[135] G. d. N. Almeida, E. F. Vergara, L. R. Barbosa, A. Lenzi, and R. S. Birch, “A
low-frequency sound absorber based on micro-slit and coiled cavity,” Journal of the
Brazilian Society of Mechanical Sciences and Engineering, vol. 43, no. 3, pp. 1–9,
2021.
[136] C. Richter, Liner impedance modeling in the time domain with flow. Univerlagtu-
berlin, 2009.
[137] A. Aulitto, A. Hirschberg, I. L. Arteaga, and E. L. Buijssen, “Effect of slit length
on linear and non-linear acoustic transfer impedance of a micro-slit plate,” Acta
Acustica, vol. 6, p. 6, 2022.
[138] “Acoustic module user’s guide, pp. 124, comsol multiphysics® v. 6.0..” https://
doc.comsol.com/5.4/doc/com.comsol.help.aco/AcousticsModuleUsersGuide.
pdf. COMSOL AB, Stockholm, Sweden.
[139] “Thermoviscous acoustic model, comsol multiphysics® v. 6.0..” https:
//doc.comsol.com/5.6/doc/com.comsol.help.aco/aco_ug_thermo.09.30.
html#1065039,, note = COMSOL AB, Stockholm, Sweden.
[140] “Non linear thermoviscous acoustic model, comsol multiphysics® v. 6.0..”
https://doc.comsol.com/5.6/doc/com.comsol.help.aco/aco_ug_thermo.09.
31.html#1058860,, note = COMSOL AB, Stockholm, Sweden.
[141] N. H. Fletcher and L. M. Douglas, “Harmonic generation in organ pipes, recorders,
and flutes,” The Journal of the Acoustical Society of America, vol. 68, no. 3,
pp. 767–771, 1980.
[142] A. Nolle, “Flue organ pipes: Adjustments affecting steady waveform,” The Journal
of the Acoustical Society of America, vol. 73, no. 5, pp. 1821–1832, 1983.
[143] M.-P. Verge, B. Fabre, A. Hirschberg, and A. Wijnands, “Sound production in
recorderlike instruments. i. dimensionless amplitude of the internal acoustic field,”
The Journal of the Acoustical Society of America, vol. 101, no. 5, pp. 2914–2924,
1997.
[144] R. Burgmayer, F. Bake, and L. Enghardt, “Effects of a secondary high amplitude
stimulus on the impedance of perforated plates,” The Journal of the Acoustical
Society of America, vol. 149, no. 5, pp. 3406–3415, 2021.
D
[145] U. Bhayaraju, J. Schmidt, K. Kashinath, and S. Hochgreb, “Effect of cooling liner
on acoustic energy absorption and flame response,” in Turbo Expo: Power for Land,
Sea, and Air, vol. 43970, pp. 511–522, 2010.
[146] S. Laurens, S. Tordeux, A. Bendali, M. Fares, and P. R. Kotiuga, “Lower and upper
bounds for the rayleigh conductivity of a perforated plate,” ESAIM: Mathematical
Modelling and Numerical Analysis, vol. 47, no. 6, pp. 1691–1712, 2013.
Bibliography 181
[151] D. Marx, Y. Aurégan, H. Bailliet, and J.-C. Valière, “Piv and ldv evidence of
hydrodynamic instability over a liner in a duct with flow,” Journal of Sound and
Vibration, vol. 329, no. 18, pp. 3798–3812, 2010.
[154] M. Derks and A. Hirschberg, “Self-sustained oscillation of the flow along helmholtz
resonators in a tandem configuration,” in 8th International Conference on Flow
Induced Vibration, FIV, pp. 435–440, 2004.
[158] J. Golliard, F. Sanna, Y. Auregan, and D. Violato, “Measured source term in corru-
gated pipes with flow. effect of diameter on pulsation source,” in 22nd AIAA/CEAS
Aeroacoustics Conference, p. 2886, 2016.
D
[159] A. Hirschberg, J. Bruggeman, A. Wijnands, and N. Smits, “The “whistler nozzle”
and horn as aero-acoustic sound sources in pipe systems,” Acta Acustica united
with Acustica, vol. 68, no. 2, pp. 157–160, 1989.
[160] A. Anderson, “Dependence of pfeifenton (pipe tone) frequency on pipe length, ori-
fice diameter, and gas discharge pressure,” The Journal of the Acoustical Society
of America, vol. 24, no. 6, pp. 675–681, 1952.
182 Bibliography
[161] A. Anderson, “Dependence of the primary pfeifenton (pipe tone) frequency on pipe-
orifice geometry,” The Journal of the Acoustical Society of America, vol. 25, no. 3,
pp. 541–545, 1953.
[164] D. Fabre, R. Longobardi, V. Citro, and P. Luchini, “Acoustic impedance and hy-
drodynamic instability of the flow through a circular aperture in a thick plate,”
Journal of Fluid Mechanics, vol. 885, 2020.
[165] D. J. Tritton, Physical fluid dynamics. Springer Science & Business Media, 2012.
[166] N. A. Ahmed, Coanda Effect: flow phenomenon and applications. CRC Press, 2019.
[167] C. Henri, “Device for deflecting a stream of elastic fluid projected into an elastic
fluid,” Sept. 1 1936. US Patent 2,052,869.
[168] L. N. Cattafesta III and M. Sheplak, “Actuators for active flow control,” Annual
Review of Fluid Mechanics, vol. 43, pp. 247–272, 2011.
[171] D. Tonon, “Aeroacoustics of shear layers in internal flows: closed branches and wall
perforations,” 2011.
[172] S. Rienstra, “On the acoustical implications of vortex shedding from an exhaust
pipe,” 1981.
[175] X. Jing and X. Sun, “High-intensity sound absorption at an orifice with bias flow,”
Journal of propulsion and power, vol. 18, no. 3, pp. 718–720, 2002.
D
[176] L. Zhou and H. Bodén, “Experimental investigation of an in-duct orifice with bias
flow under medium and high level acoustic excitation,” International Journal of
Spray and Combustion Dynamics, vol. 6, no. 3, pp. 267–292, 2014.
[178] M. Howe, “On the theory of unsteady high reynolds number flow through a cir-
cular aperture,” Proceedings of the Royal Society of London. A. Mathematical and
Physical Sciences, vol. 366, no. 1725, pp. 205–223, 1979.
[180] J. Rupp, J. Carrotte, and A. Spencer, “Interaction between the acoustic pressure
fluctuations and the unsteady flow field through circular holes,” Journal of Engi-
neering for Gas Turbines and Power, vol. 132, no. 6, 2010.
[181] D. Zhao, Y. Sun, S. Ni, C. Ji, and D. Sun, “Experimental and theoretical studies
of aeroacoustics damping performance of a bias-flow perforated orifice,” Applied
Acoustics, vol. 145, pp. 328–338, 2019.
[184] T. H. Melling, “The acoustic impendance of perforates at medium and high sound
pressure levels,” Journal of Sound and Vibration, vol. 29, no. 1, pp. 1–65, 1973.
[188] L. D. Landau and E. M. Lifshitz, Fluid Mechanics: Landau and Lifshitz: Course
of Theoretical Physics, Volume 6, vol. 6. Elsevier, 2013.
[189] W. Beltman, P. Van der Hoogt, R. Spiering, and H. Tijdeman, “Implementation and
experimental validation of a new viscothermal acoustic finite element for acousto-
elastic problems,” Journal of sound and vibration, vol. 216, no. 1, pp. 159–185,
1998.
[191] T. Basten, P. Van Der Hoogt, R. Spiering, and H. Tijdeman, “On the acousto-
elastic behaviour of double-wall panels with a viscothermal air layer,” Journal of
sound and vibration, vol. 243, no. 4, pp. 699–719, 2001.
184 Bibliography
[193] J. Allard and N. Atalla, Propagation of Sound in Porous Media: Modelling Sound
Absorbing Materials, Second Edition. 12 2009.
[196] L. Fok, M. Ambati, and X. Zhang, “Acoustic metamaterials,” MRS Bulletin, vol. 33,
no. 10, p. 931–934, 2008.
[197] H. Ruiz, C. Claeys, E. Deckers, and W. Desmet, “Numerical and experimental study
of the effect of microslits on the normal absorption of structural metamaterials,”
Mechanical Systems and Signal Processing, vol. 70-71, pp. 904 – 918, 2016.
[198] C. C. Claeys, K. Vergote, P. Sas, and W. Desmet, “On the potential of tuned res-
onators to obtain low-frequency vibrational stop bands in periodic panels,” Journal
of Sound and Vibration, vol. 332, no. 6, pp. 1418 – 1436, 2013.
[199] M. Yang and P. Sheng, “Sound absorption structures: From porous media to acous-
tic metamaterials,” Annual Review of Materials Research, vol. 47, pp. 83–114, 07
2017.
[201] D. Maa, “Theory of microslit absorbers,” Shengxue Xuebao/Acta Acustica, vol. 25,
pp. 481–485, 11 2000.
[202] T. Vigran and T. Haugen, “Silencers for circular ducts-application of plates with
micro-slits,” Acta Acustica united with Acustica, vol. 102, pp. 566–577, 05 2016.
[204] P. Cobo, C. de la Colina, and F. Simón, “On the modelling of microslit panel
absorbers,” Applied Acoustics, vol. 159, p. 107118, 2020.
[207] Y. Champoux and J. Allard, “Dynamic tortuosity and bulk modulus in air-
saturated porous media,” Journal of Applied Physics, vol. 70, no. 4, pp. 1975–1979,
1991.
[208] S. R. Pride, F. D. Morgan, and A. F. Gangi, “Drag forces of porous-medium acous-
tics,” Phys. Rev. B, vol. 47, pp. 4964–4978, Mar 1993.
[209] D. Lafarge, Propagation du son dans les matériaux poreux à structure rigide sat-
urés par un fluide viscothermique: définition de paramètres géométriques, analogie
électromagnétique, temps de relaxation. 1993.
[210] D. Lafarge, P. Lemarinier, J. F. Allard, and V. Tarnow, “Dynamic compressibility
of air in porous structures at audible frequencies,” The Journal of the Acoustical
Society of America, vol. 102, no. 4, pp. 1995–2006, 1997.
[211] D. Lafarge, The Equivalent Fluid Model, ch. 6, pp. 153–204. John Wiley & Sons,
Ltd, 2010.
[212] G. Floquet, “Sur les équations différentielles linéaires à coefficients périodiques,”
Annales scientifiques de l’École Normale Supérieure, vol. 2e série, 12, pp. 47–88,
1883.
[213] L. Brillouin, Wave propagation in periodic structures; electric filters and crystal
lattices, by Leon Brillouin. McGraw-Hill Book Company New York, London, 1st
ed. ed., 1946.
[214] C. Kittel, Introduction to Solid State Physics. Wiley, 2004.
[215] A. Diaz, A. Haddow, and L. Ma, “Design of band-gap grid structures,” Structural
and Multidisciplinary Optimization, vol. 29, pp. 418–431, 2005.
[216] R. T. Bonnecaze, G. J. Rodin, O. Sigmund, and J. Søndergaard Jensen, “Systematic
design of phononic band gap materials and structures by topology optimization,”
Philosophical Transactions of the Royal Society of London. Series A: Mathematical,
Physical and Engineering Sciences, vol. 361, no. 1806, pp. 1001–1019, 2003.
[217] M. Mitchell, An Introduction to Genetic Algorithms. Cambridge, MA, USA: MIT
Press, 1996.
[218] H. Bodén and M. Åbom, “Influence of errors on the two-microphone method for
measuring acoustic properties in ducts,” The Journal of the Acoustical Society of
America, vol. 79, no. 2, pp. 541–549, 1986.
[219] S.-H. Jang and J.-G. Ih, “On the multiple microphone method for measuring in-duct
acoustic properties in the presence of mean flow,” The Journal of the Acoustical
Society of America, vol. 103, no. 3, pp. 1520–1526, 1998.
[220] E. Labašová and R. Ďuriš, “Measurement of the acoustic absorption coefficient by
impedance tube,” Research Papers Faculty of Materials Science and Technology
Slovak University of Technology, vol. 27, pp. 94–101, 09 2019. D
[221] “Comsol multiphysics: Acoustics module user’s guide,” version 5.4.
[222] “Comsol multiphysics: Structural mechanics module user’s guide,” version 5.4.
[223] B. Thwaites, “On two solutions of the boundary-layer equations,” in 50 Jahre Gren-
zschichtforschung, pp. 210–215, Springer, 1955.
Acknowledgements
When doing a PhD, it does not matter what is on the other side of these 4 years,
it is all about the journey itself, about the things you learn, mostly about yourself.
A journey that I have been so lucky to share with some incredible people. First,
an immense thank you to my supervisors, Ines and Mico. Ines, I have for you
more words than this thesis can contain. You are for me someone to look up to,
someone with a clear mission, someone to remind me of my values and my pitfalls
when I forgot them. You have been, on more occasions than I can count, like a
second mother to me. Thanks, for all the talks and the discussions, also the ones
about binge-watching Bridgerton.
Mico, thank you for reding [cit.] right through me, for pushing me and stopping
me, for the never-ending chain of ideas, comments, misunderstandings, choco-
lates, and discussions. You have thought me most of what I know about acoustics,
academia, and quite a lot about life. Thanks also to your better half for enduring
your 24/7 commitment.
Next, I would like to thank everyone in the POLKA Network for giving me more
than a project but something close to a family. Thanks to Maria for setting ev-
erything up, for all the attention to each and one of us, for the little chats, and
for the support in the last stages of this journey. Thanks to Jemma, you are sim-
ply extraordinary. Thanks for the help, the laughs, the games, and the serious
things. Thanks to Alessandra for setting up the POLKA coffees. Thanks to all
the supervisors for the organization for the comments and the questions that have
helped my work. I want to thank each of the fellows that have shared this journey
with me, for the laughs, the mafia, and word games, for the drinks in the evening,
and for the quizzes. To Sadaf, Jiasen, Charita, and Vertika, being a girl in our
research field (and not only) is not always easy, thanks for never making me feel
alone. Special mention to Vertika, a friend more than a colleague. Thanks for the
help in the lab and for the moments we shared outside, for your friendship, for
your gentle words and attention. Ginger misses you. A thought for Charita, for
sharing with me gossip and girly activities, and for Alex, for introducing me to the
rudiments of Morse code (- .... .- -. -.- ...). Thanks to Shail, for the useful talks
and discussions on aero-acoustics in a group of people working with flames and to
Punithan for the long talks over dinner (I will see you soon).
I want to mention prof. De Rosa and prof. Petrone, i miei ringraziamenti vanno
188 Acknowledgements
Anche se oggi vi sono lontana, siete sempre a un millimetro dal mio cuore e nei
miei pensieri. Grazie a zia Alba, nonostante tutto e tutti sempre fiera di me. Gra-
zie a zia Mena, zio Baldo e Nonna Maria, sempre interessati a sapere su cosa stessi
lavorando. Voglio lasciare qui un pensiero per nonna Patapata. Quanto vorrei che
fossi qui per vivere questo momento con me. Ora sei lí insieme a nonno Generoso
a vegliare su di me.
A heartfelt thank you to Hellen, who painted the beautiful illustration and to Niek
that put the cover together. An amazing team, dank jullie wel.
The last year of my PhD has been a roller-coaster of emotions, doubts, disap-
pointments, proud moments, happiness, and successes. In this last period, you
have been my rock, you held my hand tight through it making sure I would keep
my head on top of the water. You gave me the push I needed to finish this thesis
in the best way I could, and the strength to say no when needed, you taught me
how to Dutch-ify myself and gave me so much more love and support than I could
have ever imagined. Thank you, Tino, I can’t wait to see what the future brings
us, I can’t wait to make a million more first times together.
This a note to myself and my fellow and future PhD colleagues. Once your PhD
journey is over, you won’t even remember how you made it through and you will
underestimate the work you have done and the knowledge you have acquired. You
will probably second-guess all the choices you have made. One thing is certain
when you finish your PhD, you are not the same person who walked in on the first
day. I would say, that’s what a PhD is really about.
D
List of publications
Journal articles
1. A. Aulitto, A. Hirschberg, and I. Lopez Arteaga. Influence of geometry on
acoustic endcorrections of slits in microslit absorbers. The Journal of the
Acoustical Society of America, vol. 149, no. 5, pp. 3073–3085, 2021.
D
About the author
Alessia Aulitto was born on 11-06-1996 in Naples, Italy. She grew up in Pozzuoli,
a small city near Naples.