Aero MSC Report
Aero MSC Report
Xueqing Zhang
Ad
Noise Prediction of a NACA 0015
Airfoil with Gurney Flap
A study based on Time-Resolved Particle Image Velocimetry
Xueqing Zhang
The undersigned hereby certify that they have read and recommend to the Faculty
of Aerospace Engineering for acceptance the thesis entitled “Noise Prediction of a
NACA 0015 Airfoil with Gurney Flap” by Xueqing Zhang in fulfillment of the
requirements for the degree of Master of Science.
Supervisors:
Dr. ir. Andrea Sciacchitano
This thesis is dedicated to my family, whose unconditional love is never absent to back
me up, and whose own sedulous pursuits of professional excellence always encourage me
to move on. In memory of my grandfather, who set the paragon of a qualified aerospace
engineer for me to follow. I am grateful to Teng Wang, for his help in digital signal
processing and for every moment we have spent together.
Many gratitudes to Dr.Andrea Sciacchitano, for offering me the opportunity to work on
aeroacoustics and for being an excellent mentor throughout the process. It has been an
honour to work under his supervision. Si me lo permite, quisiera darle mis más sinceras
felicitaciones y mis mejores deseos para esta nueva etapa en su vida, ¡un bébé!
Furthermore, I would like to thank Dr.Stefan Pröbsting, for the patient guidance
and discussions in aeroacoustics, as well as for sharing the perceptions he harbours
for this world that offered me a viewpoint I could otherwise ignore with my eastern
cultural background. I would also like to thank all the technicians in the aerodynamics
department for their friendly help and icecreams.
To all who shared this journey, thank you! Special mentions are due to Ilke and Zhenni,
without whom my journey at TU Delft would never have begun; to Alberto, Ananth,
Bo, Bowei, Jiale, Jiggar, Koen, Linfeng, Mudit, Remco, Shengling, Shruthi, Teng, Yifei,
Ziyuan and many others with whom I had the pleasure of crossing our paths; to Xin
Liao and Xing Meng, who are far away but always near.
Xueqing Zhang (Caddie)
张 雪晴
Delft, December 2015
People who design machines and airplanes, no matter how much they be-
lieve that what they do is good, the winds of time eventually turn them into tools
of industrial civilization. They’re cursed dreams. Beautiful yet cursed dreams.
— Hayao Miyazaki
“设 计 飞 机 和 机 械 的 人, 无 论 他 们 的 意 图 是 多 么 善 良, 时 代 之 风 会 把 它 转
化为机械文明的工具,从来都不是无害的,都是被诅咒的梦想。”
— 宫崎 骏
Air traffic noise, especially during the aircraft take-off and landing, has been uni-
versally acknowledged as a nuisance, for which reason, the International Civil Avia-
tion Organization(ICAO) has ratified articles and established specifications aiming
at the attenuation of air traffic noise, such as the Annex 16. The implementation
of large bypass ratio turbo-fan engines on civil aircrafts and the application of de-
signs such as engine nacelle and chevron nozzle have shifted the attention of noise
control to the deployment of high-lift devices during take-off and landing. Source
control is proved to be effective in the reduction of aeronautical noise. Traditional
aeroacoustic assessments rely mainly on microphone (array) measurements and
instantaneous flow fields obtained from CFD. However, the former fails in provid-
ing the mechanism of sound generation and the latter becomes prohibitive in the
operating Reynolds number range of real aircrafts.
Previous investigations have validated the causality between the unsteady trans-
verse velocity in the wake of the high-lift devices and the far-field pressure fluctua-
tion. This thesis is a continuation of the previous study, focusing on investigating
the feasibility of noise prediction of high-lift devices based on experimental data
obtained by time-resolved PIV . A model of NACA 0015 airfoil with Gurney flap
of the height of 6% the chord length has been investigated by means of 2D time-
resolved PIV in combination with simultaneous microphone measurements. The
Gurney flap is a simplified model of the high-lift devices with more complex struc-
ture in real application. The experimental parameters were selected such that the
source of sound can be treated as compact and consequentially guaranteed the
applicability of both the distributed and the lumped formulation of the Curle’s
aeroacoustic analogy.
In the first step, the time evolving pressure field was deduced from the velocity
field measured by 2D TR-PIV, by the Poisson-based solver for quasi-2D incom-
pressible flows. Pressure on the surfaces of the airfoil then constitutes the source
terms of the distributed formulation of Curle’s analogy and the corresponding far-
field noise was evaluated. In the second step, the aerodynamic loads of the airfoil
were evaluated from the velocity field by means of momentum balance, which were
then used in the lumped formulation for noise prediction. The spanwise coherence
length evaluated from the cross-plane flow visualizations was applied for the cor-
rection of 3D effect.
Both formulations of the Curle’s analogy yield noise evaluation of fair agreement
with the simultaneous microphone measurements. All the calculated far-field noise
power spectra reproduce the peak at Kármán vortex shedding frequency, which
agrees well with the microphone measurements. The pressure fluctuations yielded
by distributed formulation of the Curle’s analogy are somehow damped due to
the Gaussian smoothing applied to the velocity fields in the process of pressure
reconstruction for the elimination of spurious spatial derivatives. Such damping
effect can be circumvented by the lumped formulation, since the pressure is only
evaluated on control volume boundaries in the locating in the far-wake, where 3D
motions are weaker than in the region in direct proximity of the airfoil trailing
edge. The high frequency components in the computed acoustic spectra are the
most affected by the experimental and numerical errors. The coherence length
correction reduce the over-estimation of the predicted noise levels with respect to
the microphone measured ones. However, there still exists room for improvement
in the procedure of coherence length correction.
Concludingly, this thesis proved TR-PIV an effective approach in the noise pre-
diction of airfoil with Gurney flap and stepped out the first step in PIV-based
aeroacousitic investigations of the high-lift devices. Application of the approach
on high-lift devices with more complex configurations would be a prospective con-
tinuation of the current study.
基于时间解析粒子图像测速技术(PIV)的NACA0015翼
型Gurney襟翼噪声预测
摘 要
航空噪声,尤其是在飞机起飞着陆阶段的近场噪声,随着民用飞机数量的
剧增,日益引起世界范围的关注。 为此,在国际民用航空组织(ICAO)制定了
一系列飞机噪声适航条例,《国际民用航空公约》附件16便是其中之一。 由于大
涵道比涡扇发动机在民航客机中的广泛使用以及消声短舱、 V-型花瓣喷嘴等结
构上的改进大大地降低了发动机噪声,机体噪声飞机噪声中所占比例日益增大,
其中飞机起降阶段增升装置噪声是机体噪声的重要组成部分。 飞机降噪通常采
用声源控制的方法主动降噪,此方法需要根据噪声产生机理改进增升装置结构,
然而,传统的麦克风及阵列测量无法解释发声机理,而对于飞机设计雷诺数范
围的瞬时流场计算目前仍是数值模拟无法企及的。
前人研究中已经证实了增升装置尾流中横向速度波动与远场压力扰动的
因果关系。 本文基于这一结论,对基于时间分辨PIV速度测量数据进行增升
装置噪声预测的可行性进行了分析。 本文使用Gurney襟翼作为增升装置的简
化模型,对安装有高度为6%弦长的Gurney襟翼的NACA0015翼型同时进行二
维PIV和远场麦克风测试。 通过调整麦克风安装位置使得模型可以被看作致密
声源,从而远场噪声既可以通过对机翼表面偶极子源进行积分直接求解(下称:
分布式Curle声类比)
,也可以由速度场导出的气动载荷间接获得(下称:集总
式Curle声类比)
。
本文首先运用针对准二维不可压流的泊松方程由二维PIV测量的速度场导出
流场压力分布。进而由机翼表面压力波动通过分布式Curle声类比导出远场噪声。
除此之外,本文还通过动量方程解出机翼在每个瞬时的气动载荷,并由此运用
集总式Curle声类比求解远场噪声。 本文通过将两种形式的声类比导出的远场压
力扰动与麦克风测试结果对比,对两种情况的适用性进行了评估。 此外,通过
三维立体PIV在Gurney襟翼下游展向平面上的速度测量,可以导出展向相关长
度,从而对二维PIV的噪声预测结果进行修正。
在导出的压力场中可以观察到压力极小值点与涡量场中涡核位置相对应。
在实验过程中,可以清晰地听到音调噪音,且这一音调噪音所对应的频率峰值
可以在所有由速度场导出的噪声频谱中观察到,而这一峰值频率恰好对应卡门
涡街的漩涡脱落频率。 两种形式的Curle声类比预测的远场噪声都与相应位置的
麦克风测试有较好的吻合,所有由PIV预测出的噪声功率谱都再现了卡门旋涡脱
落频率位置的峰值。 为了避免流场测量误差对压力梯度带来的杂散,在求解压
力场之前首先需要对PIV测得的速度场进行高斯平滑处理,而这一过程使得近场
波动幅值衰减,从而导致分布式Curle声类比预测的远场压力扰动幅值下降。 相
比之下,由于导出气动载荷的控制体边界远离三维流动效应显著的区域,集总
式Curle声类比求解过程中避免了平滑处理的过程。 由本文中PIV导出的噪声频
谱观察到,噪声中高于音调噪音频率的高频组分受试验及数值的影响较其他组
分大。 相关长度修正减少了基于二维的噪声预测与麦克风测试之间的偏差,但
本文的校正过程仍有改进空间。
综上所述,本文验证了时间解析PIV在Gurney襟翼噪声预测上的可行性,迈
出了基于PIV的增升装置噪声研究的第一步,此方法在增升装置的噪声预测中将
具有可观的应用前景。
Preface v
Abstract vii
List of Figures xv
1 Introduction 1
1.1 Noise and noise control . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Aeronautical noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Research objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.1.1 Causality-correlation . . . . . . . . . . . . . . . . . . . . . . . 28
3.1.3 Coherence . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.2 Evaluation of pressure from velocity field . . . . . . . . . . . . . . . . 31
4 Experimental techniques 39
4.2.1 Seeding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.2.2 Illumination . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.2.3 Imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.2.4 Recording . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5 Experimental set-up 59
5.2 Experimental set-up for synchronized planar PIV and microphone mea-
surements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.3 Experimental set-up for stereoscopic PIV measurements . . . . . . . . 68
Bibliography 109
1.1 Drawing by Keely McGann, winning artist of the 2009 drawing contest
organized by ICAO on the theme: Aviation in a Green Environment. . . 2
2.3 Theoretical directivity patterns for far-field sound pressure levels radi-
ated from (a) monopole, (b) dipole, (c) lateral quadrupole, and (d)
longitudinal quadrupole sound sources. [59] . . . . . . . . . . . . . . . 22
2.4 Root mean square of the source terms retrieved from PIV measurements
of a 2D cavity by Koschatzky et al. (2010)[36] . . . . . . . . . . . . . 24
2.5 Power spectra of sound signal from a microphone and of sound com-
puted from PIV at the same locations with the Curle’s analogy and the
vortex sound theory by Koschatzky et al. (2010)[36] . . . . . . . . . . 24
2.6 Power spectra computed using Curle’s analogy with the microphone
measurement. Predictions based on full span coherence assumption and
on the measured spanwise coherence length by Lorenzoni et al.(2012)
[43] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.1 Control volume and Control surface around the cylinder [37] . . . . . . 34
3.2 Definition of Inviscid and Viscous region on the Control surface [37] . . 34
4.1 The technical chart, the dimensions and the equivalent schematic of a
Linear X-M53 microphone[31] . . . . . . . . . . . . . . . . . . . . . . 45
4.8 Timing diagram for PIV recording based on Double Frame/Single Ex-
posure frame straddling mode [44] . . . . . . . . . . . . . . . . . . . 55
5.3 NACA 0015 and Gurney flap models designed by J.Shah [63] and man-
ufactured at TU Delft . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.4 Photograph of the experimental set-up for synchronized planar PIV and
microphone measurements . . . . . . . . . . . . . . . . . . . . . . . . 63
5.5 Schematic of set-up for synchronised planar PIV and microphone mea-
surements(side view) . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.6 Schematic of set-up for synchronised planar PIV and microphone mea-
surements(top view) . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
6.7 An excerpt of time history of lift and the components of lift evaluated
from PIV measurement:GF6-AOA4-V20-FOV2 . . . . . . . . . . . . . 90
6.8 An excerpt of time history of drag and the components of drag evalu-
ated from PIV measurement:GF6-AOA4-V20-FOV2 . . . . . . . . . . 91
6.9 Power spectra (∆f = 2.4Hz) of lift and drag evaluated from PIV
measurement:GF6-AOA4-V20-FOV2 . . . . . . . . . . . . . . . . . . . 92
6.10 An excerpt of time history of far-field acoustic pressure evaluated by
lumped formulation of Curle’s analogy based on full span coherence
assumption from PIV measurement (GF6-AOA4-V20-FOV2) in com-
parison with the simultaneous microphone measurement at the same
location . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
6.11 Power spectra (∆f = 2.4Hz) of far-field acoustic pressure evaluated
by lumped formulation of Curle’s analogy based on full span coherence
assumption from PIV measurement (GF6-AOA4-V20-FOV2) in com-
parison with the simultaneous microphone measurement at the same
location on GF6 and GF0 model. . . . . . . . . . . . . . . . . . . . . 94
6.12 Stereoscopic PIV measurement: GF6-AOA4-V20, z = 15mm (First
row: mean velocity, Second row: velocity RMS, Third row: a snapshot
of instantaneous velocity, Left column: transverse u-component, Right
column: streamwise v-component) . . . . . . . . . . . . . . . . . . . 97
A.2 Extrapolated Gaussian curve and the predicted length of coherence . . 120
B.3 The first three resonance modes in an open cylindrical tube. The hori-
zontal axis is pressure.[73] . . . . . . . . . . . . . . . . . . . . . . . . 124
B.4 Power spectra (∆f = 2.4Hz) of the background noise of V-tunnel with
empty plexiglass-Kevlar test section and incoming flow at 20m/s. . . . 125
Introduction
Figure 1.1: Drawing by Keely McGann, winning artist of the 2009 drawing contest organized
by ICAO on the theme: Aviation in a Green Environment.
Noise is often referred to as unwanted sound, from the quiet but annoying sort such
as the barking of dog, to the loud and harmful ones, such as jet engine in close
quarter. The physics of the noise is the same as that of any other sound: in the case
of fluid (gas, plasma or liquid), the sound source creates vibration that propagates
longitudinally in the surrounding medium as sound waves. Sound waves are weak
perturbations that cause the local pressure in the medium to deviate from the
equilibrium one. This deviation is referred to as sound pressure, noted as p0 .
Loudness, commonly used to refer the auditory sensation, is a subjective measure
resulting from psychological factor and physical strength of the sound.[32] The
physical strength of the sound is often measured by sound pressure level, which
is a logarithmic measure of the sound pressure with respect to a reference value.
Denoted by SP L and measured in decibel dB, the sound pressure level is defined
as:
P
SP L = 20log10 (1.1)
P0
where P is the root mean square of the local sound pressure, P0 is the reference
sound pressure. In air, P0 = 20µP a, which is the threshold of human hearing.[21]
Figure 1.2 illustrates the decibel of a variety of sources, and the uneasiness and
damage caused to human hearing.
In free space, the sound pressure of a spherical sound wave decreases with the
inverse of the distance travelled, and thus the sound intensity decreases with the
inverse square, known as the ”inverse square law”. The propagation of noise is
also influenced by the convection of the medium, the obstacle presented on the
path, and the absorption effect due to viscosity. Approaches of noise abatement
therefore fall into two categories: acoustic insulation and source control. Since the
efficacy of acoustic insulation, with the installation of sound absorbing materials, is
relatively low and is often confined to small scale (such as room-acoustic), source
control is more effective and practical in the case of aviation operations. The
source control is fulfilled by adjusting the geometric or kinetic parameters such as
to altering the noise generating flow features, and thus the understanding of the
Fixed-wing aircraft powered by jet engine is prototyped for airlines at the moment.
Two major noise sources are identified for this configuration of aircraft: the jet
engine and the airframe. Early generation of turbo-jets were major source of
aeronautical noise, because of the increasing power generated for takeoff. The
noise from the engine is positively related to exhaust velocity, the relation of which
is described by the eighth power law (Lighthill,1952). With the development and
implementation of the large by-pass ratio turbo-fan engines on the civil aircrafts,
the exhaust velocity of the engine is lowered. Additional designs such as engine
nacelles with acoustic liner and chevron nozzle further contribute to the engine
noise attenuation. Nowadays, the engine noise has been markedly reduced to
the level comparable or even lower than the level of airframe noise. To reduce the
overall noise level of the aircraft, attention should be shifted to the noise generated
by airframe. The main noise generating airframe elements are:[62]
Figure 1.3: Example of AFN source localizations of a landing A340 with high-lift device
and landing gears (from flight test campaign)[62]
This study aims at the development and validation of a technique which predicts
the far-field noise based on the velocity field measured by the PIV system in
the vicinity of the solid body immersed in the flow, based on the measurements
acquired by time-resolved PIV and microphones. The outcome is a first step to
the source control of the airframe noise.
A Gurney flap is a lift-enhancing device of simple configuration, typically with
the shape of a small tab protruding perpendicularly into the flow. It was first
implemented on race car by Daniel Gurney to enhance the turning capability and
nowadays is utilized on wind turbines for lift increment [61]. In this preliminary
design of the implementation of PIV in flap noise measurement, the Gurney flap
serves as a simplified model of the flap on the aircraft, such as to facilitate the
model manufacturing and experimental set-up.
Chapter 2 starts with an introduction of the Gurney flap aerodynamics and then
gives an overview of the aeroacoustic analogies. In chapter 3, the mathematical
tools applied in the data reduction procedure of this thesis work are introduced,
following which, the bridging techniques between the flow measurements and the
application of acoustic analogies are presented. Chapter 4 continues with a descrip-
tion of the experimental techniques employed in the aerodynamic and aeroacoustic
measurements in the course of the present study, before the detailed experimental
set-ups are described in chapter 5. Results are presented in chapter 6 and the
conclusions are reached in chapter 7.
Aerodynamics is a study of the motion of the air, particularly its interaction with
the solid body submerged in the flow, which not only enables the calculation of
forces or energy that can be harnessed, but also is essential in understanding the
methodology behind the design and construction of airborne crafts and fluid ma-
chines.
Aeroacoustics studies the mechanism of noise generation aerodynamically, either
by the turbulent motion (e.g. noise emitted by high speed air jet), or by the in-
teraction of aerodynamic forces with the surfaces (e.g. ’singing’ telephone wires).
Although the scientific theory of the noise generation by aerodynamic flows has
not been fully established, yet practical analysis that relates the hydrodynamic
flow phenomena and the wave-like propagating disturbances has been investigated
extensively for various flow conditions.
This chapter first gives an overview of the flow features and lift-enhancing mech-
anism of the Gurney flap in section 2.1. Thereafter, a general introduction to
aeroacoustic analogies is given in section 2.2 followed by detailed information on
Lighthill’s analogy and Curle’s analogy.
Since the first implementation on the race car by Daniel Gurney in the 1970s,
the flow properties of the Gurney flap are widely investigated in literatures.
Lieback(1978) [40] was the pioneer of investigating the aerodynamic performance
by experiments. His experiments on a Newman airfoil with 1.25% the chord length
proved the Gurney flap to be lift-enhancing. At the same time, though counter-
intuitive, the Gurney flap was found to be drag-reducing. Lieback further sug-
gested the height of the Gurney flap should be kept between 1% to 2% of the
chord length to optimize the aerodynamic benefits, while the flaps with height
exceeding 2% chord length will substantially increase the drag. A more detailed
study on the scaling of Gurney flap by Giguère concludes that the Gurney flap
height should be less than the local boundary layer thickness to avoid the increase
in drag.[52]
The efficacy of Gurney flap in improving the lift characteristics of airfoils, wings
and aircraft models has been confirmed with the measurements from various stud-
ies, such as [48], [33] and [39]. Pressure and velocity measurements have helped
shedding light on the mechanism responsible for the lift-enhancement. However,
compared to the widely observed improvements in aerodynamic performance, the
mechanism is not equally unveiled. Pressure and velocity measurements have
helped shedding light on the issue. Jeffrey et al. (2000) [33] attributed the lift-
enhancement for subsonic airfoils to two major causes:
1. The Gurney flap decelerates the flow upstream, and consequently the pressure
difference of the lower surface and upper surface of the airfoil is increased at
the trailing edge. In such a way, Gurney flap adds to the effective camber of
the airfoil.
2. The long wake downstream the Gurney flap increases the trailing edge suction
and thus delays the flow separation on the upper surface of the airfoil.
Either cause leads to the increase in airfoil circulation, which results in the increase
in lift. Notably, Jeffrey et al. suggested in the same article that the flow structure
Figure 2.1: Visualization of counter-rotating vortices downstream the trailing edge for 4%
Gurney flap with α = 0◦ ,4◦ and 8◦ [64]
2. Second mode: this mode is the result of the intermittent ejection of the
circulating fluid in the cavity formed by the Gurney flap and the airfoil.
According to Troolin et al., the second shedding mode induces negative normal
velocity in the wake of the airfoil and thus contributes to the increase in circula-
tion. However, the observation of the second shedding mode has, curiously, not
been reported by other researchers until the writing of this thesis.
For further information and detailed theories about Gurney flap, one can refer to
[69], which is a compendium summarizing the aerodynamic characteristics, lift-
enhancing mechanism, and optimum design of Gurney flap, as well as the appli-
cation of Gurney flap on airfoils, wings and aircraft models ranging from subsonic
to supersonic.
Aeroacoustic analogies describe the dependence of the flow generated sound field
and its causes, the fluctuations in the flow. The notion of aeroacoustic analogy
(also mentioned as ’acoustic analogy’) was first introduced by Sir James Lighthill
in 1952, with his ’On sound generated aerodynamically’ [41], which placed the
noise from jet engine under the scientific scrutiny and later is acknowledged as the
origin of modern aeroacoustics. [38]
In several decades, aeroacoustic analogy has burgeoned into a big family. Despite
the difference in the choice of field variables, the way of interpreting source terms
Lf = g (2.1)
In the general form, f is the field variable used by the analogy, such as density,
velocity or enthalpy. L is the wave operator that relates the field variables in such
a way that the fluctuation of these variables propagates like waves. g is the source
term that has to be defined a priori, or the equation of aeroacoustic analogy should
be solved iteratively to increase the accuracy of the source term. In most cases,
instead of using the true acoustic source distributions, the source term is defined
by collecting from the field variable terms left over by the wave operator on the
left hand side.
The aeroacoustic analogies can be sorted into three categories based on the field
variables applied. The complexity of the application of the following analogies
grows in the order they are presented below.[65]
1. Density-based analogies
• Lighthill’s analogy
• Curle’s analogy
• Powell’s analogy
• The Ffowcs Williams-Hawkings analogy
• Phillip’s analogy
• Lilley’s analogy
3. Enthalpy-based analogies
• Howe’s analogy
• Doak’s analogy
2. Theories designed for particular model or relating the sound field to the
pressure distribution model on or near the edge of the plate
Gurney flap noise is a type of trailing edge noise and falls into the realm of the
first category, where the far-field radiations determined in terms of the turbulent
velocity field. Examples of the theories in this category are developed by Ffowcs
Williams and Hall[74], Crighton and Leppington[8] and Howe [29].
Since the present work is a first step towards the prediction of trailing edge noise
of the airfoil with Gurney flap, the Curle’s analogy, a special (and simpler) case of
Ffowcs Williams-Hawkings analogy, is chosen to educe the far-field noise from the
PIV measured velocity field for the simplicity of numerical application. Therefore
for the reason of brevity, only the physical background and working equations of
the Lighthill’s and Curles analogy are presented. Note that the Lighthill’s analogy
is the antecedent of Curle’s analogy. For information of other acoustic analogies
and for detailed mathematical derivations, one can refer to[41], [9], [17], [27], [28]
and [65].
Since both Lighthill’s analogy and Curle’s analogy are density-based aeroacoustic
analogy, the field variable f in Equation 2.1 is substituted with ρ0 , the fluctuation
of the density ρ. In the case of current project, it is equivalent to choose the
pressure perturbation (sound pressure) p0 as field variable. The argument goes as
follows.
Apart from the problems where energy is transferred by heat conduction, such as
frictional dissipation of sound, highly non-linear events caused by the shock waves
or sound generation by heat sources(e.g. combustion), it is sufficient to consider
the flow to be homentropic.[28] Namely, the entropy s of the fluid is uniform and
constant, and thus the energy equation can be substituted by:
s = s0 = constant (2.2)
where R is the gas constant, γ = Cp /Cv is the specific heats ratio and T is the
absolute temperature in K.
Under the homentropic condition, local pressure and density obey the Poisson
relation: p ∝ ργ . Combined with Equation 2.3, results in a linear relation between
small fluctuation in density and pressure:
p0 = c20 ρ0 (2.4)
where σij is the (i, j)th component of the viscous stress tensor. For a Stokesian
gas, it is defined by the constitutive equation reads:
∂vi ∂vj 2 ∂vk
σij = µ + − δij (2.7)
∂xj ∂xi 3 ∂xk
where µ is the viscosity of the fluid. Multiplying Equation 2.5, adding the result
to Equation 2.6, the result of which upon adding and subtracting c20 ∂ρ/∂xi , yields:
∂ρvi ∂ρ ∂Tij
+ c20 =− (2.8)
∂t ∂xi ∂xi
where
is the Lighthill’s stress tensor. Differentiate Equation 2.5 with respect to t, take the
divergence of Equation 2.8, and then subtract the results to obtain the Lighthill’s
equation:
2
∂ 2 Tij
∂ 2 2 0
− c 0 ∇ ρ = (2.10)
∂t2 ∂xi ∂xj
The subscripts ”0” used in p0 and ρ0 in Equation 2.9 denotes the constant refer-
ence value, which is taken the value of corresponding property at large distances
from the sound generating turbulent region. The reason for this arrangement
is that, at the distance sufficiently far away from the turbulent region, the flow
recovered to be homogeneous, with uniform equilibrium state: speed of sound c0 ,
density ρ0 and pressure p0 .
...the pressure fluctuations have some of the properties of both the localized
fluctuations and those in the sound field. Thus, in this region, the pressures are as
weak as in the sound field, but the distances involved are small enough so that the
effects of finite propagation speed, and hence of compressibility, can be neglected.
Tij ≈ ρ0 vi vj (2.12)
With the approximations on Lighthill’s stress tensor (Equation 2.12), and the
relation between fluctuation in pressure and density (Equation 2.4), the solution
reduces to:
1 ∂2
Z
0 r dV
p (x, t) = ρ0 vi vj y, t − (2.16)
4π ∂xi ∂xj Vy c0 r
Discussion
Lighthill’s equation treats the sound sources convecting with the same speed of
the flow in a way as if they were moving in the stationary medium. It incorporates
the sound generation, convection with the flow, propagation with different speed,
and dissipation due to conduction and viscosity.
Under the flow conditions:
3. the spatial gradients of the speed of sound and the static hydrodynamic
parameters are, at most, of the order of the perturbation
The solution given to Equation 2.10 by Lighthill is for unbounded flow, which does
not include the influence of solid boundaries submerged in the flow, although at the
time Lighthill has remarked on the importance of solid boundary in the generation
of sound. In fact, in many cases of technological interest, solid boundaries play
a vital role in sound generation, such as high-lift devices, propellers, helicopters,
windmills and turbofans.
The presence of solid boundaries in the flow influence the sound radiation in two
ways, the mechanisms of which lie respectively in theories of classical acoustics
and aeroacoustics:
• The solid boundaries reflect and diffract the sound generated by the volume
quadrupole sources that appear in Lighthill’s solution.
• The solid boundaries support the distribution of dipole sources; in the case
where the solid boundaries are in motion, monopole sources can be generated.
The limitation was overcome by N. Curle in 1955 [9]. The solution given to the
Lighthill’s equation by Curle takes the effect of solid boundaries into account and
is later known as the Curle’s equation.
writes:
t+ t+
∂ 2G
Z Z Z Z
0 ∂Tij ∂G
p (x, t) = Tij dV dτ + G − Tij ni dSdτ
−∞ Vy ∂yi ∂yj −∞ ∂Vy ∂yj ∂yj
Z t+ Z
∂ρ0
2 0 ∂G
−c0 ρ −G ni dSdτ
−∞ ∂Vy ∂yi ∂yi
(2.17)
When the surface of the solid body is rigid and stationary, the solution of far-field
pressure can be written as:
1 ∂2
Z Z
0 Tij 1 ∂ Pij
p (x, t) = dV − ni dS (2.18)
4π ∂xi ∂xj Vy r t=te 4π ∂xj δVy r t=te
where Pij = p0 δij −σij is the generalized pressure tensor formed by the combination
of pressure fluctuation and viscous stress. te is the retarded time.
The volume integral and surface integral of Equation 2.18 (respectively the first
and second term on the right hand side) are often referred to as quadrupolar and
dipolar source terms.
It can be noted that the presence of solid boundaries engenders a dipole source.
This phenomenon is caused by the prevention of pressure recovery by fluid
acceleration due to the solid boundaries. The flow in turn exerts a fluctuating
force on the solid body, in the form of time-varying surface pressure. The
fluctuating force is directly related to a dipole acoustic source and causes a higher
level of noise emission compared to the free turbulence.
Figure 2.3 shows the radiation patterns of monopole, dipole and quadrupole
sources. It can be seen that the monopole source is the most efficient for low
Mach number and is omnidirectional, while the emission of dipole and quadrupole
is directional due to mutual cancellation. Take dipole as an example: a dipole is
a separation of a sink and a source. The line connecting the two coupled source
in dipole is called the axis of dipole. Along the axis, the direction of the dipole is
defined as pointing from sink to source. The positions on the line perpendicular
to the axis of dipole do not perceive any disturbance from the source.
As mentioned previously the fluctuating of aerodynamic force is directly related
to the dipole source. It can be further argued that, the strength and direction of a
dipole source is equivalent to the strength and direction of the force exerted by the
solid body on the external fluid. [75] It can be thus concluded that for an airfoil,
the strongest noise emission can be expected on the direction perpendicular to
the chord.
Figure 2.3: Theoretical directivity patterns for far-field sound pressure levels radiated from
(a) monopole, (b) dipole, (c) lateral quadrupole, and (d) longitudinal quadrupole sound
sources. [59]
rying the sound generating vortices. The compact relation can be written as:
L/λ << M << 1, which indicates that at low Mach number, the flow induced
sound source are more likely to be compact.
For low Mach number (M < 0.05), the compressibility effect of the flow can be
neglected. Furthermore, for chord Reynolds number (105 ), though the viscosity
plays an important role in the near wall region, the overall effect can still be ne-
glected when compared to the normal stresses. Then, the contribution of viscous
term σij in the dipolar source can also be neglected [70].In addition, under such
circumstances, the magnitude of the quadrupolar source is M 2 order lower than
the dipolar term and therefore can be neglected [17]. Thus the solution of far-field
pressure fluctuation reduces to:
p0 δij
Z
0 xj ∂
p (x, t) = − ni dS (2.19)
4πc0 |x| ∂t δVy r t=te
Discussion
Curle’s analogy is the solution to the Ligththill’s equation Equation 2.10 when
the effect of rigid, stationary solid boundaries is taken into consideration. The
presence of solid boundaries engenders a dipolar source term in the solution,
which radiates sound more efficiently than the quadrupolar source introduced by
free turbulence. Especially, at low Mach number, the contribution of quadrupolar
source can be neglected at the presence of dipolar source.
When the source satisfies the condition of compactness, with the application of
Gutin’s principle, the term of the integral of surface pressure in the dipolar source
term can be substituted by the instantaneous aerodynamic forces. Since there are
ready techniques for evaluating or measuring fluctuating aerodynamic loads, the
simplification of Curle’s analogy not only facilitates the numerical application,
but also lowers the demand for experimental set-up.
Figure 2.4: Root mean square of the source terms retrieved from PIV measurements of a
2D cavity by Koschatzky et al. (2010)[36]
Figure 2.5: Power spectra of sound signal from a microphone and of sound computed
from PIV at the same locations with the Curle’s analogy and the vortex sound theory by
Koschatzky et al. (2010)[36]
Figure 2.6: Power spectra computed using Curle’s analogy with the microphone measure-
ment. Predictions based on full span coherence assumption and on the measured spanwise
coherence length by Lorenzoni et al.(2012) [43]
The following sections offer a general introduction on the techniques for analysing
the experimental data. First, the mathematical background of statistical data
analysis methods are introduced, including the causality-correlation, the spectral
density calculation and the variance and covariance. Then, the techniques for
evaluation of pressure field and aerodynamic loads are presented, which are the
bridging techniques between the PIV measured velocity and the application of
Curle’s analogy.
This section gives a brief introduction to the data analysis techniques applied in
the current study for analysing the relationship between near-/far-field.
3.1.1 Causality-correlation
where ψ 0 (x, t) is the zero-mean part of the near-field quantity, such as velocity
fluctuation u0 and v 0 , measured at position x and time t; p0 represents the acoustic
pressure measured in the far-field; τ is the time shift between between the far-field
pressure and near-field quantity.
The cross-correlation is similar to the covariance of two functions except for the
terms in denominator, σψ (x) and σp , which are root-mean-square (RMS) of the
near and far field variables. The normalization with the RMS results in bounding
the value of cross-correlation coefficient Rψ,p (x, τ ) within [−1, 1]. The bounding
value 1 or -1 is taken when the near and far field signals display a perfect linear
relationship. Non-linear relationship or data scattering due to measurement noise
will force the value Rψ,p to zero.
It should be noted that correlation does not necessarily imply causality, while the
latter should be validated by other proof. In the current study of far-field noise
generated by an airfoil with Gurney flap, the proof of causality relation is the
Curle’s analogy described by Equation 2.18. Although Equation 3.1 is given in a
time domain, yet it is equivalent to the spatial correlation between near and far
field variables in this study, since the retarded time and the spatial distance are
related with a constant speed of sound.
Since the experimental data is acquired at discrete time instants tn , Equation 3.1
should be applied in discrete form:
N
1 X 0
Cψ,p (x, τ ) = [ψ (x, tn )p0 (tn + τ )] (3.2)
N n=1
with
v
u
u1 X N
σψ (x) = t ψ 02 (x, tn ) (3.3)
N n=1
The discussion in this section will focus on the first two approaches.
Assume the input and output signal are mono-dimensional continuous time signals
x(τ ) and y(τ ), then the general formulation of cross-correlation function is written
as:
Z T
1 2
Rxy (τ ) = lim x(t)y(t + τ )dt (3.4)
T →∞ T − T2
when x(τ ) = y(τ ), Rxy (τ ) becomes the auto-correlation function Rxx (τ ) (or
Ryy (τ )).
The Fourier transform of the correlation function yields the Power Spectral Den-
sity:
Z
Sxy (f ) = Rxy (τ )e−2iπf τ dτ (3.5)
where Sxy (f ) is the two-sided cross-spectral density function and is an even func-
tion of frequency f . Here, the value of f can be both positive and negative.
Since the negative frequency and the positive frequency is equivalent, one-sided
cross-spectral density can be defined for f ≥ 0 as:
(
2Sxy (f ) f >0
Φxy (f ) = (3.6)
Sxy (f ) f =0
The auto-spectral density Φxx (or Φyy ) can be defined similarly. Equation 3.5 is
often referred to as the Wiener-Khinchine relations in honour of the two mathe-
maticians N. Wiener and A. I. Khinchine.
Alternatively, the spectral density can be derived based on finite Fourier trans-
form, which is better for engineering application since the measurements in real
world can only be acquired in finite time interval 0 ≤ t ≤ T . The cross-spectral
density is defined as:
1 ∗
Sxy (f, T, k) = X (f, T )Yk (f, T ) (3.7)
T k
Xueqing Zhang M.Sc. Thesis
3.2 Evaluation of pressure from velocity field 31
where Xk (f, T ), Yk (f, T ) are the Fourier transforms of the sample records xk (t),
yk (t), with Xk∗ (f, T ) the complex conjugate of Xk (f, T ).
To extrapolate Sxy (f, T, k) to the infinite time range, expected value of Sxy (f, T, k)
should be used for statistical consistency:
The definition given by Equation 3.7 and Equation 3.8 is often referred to as the
Welch’s method, named after P. D. Welch.
3.1.3 Coherence
Coherence, similar to cross-correlation, evaluates how well the two signals are
related, except that cross-correlation evaluates the phase lag of the two signals in
the time domain, while coherence evaluates the dependency of the average phase
angle for each frequency components of the two signals.
The coherence function is defined based on the cross-spectral and auto-spectral
density function, which can be used to examine the relation between two signals,
such as the investigation of the power transfer between input and output of a linear
system or estimation of the causality.
The coherence between two signals is defined as:
2 |Φxy (f )|2
γxy (f ) = (3.9)
Φxx (f )Φyy (f )
Φxy (f ), Φxx (f ) and Φyy (f ) are the cross- and auto-spectral density as defined in
subsection 3.1.2.
Equation 2.19 and Equation 2.20 both require the pressure distribution on the
airfoil. The pressure is related to the velocity by the Navier-Stokes equations and
the spatial gradients of pressure can be expressed in either the Eulerian ( Equation
3.10) or the Lagrangian formulation ( Equation 3.11) as:
∂u
∇p = −ρ + (u · ∇)u − ν∇2 u (3.10)
∂t
Du
∇p = −ρ − ν∇2 u (3.11)
Dt
For the data acquired from PIV measurements, the pressure distribution needs
to be reconstructed from the velocity field. Two strategies are available for the
computation of the pressure field:
In strategy (1), the pressure gradients are integrated with spatial marching scheme
from the boundaries where Dirichlet boundary condition is assigned. The method
introduces the accumulation of random errors along the integration path and thus
a dependence of the accuracy on the path. Liu and Katz (2006) [42] have de-
veloped an omni-directional virtual boundary integration scheme to integrate the
acceleration while minimizing the effect of local random error. Strategy (2) takes
the in-plane divergence of Equation 3.10 or Equation 3.11 such that the equation
about pressure is rearranged into a Poisson equation:
∂ 2p ∂ 2p
+ = −ρfxy (3.12)
∂x2 ∂y 2
where the source term fxy is a function of velocity and can be decomposed into
the component of 2D and 3D. For high Reynolds number, the viscous term can be
Du
fxy = f2D + f3D = −ρ∇xy
" Dt 2 #
2
∂u ∂v ∂u ∂v
= +2 +
∂x ∂x ∂y ∂y
∂(∇xy · u) ∂w ∂u ∂w ∂v
+ + (u · ∇)(∇xy · u) + +
∂t ∂x ∂z ∂y ∂z
(3.13)
Test cases applying Equation 3.12 can be found in literatures such as [11] [12].
The different approaches (Lagrangian and Eulerian) yield different performance.
Violato et al. (2011) [68] compare the two approaches on a rod-airfoil configu-
ration and offer guidelines to the choice of experimental parameters according to
the approach adopted. According to the authors, the Lagrangian approach is less
sensitive to measurement noise. De Kat and van Oudheusden (2012) [10] fur-
ther assessed the experimental viability of the approaches in the reconstruction
of instantaneous pressure field, giving emphasis on the flow features of convecting
vortices.
Figure 3.1: Control volume and Control surface around the cylinder [37]
Figure 3.2: Definition of Inviscid and Viscous region on the Control surface [37]
With the pressure field reconstructed in section 3.2, there is still one step
before the application of the lumped form Curle’s analogy ( Equation 2.20),
the evaluation of instantaneous aerodynamic forces exerted on the solid body.
This section introduces two method for evaluating aerodynamic loads from PIV
measured velocity field.
In this case the planar PIV measurement cannot resolve the out-of-plane velocity,
only 2D components are considered. In addition, with the outer boundary of the
control volume sufficiently far from the body surface, the viscous effect is neglected,
resulting in:
" # Z Z " ∂u # I " 2 # I " #
D ∂t
dxdy −u dy + uvdx −pdy
= −ρ ∂v
+ρ 2
+ (3.15)
L V ∂t
dxdy S −uvdy + v dx S pdx
The authors of [37] proposed a robust procedure for the evaluation of the pressure
terms on the control surface in Equation 3.15. As shown in Figure 3.2, the part
of control surface hit by the wake is defined as ’viscous region’, where pressure is
obtained by integrating Equation 3.11, but neglecting the viscous contribution.
To minimize the effect of error propagation in the process of spatial integration,
outside the wake, pressure is estimated by Bernoulli equation. In such a way,
pressure in the wake region can be resolved by integrating from both edges of
the wake, and the end point of integration, lying in the inviscid region, can be
compared with the value given by Bernoulli equation. The discrepancy between
the integration result and the Bernoulli equation can then be used to redistribute
the discrepancy along the viscous region with a linear weighting function. The
above mentioned technique is derived and applied to the incompressible scenario,
an extension of this method to the application of compressible flow can be found
in [66], where a test-case with the supersonic flow around a bi-convex airfoil is
presented.
with
1 1 1
γf lux = u2 I − uu − u(x ∧ ω) + ω(x ∧ u)
2 N − 1 N −1
1 ∂u ∂u ∂u
− x· I −x + (N − 1) x (3.17)
N −1 ∂t ∂t ∂t
1
+ [x · (∇ · T)I − x(∇ · T)] + T
N −1
Xueqing Zhang M.Sc. Thesis
3.3 Evaluation of aerodynamic forces from velocity field 37
where n̂ is the surface normal, N is the spatial dimension, ω is the vorticity, x is the
position vector, I is the unit tensor and T is the viscous stress tensor. The authors
of [50] recommended that the forces be evaluated from the data across the region
of the wake where the vorticity and viscous effects are present since it is probable
that the surface integration of the product of x and ∂u/∂t may not converge in
the far-wake. Although in 2-D measurements, the divergence-free condition may
not be satisfied for the ”Flux equation”, yet concerning the spurious divergence of
velocity field generated by the measurements in the boundary layers of the solid
body, the ”Flux equation” is still strongly preferred in such situation.
Experimental techniques
For engineers, theories are not the central objectives, nor an end in itself,
but a means to the utilitarian end: the incarnation of the idea of people.
Quotes of statement made by a British engineer to the Royal Aeronautical So-
ciety in 1922 acknowledged the gap between scientific theories and engineering [67]:
Aeroplanes are not designed by science, but by art in spite of some pretence and
humbug to the contrary. I do not mean to suggest for one moment that engineering
can do without science, on the contrary, it stands on scientific foundations, but
there is a big gap between scientific research and the engineering product which
has to be bridged by the art of the engineer.
Experimental research is one of the art of engineer, which not only is effective in
validating the prediction made by scientific theories, but also provide information
that are out of the reach of modern theories, furthermore, can reversely produce
analytical concepts and way of thinking, as with the case of Durand and Lesley’s
idea of propulsive efficiency of a propeller in front of an obstruction.
where is the dielectric constant of the substance between the two plates, d and
A are the dimensions of the capacitor.
The diaphragm is excited to vibrate by the impinging pressure fluctuation caused
by the acoustic wave, and thus the distance between the two plates d is changing
with the variation of impinging pressure on the diaphragm. When the charge stored
by the capacitor Q is kept constant, according to the capacity law, the voltage
V between the two plates responses linearly and continuously to the pressure
fluctuation:
Q Q
V = = d (4.2)
C A
The resistor functions as a high pass filter. In such a way, the output voltage
measures only the fluctuating part of the pressure signal.
Microphone sensitivity defines the ratio between the output voltage and the input
pressure signal. The sensitivity is depicted in two dimensions:
• Directional sensitivity
Directional sensitivity, or directionality, describes the microphone’s sensitiv-
ity to sound from various directions.
• Frequency sensitivity
Frequency sensitivity, or angular sensitivity, describes the amplitude response
of microphone to signal of different frequencies.
Figure 4.3 shows the directional sensitivity of Linear X-M53. It can be observed
that in audible frequency range (20 Hz to 20 kHz), the model is essentially omni-
directional. When moved to higher frequencies, the sensitivity lobe moves to the
front and the directionality becomes hypercardioid.
Figure 4.4 shows the frequency sensitivity of Linear X-M53. It can be observed
that from 50 Hz to 10 kHz, the ratio between output and input signal is roughly
constant. The reason behind this phenomenon is due to the schematic of the
condenser microphone. In the range of frequency below 50 Hz, the amplitude
of fluctuation is large due to the large amount of energy contained by the low
frequency components. The displacement of diaphragm is insufficient to measure
the amplitude of fluctuation in this frequency range, resulting in the low sensitivity.
In the frequency range above 10 kHz, the response first increases then quickly drops
because of the insufficiency of the dynamic response. The time lag introduced by
the diaphragm-condenser system causes out-of-phase cancellation.
Microphone response can be corrected to acquire a constant response with the
frequency, which requires a source in anechoic environment with known constant
amplitude and position.
only offer a comparison of acoustic pressure at certain locations and thus does not
include the mapping of acoustic source. Therefore the major concern of micro-
phone placement lies in avoiding the near-field effect while having sufficient signal
to noise ratio.
Microphones placed in the near-field region measure the radiation sound and the
pseudo-sound at the same time, resulting in the loss of representation of the sound
propagating to the far-field. To avoid this effect, acoustic measurements are usu-
ally conducted in the far-field, in which the microphone is placed at least one
acoustic wavelength and two source dimensions away from the source.
On the basis of far-field measurement, the microphone should be placed in the
region close to the source, such that the sound level of the source is strong enough
to dominate the background noise as well as the reflections. The signal-to-noise
difference is preferred to be above 10 dB. In occasions where 10 dB is not achiev-
able, a lesser difference is tolerable as long as the interfering noise can be separated
from the data.[47]
From Equation 1.1, the root mean square of sound pressure can be calculated by:
SP L
P = P0 · 10 20 (SP L = 114dB) (4.3)
And with the root mean square of voltage V acquired from the microphone record-
ing, the output-input ratio H can be thus written as:
V
H= (4.4)
P
Figure 4.1: The technical chart, the dimensions and the equivalent schematic of a Linear
X-M53 microphone[31]
MSc. Thesis Xueqing Zhang
46 Experimental techniques
δx
U= (4.5)
δt
The idea of PIV is derived from the laser speckle velocimetry used by several
research groups in late 1970s, and was greatly developed in the following decades.
The availability of high-speed lasers and cameras allows for variations upon the
original idea, such as:
• Stereoscopic PIV
Stereo-PIV measures three velocity components in a 2D plane with two cam-
eras.
• Tomographic PIV
Tomo-PIV measures three velocity components in a volume using two, or
more cameras.
• Time-Resolved PIV
Time-Resolved PIV benefits from the advances in CMOS camera technology
to enable the measurement with high time-resolution.
4.2.1 Seeding
Tracer particles play a vital role in PIV. There are essentially three requirements
for the particles:
• The particles should exactly follow the motion of the fluid flow.
To choose the proper tracer particle that meets the demands, the mechanical
properties and the scattering properties of the particle should be considered.
Mechanical properties
For very small particles in low speed motion, the quasi steady viscous term, namely
the Stokes drag dominates the particle dynamics. The difference between the
particle velocity V and that of the surrounding fluid U is used to describe how
well the tracer particle can follow the flow. This difference is often referred to as
slip velocity and is defined as:
2 ρp − ρf dV
V − U = a2 (4.6)
9 µ dt
where
1 2 ρp − ρf
τp = d (4.8)
18 p µ
Theoretically, Stokes number should be St << 1. In gas flow, for practical
limitations, St < 0.1 is used.
Scattering properties
The scattering efficiency of the tracer particles depends on:
• particle diameter dp
Wavelength is restricted by laser and camera sensor and thus cannot be chosen
freely. To acquire a high refractive index ratio, material of high index of refraction
is preferred. Furthermore, according to Mie’s theory, particle of large diameter is
advantageous.
It can be seen that the indications in the choice of particle diameter given by me-
chanical properties and scattering properties are contradictory. The most common
approach is to choose the largest particle that can follow the flow accurately.
4.2.2 Illumination
Due to the working principle of PIV measurements, there are several requirements
on the illumination. First, the particles should be sufficiently illuminated such that
the scattered light can be captured by the camera sensor. Second, the illumination
should be in short duration, such that the particle images do not appear as streaks.
Third, the light provided by the light source can be shaped into thin sheet that
only particles lying within can be imaged in focus.
For the requirements mentioned above, in PIV setups, lasers are often used as the
light source due to their ability to provide light beam that is:
4.2.3 Imaging
The imaging system is characterized by the focal length f , the aperture number f#
and the magnification factor M . Focal length and magnification factor are related
to image distance di and object distance do in the thin lens approximation, by:
1 1 1 di
= + ;M = (4.9)
f di do do
The aperture number (also referred to as f-number), is defined as the ratio of focal
length f and the aperture diameter D. However, in practice, f# is determined
by the magnification factor and the thickness of the light sheet. The relation is
explained as follows.
The focal depth δz is the depth within which the particles are in focus, and is de-
scribed by the wavelength of the light source λ, the f-number and the magnification
factor as:
2
M +1
δz = 4.88λf#2 (4.10)
M
δz should be larger or equal than the light sheet thickness ∆z, which is about 1mm
in practice.
There is a trade-off in the choice f-number, since a large f-number allows a higher
depth of field while a small f-number enhances the light intensity collected by the
sensor. This issue is far from trivial for recordings with high repetition rate where
the power per pulse decreases.
The diameter of particle image dτ depends on the geometric optics effect and the
diffraction effect:
q
dτ = (M dp )2 + (ddif f )2 (4.11)
where
Typically, the term (M dp ) is much smaller than ddif f and hence can be neglected.
The particle image diameter is a parameter concerning the accuracy of velocity
measurements. When the particle image occupies less than one pixel, it is referred
to as peak locking, which lowers the accuracy of measurement since the particle
positions are locked to integer values and cannot be detected with subpixel accu-
racy.
4.2.4 Recording
High-speed cameras used in PIV measurements usually store the images in memory
on board during the measurement and later transfer to the computer. For the
current study, cameras with CMOS sensors are used. The CMOS cameras available
offer frame rates of over 5000 frames per second for full mega pixel resolution
(1024 × 1024 pixels). Higher frame rates can be reached when sensor is cropped.
For example, at 512×512 pixel resolution, 10,000 frames per second can be reached.
In the current study, Double Frame/Single Exposure acquisition is used. The
camera shuttering time and the laser pulses are synchronised in so-called frame
straddling mode, as shown in Figure 4.8, where Tshut is the camera shutting and
data transfer time, δt is the laser pulse separation and ∆t is the acquisition time
of an image couple.
Figure 4.8: Timing diagram for PIV recording based on Double Frame/Single Exposure
frame straddling mode [44]
To evaluate the velocity of the flow field, the displacement of particles during the
pulse separation δt should be evaluated from the two subsequent image frames.
The steps and rules to follow are described as follows.
1. Image windowing
Since the displacement of particles is usually not uniform on the entire region,
the image is first partitioned into small windows, inside which the displace-
ment is assumed to be uniform. To optimize between the pulse separation
and window size, there are a few rules:
2. Cross-correlation analysis
Cross-correlation is applied the image pair to determine the particle displace-
ment in each window. Let I(x, y, t) and I 0 (x, y, t + δt) the two interrogation
window at time t and time t+δt as shown in Figure 4.9. The cross-correlation
function Φ is computed in its discrete form reads:
PI PJ
i=1 j=1 IA (i, j) · IB (i + m, j + n)
Φ(m, n) = qP P (4.13)
I J 2
PI PJ 2
i=1 j=1 IA (i, j) · i=1 j=1 IB (i, j)
3. Sub-pixel interpolation
The displacement in pixels defined by correlation coefficient Φ defines only
discrete pixel locations. In order to reach sub-pixel precision, the correlation
peak is interpolated, typically by means of Gaussian interpolator, around the
maximums as shown in Figure 4.10. There are additional peaks besides the
main peak on the correlation coefficient plot due to noise and correlation of
non-paired particles. Assume the value of highest peak be Φ1 , the second
highest peak Φ2 , and the correlation signal-to-noise ratio can be defined as
shown in Equation 4.14.
Φ1
SN = (4.14)
Φ2
Usually, small interrogation window is preferred for higher vector resolution, espe-
cially in the case where fine flow patterns are involved. However, this requirement
give rise to the problem of the reduction in image density NI and hence the infor-
mation contained in one interrogation window. Another disadvantage with small
interrogation window is particles escaping the computation domain, namely Loss-
of-pairs. The above mentioned two defects both lead to rise in the uncertainty of
particle image displacement.
The Multi-pass interrogation is a solution to this issue. The method is an iterative
process, in which a relatively large interrogation window is used to produce an
initial velocity approximation; then, the approximated velocity is used to shift (or
deform) the correlation windows. In each iteration, the window size decreases and
corrections are applied to the location of correlation peak.
Peak-locking should be avoided for reliable PIV measurement, since particle image
much smaller than pixel size leads to discontinuous velocity field. Theoretically,
f# should be increased for larger particle image. However, in the case where the
pulse duration is short due to high acquisition frequency, small f# is required for
the sake of image contrast. In such case, a slight defocusing of the camera lenses
helped to avoid this effect in practice.
Experimental set-up
All experiments of this study are carried out with the facilities and equipment
available on campus of Delft University of Technology. The following sections give
a brief introduction on the experimental facilities and models. The wind tunnel
and test sections used for this study are introduced in subsection 5.1.1, following
which, the airfoil and trailing edge models are presented in subsection 5.1.2.
The wind tunnel test was carried out at the vertical low turbulence wind tun-
nel (V-tunnel) located in the Low Speed Laboratory (LSL) at Delft University of
Technology. The V-tunnel has an open test section with a vertical inflow coming
from below, with an operating velocity range of 5 - 45 m/s. Figure 5.1 shows the
schematics of the V-tunnel.
A circular exit with a diameter of 0.6m leads the inflow to the open section. In this
study, due to the requirements of model installation and acoustic measurement, a
transition duct was installed on top of the circular exit to accommodate the square
test section sized 0.4 × 0.4m2 .
The reasons for choosing V-tunnel for this study are: (i) the low turbulence level
of the inflow (below 0.1%) due to the high contraction ratio between the settling
chamber and the open-section exit, such as to provides incoming flow with high
flow quality, and (ii) the relatively low operating noise level, which makes the tun-
nel suitable for accommodating acoustic measurements in absence of an anechoic
wind tunnel. For PIV measurement alone, the closed test section was built from
plexiglass. For synchronized PIV and acoustic measurement, two side walls of the
test section were substituted with Kevlar, which is transparent for acoustic waves
Figure 5.1: Schematic of vertical low turbulence wind tunnel (V-tunnel) in TU Delft [55]
A NACA 0015 airfoil with detachable Gurney flap sized 0%, 2%, 4% and 6% the
chord length were designed and manufactured for the wind tunnel experiments.
The airfoil have a chord length of 20cm and a span of 40cm. The detachable
trailing edge were made from plexiglass to allow the passage of laser light, by
which means the velocity field on both pressure and suction side of the trailing
edge can be captured simultaneously.
The 0% (or ’clean’) and 6% Gurney flap ( Figure 5.3) were chosen for this PIV-
based noise investigation by evaluating the results of the previous study given
by J.Shah. In addition, roughness elements (height around 0.841mm) made of
Figure 5.2: Test sections for stereoscopic PIV measurements(left) and synchronized
acoustic-planar PIV measurements(right) designed by J.Shah [63]
Figure 5.3: NACA 0015 and Gurney flap models designed by J.Shah [63] and manufactured
at TU Delft
CMOS sensors, placed facing each other on both sides of the test section, were
used for image acquisition. The FOVs of the two cameras have an overlap of 25%
(see Figure 5.7), which is used to stitch the two FOVs together in the procedure
of image processing. For the concern of high temporal resolution, the sensors of
the cameras were cropped to 704 × 704 to reach an acquisition frequency at 5kHz.
The flow is seeded with water-glycol based fog with size of particles dp ≈ 1µm.
PIV images are acquired in double frame mode at the acquisition frequency of
Table 5.1: Synchronized planar PIV and microphone measurement: experimental parame-
ters
Table 5.2: Synchronized planar PIV and microphone measurements: experimental devices
5kHz, with a delay between two frames of 50µs. Detailed parameters of PIV
recording can be found in Table 5.3.
Simultaneous velocity and acoustic measurements were performed since previous
works (reference [63]) found the quality of causality correlation to be poor when
velocity and acoustic measurements are not performed simultaneously. The two
types of measurements can be synchronized by recording the trigger input of the
CMOS cameras. Usually, the duration of microphone measurement is much longer
than that of the PIV measurement and thus, the trigger input can be used to crop
from microphone measurement the portion of recording that is simultaneous with
Table 5.3: Synchronized planar PIV and microphone measurements: recording parameters
1
In experiments, cameras were slightly defocused to make the particle image larger such as to avoid
peak-locking
5mm
4 8
1.3m
3 20cm
9
2
1
10
11
Figure 5.5: Schematic of set-up for synchronised planar PIV and microphone measurements(side view)
Planar PIV set-up: Top view
1 NACA 0015 with Gurney flap 5b Test section: plexiglass
2 Laser sheet 6,7,8,9 Microphone #1,2,3,4
3 Laser head 12 Camera #1
4 Mirrors and lenses 13 Camera #2
5a Test section: Kevlar
12
5b
5a 6,7,8,9
4
3 40cm
40cm
2
1
13
Figure 5.6: Schematic of set-up for synchronised planar PIV and microphone measurements(top view)
68 Experimental set-up
1.5c
1.2c
FOV Camera #2
FOV Camera #2
FOV Camera #1
FOV Camera #1
80mm
80mm
0.8c
0.5c
80mm
80mm
Figure 5.7: Two versions of Field of View for two-camera recording: FOV1(left) and
FOV2(right)
The experimental parameters and devices for the stereoscopic PIV measurements
are identical as those applied for planar PIV measurements (refer to Table 5.1 and
5.2), except that the stereoscopic PIV measurements were not synchronised with
acoustic measurements and therefore the microphones were not used.
The aim of stereoscopic measurements is to assess the spanwise coherence of the
flow structures. The spanwise coherence can be used for the correction of 3D effect
in noise generation, since noise predictions based on planar PIV measurements are
essentially based on 2D assumptions.
In the stereoscopic PIV measurements, the sheet of laser sheet was placed normal
to the free stream. The distance between the laser sheet and the Gurney flap is
15mm, which is approximately one Gurney flap height downstream the Gurney
flap. Velocity components of all three dimensions were captured, from which the
spatial derivatives in the spanwise and transverse dimension can be resolved.
The front and top views of the set-up of the stereoscopic PIV measurements are
illustrated in Figure 5.9. Detailed recording parameters of the PIV system can be
found in Table 5.4.
Figure 5.8: Photograph of the experimental set-up for stereoscopic PIV measurements
Camera 1 Camera 2
15mm
Laser sheet
NACA0015 with Gurney flap
(view from pressure side)
Camera 1 Camera 2
Pressure side
Plexiglass
Laser head
Pre-processing of the raw images lies in minimum intensity subtraction from the
time series such as to remove the unwanted reflections. Sequential cross-correlation
algorithm with multi-pass iterations of decreasing interrogation window sizes (from
64 × 64, 50% overlap, 2-pass, square-shaped window to 32 × 32 to 75% overlap,
3-pass, round-shaped window) is applied for vector field calculation.
Vector post-processing is applied to remove the outliers. The spurious velocity vec-
tors are detected by means of the normalized median test that yields a ‘universal’
probability density function for the residual and hence a single threshold value can
be applied to variety of flow conditions without a priori knowledge of the flow char-
acteristics to detect spurious vectors. The method is developed by Westerweel and
Scarano in 2005 [71]. The velocity fields evaluated by DAVIS 8.2 are shown in Fig-
ure 5.10.
To obtain the combined velocity field, the two velocity fields are first trimmed to
remove the verge effect (measurements of poor quality at the verge of the camera
sensor). The trimmed velocity fields are then concatenated with MATLAB pro-
gramme, which detects the overlap region and averages the vector values given by
the two vector fields. An example of the combined velocity field is given by Figure
5.11.
Figure 5.10: Velocity field obtained by camera 1(left) and camera 2(right)
The image processing of the stereoscopic PIV measurements basically follows the
same procedures as that of the planar PIV, except that the cross-correlation al-
gorithm applied should be stereo cross-correlation algorithm. In this study, the
interrogation window size set for multi-pass iterations for stereoscopic PIV image
processing is: 75% overlap, 2-pass, from 64 × 64 square-shaped to 16 × 16 round-
shaped.
Correct calibration is essential for PIV velocity measurements. An additional step
in calibration of stereoscopic PIV is referred to as the ’self-calibration’, developed
and validated by Wienke in 2005 [72]. Self-calibration involves computation of
a disparity map based on the real particle images taken by camera 1 and 2 at
the same time (in practice, very short pulse separation, e.g. 1µs). The cross-
correlation of the two images determines whether the calibration plate coincides
with the light sheet. The true position of the light sheet and the mapping func-
tions are then corrected according to the disparity vectors, and thus improve the
calibration of the stereoscopic PIV.
This chapter presents the results of the velocity measurements, acoustic measure-
ments and the deducted far-field acoustics with the implementation of different
formulation of the Curle’s analogy. To start with, section 6.1 presented the veloc-
ity fields measured by the time-resolved planar PIV, from which, the pressure in
the flow field is evaluated by in-plane Poisson solver. On the basis of evaluated
pressure fields, in section 6.2, the far-field acoustic pressure is evaluated at each
location of the four microphones with the distributed formulation of the Curle’s
analogy. The far-filed acoustic pressure is evaluated based on the PIV measure-
ments obtained from both FOV 1 and FOV 2, and the results are compared.
Section 6.3 continues with the far-field acoustic prediction, but with the lumped
formulation of Curle’s analogy. The section starts with the evaluation of aerody-
namic loads. Thereafter, the lumped formulation is applied and the predictions are
compared to the simultaneous microphone measurements both in time series and
in power spectra. Finally, in section 6.4, the 2D acoustic predictions are corrected
with the length of coherence in order to narrow the discrepancies between the PIV
and microphone measurements.
The velocity fields and reconstructed pressure fields presented in this section are
retrieved from the PIV measurement at free stream velocity V∞ = 20m/s and with
the field of view FOV 2 (see Figure 5.7).
Previous study (reference [63], p.74) has shown that for airfoil with Gurney flap,
the velocity fluctuation and the far-field microphone measurement share the same
peak frequency on power spectra. Hence, the Kármán vortex shedding period
Tshed = 0.32ms can be evaluated from the peak frequency 314.9Hz in the power
spectra of microphone measurement (see Figure 6.11, subsection 6.3.2).
The sequence of instantaneous snapshots of the transverse velocity (u-component)
and the streamwise velocity (v-component) in one shedding period (0 to 0.32ms)
are presented respectively in Figure 6.1 and 6.2. The shadow region (as shown in
Figure 5.11), which is caused by refraction of laser light when passing through the
airfoil trailing edge, is filled with velocity values evaluated from nearest neighbour-
interpolation. Such interpolation is justifiable for the following reasons:
• Outside the boundary layer, the flow is essentially irrotational and effectively
inviscid [1], and thus the relation between pressure and velocity is governed
by Bernoulli equation.
• The Gurney flap moves the stagnation point downstream [69] and the spatial
variation of pressure on the suction side is small within the last 3% of the
trailing edge. Consequently, the spatial variation in velocity is expected to
be small.
It can be observed from Figure 6.2 the velocity deficit in the wake region down-
stream the Gurney flap. The alternation of positive and negative transverse veloc-
ity (u-component) observed in Figure 6.1 indicates the existence of large vortical
structures.
Correspondingly, the sequence of vorticity at the same time instants are shown in
Figure 6.3, where red (counter-clockwise vorticity) and blue (clockwise vorticity)
blobs indicate the counter-rotating vortices. The vortices shed at the trailing edge
are conveyed downstream at an average spacing of 5cm with a velocity around
16.7m/s, which is approximately 83% of the free stream velocity.
The in-plane pressure is reconstructed from the PIV measured velocity by Poisson
solver as described in section 3.2 (Equation 3.12 and 3.13). As for the velocity
measured by planar PIV, the spanwise velocity component is not measured and
the flow is assumed as 2D flow. Consequently, the source term is reduced to:
2 2
∂u ∂v ∂u ∂v
fxy = +2 + (6.1)
∂x ∂x ∂y ∂y
As shown in Equation 6.1, the source term is a combination of spatial velocity gra-
dients. However, the velocity gradients retrieved from planar PIV measurement
in the region directly downstream the Gurney flap are unreliable due to the oc-
currence of high intensity 3D motion. Such velocity gradients can cause spurious
oscillations in the solution and may spoil the whole pressure field. To avoid this
problem, the velocity field is smoothed with the Gaussian kernel filter available in
Matlab R , with standard deviation σ = 10 and window size [10 10].
To facilitate the explanation about the boundary conditions, the outer boundary
of the domain is again divided into ’viscous region’ and ’inviscid region’ as in sec-
tion 3.3. The criterion applied in this case is the local value of enstrophy, defined
as:
ε = |→
−
ω |2 (6.2)
In this case, the threshold for viscous region is set as ε > 1.5 × 104 [Hz 2 ].
Dirichlet boundary condition is assigned to the inviscid region of the boundary,
where pressure is calculated by Bernoulli equation:
1 1
p = P∞ + ρV∞2 − ρ(u2 + v 2 ) (6.3)
2 2
Neumann boundary condition is assigned to the airfoil surface and to the viscous
region of the outer boundary. The pressure gradient in the Neumann condition is
defined as in Equation 3.10.
The sequence of instantaneous snapshots of pressure fields corresponding to the
previously shown velocity and vorticity are presented in Figure 6.4. The pressure
downstream the Gurney flap exhibits a coherence pattern as well as a distinguish-
able convection process as that exhibits by the vorticity. The locations of the local
minima of the pressure coincide with the locations of the cores of vortices as shown
in Figure 6.3, while the local maxima indicate the stagnation region caused by flow
impinging the solid surface (upstream the Gurney flap) or the interaction of the
vortices (downstream the Gurney flap).
This section presents the results of acoustic prediction obtained by the imple-
mentation of the distributed formulation of the Curle’s analogy (Equation 2.19),
based on the PIV measurements performed on both FOV 2 (more information
downstream of the Gurney flap) and FOV 1 (more information upstream of the
Gurney flap). Comparison between the two aeroacoustic computations as well as
to the far-field microphone measurements are presented in power spectra of 2.4Hz
resolution.
Table 6.1: Value of t − te for microphones with respect to the sources distributed on the
airfoil surface
It can be observed that the largest difference in retarded time is 0.1ms along the
surface of airfoil. Under the sampling rate of the current study (5000Hz), the time
interval between two snapshots of velocity field is 0.2ms. For the source locations
with a different retarded time from the mean value, the treatment can either be
interpolation of the local time series of pressure to obtain the value in between the
two snapshots, or approximating with the value of the nearest snapshot in time.
With further observation of the mean value of (t − te ), it can be inferred that
source locations with a value of (t − te ) different from the mean value only account
a small portion of airfoil surface. Concerning the simplicity of application, each
microphone is assigned with the respective mean value of (t − te ) for the calcula-
tion of retarded time. The observation of retarded time inversely testifies that the
experimental set-up for the current study satisfies the compact source assumption
as described in section 2.4.
The resulting power spectra of the acoustic prediction based on the PIV mea-
surement on FOV 2 are presented in Figure 6.6 (blue curve). As comparison, the
measurements by corresponding microphones are plotted on the same schematic
(red curve).
The spectra of microphone measurements and the acoustic prediction all feature
a local peak at 315Hz in correspondence to the vortex shedding frequency, with
the drift of peak frequency no more than 10Hz. The first harmonic 630Hz can
be observed but not obvious for microphone measurements, yet it is more remark-
able on the spectra predicted from PIV measurement. The noise level at the peak
frequency is slightly over-estimated by PIV measurement for listener locations at
microphone 1, 2 and 3. The over-estimation of noise level can be attributed to the
2D flow assumption, or full spanwise coherence assumption. However, the over-
estimation in noise level is not observed for microphone 4. Thus, the full spanwise
coherence assumption is only one (though important) of the factors that may influ-
ence the prediction of noise level. In other words, the extent of over-estimation in
noise level should be larger if the full coherence assumption is the only contributing
factor. The reduce in estimated noise level can possibly be ascribed to:
To reduce the influence of the above mentioned two factors, one can either move
the field of view upstream, or avoid the smoothing of velocity field by applying
the lumped formulation of Curle’s analogy. The viability of the two options is
discussed respectively in subsection 6.2.2 and section 6.3.
Less agreement between the PIV prediction and the microphone measurements
is retrieved in broadband components. The microphone spectra feature a rise in
amplitude in frequency range below 100Hz. The low frequency components in
the microphone spectra have been shown to be caused by the wind tunnel oper-
ation noise, concluded from the preliminary background measurements and the
measurements on the airfoil model without Gurney flap (Figure 6.11,green curve).
Self-evidently, the background noise of wind tunnel operation cannot be revealed
by acoustic computation based on velocity measurements.
The sharp decay in magnitude observed at 1445Hz in Figure 6.6 has no physi-
cal meaning. The artificial magnitude reduction is mainly due to the numerical
smoothing of the signal in the calculation of the time derivatives. Comparison
of the effect of the discrete method used for time derivatives was presented by
Lorenzoni (2008)[44], suggesting that the central scheme (which is applied in the
present study) is more effective in magnitude reduction since a wider kernel is used
for the computation of the differences.
The amplitude of computed acoustic spectra does not decay in the high frequency
range as in the spectra of microphone measurements since the high frequency com-
ponents are more susceptible to experimental and numerical noise which can be
attributed to:
• The effect of 3D motion with high intensity directly downstream the Gurney
flap.
As discussed in section 6.2, due to the damping effect of velocity fields smoothing
as treatments for PIV measurements noise and the local 3D motions, the far-
field noise level is unavoidably under-estimated by the distributed formulation of
Curle’s acoustic analogy with the pressure field reconstructed by Poisson solver.
To circumvent the above-mentioned issues, given that the source of sound has been
proved to be compact under the experimental set-up of this study, in this section,
the acoustic prediction results from the implementation of the lumped formulation
of the Curle’s analogy (Equation 2.20) are presented.
The first step towards the application of Equation 2.20 is the evaluation of
aerodynamic loads of the airfoil. The velocity measurement on FOV 2 is chosen
since the outer boundaries of the control volume should preferably be away from
the airfoil surface to avoid the influence of local 3D motions on the reconstructed
pressure.
Lift and drag of the airfoil model are estimated with the momentum balance
equation (Equation 3.15). The pressure in the momentum balance is evaluated by
the method proposed by Kurtulus et al. (2007) as described in section 3.3. In this
case, the threshold for viscous region is set as ε > 1.5 × 104 Hz 2 as in subsection
6.2.1.
The excerpts of lift and drag in time series are plotted respectively in Figure 6.7
and 6.8, with the aerodynamic loads plotted in black and the first, second and
third term in Equation 3.15 plotted respectively in green, blue and red. The plots
also show the mean value of the aerodynamic loads, plotted in magenta.
Periodicity can be observed in the reconstructed aerodynamic loads and the com-
ponents. For lift, the fluctuations are majorly attributed to the volume integral
Figure 6.7: An excerpt of time history of lift and the components of lift evaluated from
PIV measurement:GF6-AOA4-V20-FOV2
of the time derivatives of the velocity and the surface integral of the convection,
while for drag, the major components of fluctuations are the two surface integral
terms. However, it would be arbitrary to neglect any of the three components in
both lift and drag, when looking closely at the order of magnitude and considering
their contribution to the phase of the overall fluctuation.
The resulting power spectra of the aerodynamic loads evaluated from the PIV
measurement on FOV 2 are presented in Figure 6.9. Both the lift and the drag
peak at the frequency of 314.9Hz, which is consistent with the shedding frequency.
Based on the aerodynamic loads evaluated from the PIV measurement, Equation
2.20 is implemented to reconstruct the far-field acoustic pressure. The distance
Figure 6.8: An excerpt of time history of drag and the components of drag evaluated from
PIV measurement:GF6-AOA4-V20-FOV2
from the airfoil trailing edge and the certain microphone is used to calculate the
uniform retarded time for that specified microphone. The values of t − te of the
microphones coincide with the average values listed in Table 6.1.
The excerpts of the far-field acoustic pressure in time series, both the micro-
phone measurements and the reconstruction from PIV measurements based on
full spanwise coherence assumption are shown in Figure 6.10. The two measure-
ments roughly share the same period and phase. However, due to the many stages
in between, such as the smoothing of velocity fields, the computation of time and
spatial derivatives and the acoustic analogy, the maximum coefficients obtained
from the cross-correlation of the far-field pressure measured by microphone and
that evaluated by PIV is barely 0.19.
As expected, the amplitude predicted by PIV measurements is larger than that
measured by the microphones at all four locations of microphones, since in reality
the flow is intrinsically 3D and the spanwise coherence length is shorter than the
span.
Figure 6.9: Power spectra (∆f = 2.4Hz) of lift and drag evaluated from PIV
measurement:GF6-AOA4-V20-FOV2
Figure 6.11 shows the power spectra of the far-field acoustic pressure. The blue
curve and the red curve show respectively the predictions of PIV and the mea-
surements of microphones, and both peak at frequency around 315Hz, which has
been observed in the power spectra of the prediction given by distributed form
of Curle’s analogy (Figure 6.6) and is consistent with the peak frequency of aero-
dynamic loads. The latter consistency is expected, since at the Mach number of
the current study (M a∞ = 0.059), the Doppler effect is negligible, the frequency
reaching the receiver should be equal to the frequency emitted by the source.
The extra green curve in Figure 6.11 offers a comparison of the power spectra
of the clean NACA 0015 (microphone recordings performed at same locations),
which consists of only broadband components. The absence of frequency peak on
the clean NACA 0015 power spectra serves as a side proof that the tonal noise
generated by the model with Gurney flap should be attributed to the presence of
the Gurney flap.
It can be observed from Figure 6.6 and Figure 6.11 that, based on the full span-
wise coherence assumption, where the intensity and phase of the fluctuations are
presumably the same along the span, the PIV measurements over-estimate the
far-field acoustic pressure. In fact, it is reasonable to assume the intensity of the
fluctuation to be uniform along the span except for the region close to the test
section walls where horse shoe vortex develops. However, the assumption of uni-
form phase angle along the span could be the dominating factor leading to the
overprediction as much as 20dB in the frequency range of lumped formulation of
the Curle’s analogy.
Since the planar PIV measurements only offers the intensity and phase information
at the location of the thin light sheet, assumptions should be made on the pressure
fluctuations out of the measuring plane. The coherence function (as introduced in
subsection 3.1.3) provides a good measure of the dependency of the average phase
angle at two locations. An equivalent coherence length can be defined by choos-
ing a critical value for the coherence function. When the value of the coherence
function is above the chosen critical value, the fluctuation can be regarded as in
the same phase, and otherwise, the phase angle is completely independent.
The method is originally proposed by Kato et al.[35] and is proven to work rea-
sonably well by literatures such as [43], [51] and [56].
The measurements by the stereoscopic PIV set-up(see section 5.3) are used for the
assessment of the spanwise length of coherence. The mean velocity, RMS of the
velocity fluctuation and a snapshot of the instantaneous cross-plane velocity fields
of u-component and v-component are shown in Figure 6.12.
The length of coherence is evaluated for both transverse (u) and streamwise (v)
components. The spanwise coherence was calculated along the line parallel to the
airfoil trailing edge, where the maximum RMS of velocity fluctuation locates. In
Figure 6.12, the trailing edge of the airfoil locates at x = 0, while the lines for
calculation of coherence are marked with grey dash dots.
For the calculation of the coherence of fluctuations at two spatial locations, Equa-
tion 3.9 is adapted as:
| hû0 (z0 , f )û0∗ (z, f )i |2
γ 2 (z, f ) = (6.4)
| hû0 (z0 , f )û0∗ (z0 , f )i hû0 (z, f )û0∗ (z, f )i |
In this case, z0 is chosen as the mid-span location.
The equivalent coherence length is dependent on the frequency f . In the current
study, a Gaussian function is fitted for all the frequency components around the
spanwise location z = z0 (for detailed information, see Appendix A). The critical
value of coherence function is chosen as γc = e−2 . The coherence length l is thus
defined as the width of the coherence curve at the critical value.
As reported by literatures such as [35] and [43], the equivalent length indicates
high values only at the frequency associated with large vortex shedding phenom-
ena (in this case, the shedding frequency 314.9Hz) and should drops drastically
below or above the tone. However, for some reasons, the coherence length peaks at
271Hz based on the stereoscopic measurements of this study. For the reason that
the coherence length is later applied for the correction of the planar PIV acoustic
prediction, it is rescaled such that the peak coherence length locates at 314.9Hz.
The length of coherence is plotted with respect to frequency in Figure 6.13.
The occurrence of 271Hz frequency in the coherence length is incongruous when
compared to the shedding frequency measured by planar PIV, however, it is con-
sistent with the stereoscopic PIV measurement, since under the set-up of stereo-
scopic PIV, the shedding frequency is measured as 271Hz. The experimental data
obtained by J.Shah[63], who applied similar experimental set-up, also reports dis-
crepancy in shedding frequencies measured by planar and stereoscopic PIV set-ups.
The phenomenon can be a result of change in test section material in two set-ups.
The discussion on the discrepancy of shedding frequency is dilated in Appendix B.
The values of the coherence length are used for the corrections of the 2D predictions
of the far-field acoustic pressure, which corresponds to adding to each specified
frequency a correction value equals to:
L
−10log10 (6.5)
l(f )
where L is the full span length, and l(f ) is the coherence length at the specified
frequency f .
The corrections of both the coherence length calculated from u-component(green
prisms) and that from v-components(black diamonds) are applied. Figure 6.14 to
6.17 illustrate the results of corrections on the acoustic prediction from distributed
formulation of Curle’s analogy, and figure 6.18 to 6.21 show the results of correc-
tions on the acoustic prediction from lumped formulation of Curle’s analogy.
It can be observed that for both formulations of the Curle’s analogy, the coherence
length correction improves the acoustic prediction in the right tendency, that is
to decrease the amplitude. The corrected predictions by distributed formulation
yield amplitudes lower than the microphone measurements, while for the predic-
tion by lumped formulation, the over-estimation is not completely removed by the
correction.
However, the results of coherence length correction presented in this section is
only illustrative of the viability of this method. Within the scope of current study,
the effectiveness of the coherence length correction cannot be evaluated, since
the above-mentioned shift in shedding frequency observed in the stereoscopic PIV
measurements has rendered the evaluation of coherence length questionable.
Figure 6.14: Power spectra (∆f = 2.4Hz) of acoustic pressure predicted by distributed for-
mulation of Curle’s analogy at the location of Microphone 1: correction based on measured
spanwise coherence of transverse (u) and streamwise (v) velocity
Figure 6.15: Power spectra (∆f = 2.4Hz) of acoustic pressure predicted by distributed for-
mulation of Curle’s analogy at the location of Microphone 2: correction based on measured
spanwise coherence of transverse (u) and streamwise (v) velocity
Figure 6.16: Power spectra (∆f = 2.4Hz) of acoustic pressure predicted by distributed for-
mulation of Curle’s analogy at the location of Microphone 3: correction based on measured
spanwise coherence of transverse (u) and streamwise (v) velocity
Figure 6.17: Power spectra (∆f = 2.4Hz) of acoustic pressure predicted by distributed for-
mulation of Curle’s analogy at the location of Microphone 4: correction based on measured
spanwise coherence of transverse (u) and streamwise (v) velocity
Figure 6.18: Power spectra (∆f = 2.4Hz) of acoustic pressure predicted by lumped for-
mulation of Curle’s analogy at the location of Microphone 1: correction based on measured
spanwise coherence of transverse (u) and streamwise (v) velocity
Figure 6.19: Power spectra (∆f = 2.4Hz) of acoustic pressure predicted by lumped for-
mulation of Curle’s analogy at the location of Microphone 2: correction based on measured
spanwise coherence of transverse (u) and streamwise (v) velocity
Figure 6.20: Power spectra (∆f = 2.4Hz) of acoustic pressure predicted by lumped for-
mulation of Curle’s analogy at the location of Microphone 3: correction based on measured
spanwise coherence of transverse (u) and streamwise (v) velocity
Figure 6.21: Power spectra (∆f = 2.4Hz) of acoustic pressure predicted by lumped for-
mulation of Curle’s analogy at the location of Microphone 4: correction based on measured
spanwise coherence of transverse (u) and streamwise (v) velocity
An aeroacoustic investigation on NACA 0015 with Gurney flap has been performed
based on the measurements of time-resolved PIV. This chapter first highlights the
main findings of this study in section 7.1, following which, the recommendations
for the future investigations are given in section 7.2.
7.1 Conclusions
The main findings from the literature study, the experimental campaigns and the
analysis of the present work, as listed below:
1. The Kármán vortex shedding is the dominating vortex shedding mode of the
Gurney flap. The interaction between the vortices and solid boundaries forms
dipolar sources resided on the solid surface, which is the major source of the
noise generated by trailing edge with Gurney flap. The frequency of the tonal
noise is consistent with the frequency of vortex shedding. The power spectra
of the fluctuations of velocity, aerodynamic loads and far-field pressure did
not indicate the existence of the second shedding mode for this configuration
2. The in-plane pressure was reconstructed from PIV measured velocity with the
in-plane Poisson solver proposed by Gurka et al.[18] for both the measure-
ments on FOV 1 and FOV 2. Coherence pattern and distinguishable con-
vection can be observed downstream the Gurney flap in the reconstructed
pressure field. In addition the locations of the local minima of the recon-
structed pressure are consistent with the locations of vortex cores identified
in the vorticity field. (chapter 6: section 6.1)
4. The noise level predicted by the lumped formulation is higher than that of
the distributed formulation, and at all four microphone locations, the PIV
measurements yielded over-estimation in noise level when compared to the
simultaneous microphone measurements. Such over-estimation was expected
under the assumption of full span coherence. The lumped formulation of the
Curle’s analogy was implemented to avoid the smoothing of velocity field, and
therefore avoiding the concomitant damping of the near-field fluctuations.
(chapter 6: section 6.2, 6.3)
5. The over-estimated noise level under the full spanwise coherence assumption
was corrected with the method of equivalent coherence length, proposed by
Kato et al.[35]. The correction reduced the noise level on the power spectra
for all frequencies. The correction was in the right direction, but the effec-
tiveness cannot yet be estimated in this study for the reason of the observed
inconsistency in shedding frequencies measured by planar PIV and stereo-
scopic PIV. The values of coherence length were rescaled before applied to
7.2 Recommendations
The present study proved the applicability and revealed the limitation of the PIV
techniques in the acoustic investigations of an airfoil with Gurney flap. Hereby
are recommendations on the further researches on the topic:
forces can be computed directly from the velocity fields with the ”flux equa-
tion” proposed by Noca et al. (Equation 3.16), circumventing completely the
pressure reconstruction. Simultaneous PIV and balance (with high frequency
response) measurements is a continuation of the current study, from which
the accuracy of the two approaches can be evaluated.
[2] T Baur and J Köngeter. Piv with high temporal resolution for the deter-
mination of local pressure reductions from coherent turbulence phenomena.
In 3rd international workshop on particle image velocimetry, Santa Barbara,
CA, USA, 1999.
[3] Julius S Bendat and Allan G Piersol. Random data: analysis and measurement
procedures, volume 729. John Wiley & Sons, 2011.
[5] RD Blevins. The effect of sound on vortex shedding from cylinders. Journal
of Fluid Mechanics, 161:217–237, 1985.
[6] YN Chen. Criteria for the cross-flow-induced tube vibrations in tube bank
heat exchangers. In Vibration in nuclear plant. Proceedings of international
conference held at Keswick, UK in May 1978, 1979.
[11] Roeland de Kat and Bas W van Oudheusden. Instantaneous planar pressure
from piv: analytic and experimental test-cases. In Proceedings of the 15th in-
ternational symposium on applications of laser techniques to fluid mechanics,
Lisbon, Portugal, 2010.
[12] Roeland De Kat, Bas W Van Oudheusden, and Fulvio Scarano. Instantaneous
planar pressure field determination around a square-section cylinder based on
time resolved stereo-piv. In Proceedings of the 14th International Symposium
on Applications of Laser Techniques to Fluid Mechanics, Lisbon, Posrtugal,
07-10 July, 2008, paper No. 1259. Calouste Gulbenkian Foundation, 2008.
[14] John M Eargle. Handbook of recording engineering. Springer Science & Busi-
ness Media, 2012.
[18] Roi Gurka, Alex Liberzon, D Hefetz, D Rubinstein, and U Shavit. Computa-
tion of pressure distribution using piv velocity data. In Workshop on particle
image velocimetry, volume 2, 1999.
[19] Richard Haberman. Applied partial differential equations with fourier series
and boundary value problems. AMC, 10:12, 2004.
[21] Colin H Hansen and David A Bies. Engineering noise control. Spon, 1995.
[22] Arne Henning, Kristian Kaepernick, Klaus Ehrenfried, Lars Koop, and An-
dreas Dillmann. Investigation of aeroacoustic noise generation by simultane-
ous particle image velocimetry and microphone measurements. Experiments
in fluids, 45(6):1073–1085, 2008.
[23] Arne Henning, Lars Koop, and Klaus Ehrenfried. Simultaneous particle image
velocimetry and microphone array measurements on a rod-airfoil configura-
tion. AIAA journal, 48(10):2263–2273, 2010.
[24] Arne Henning, Andreas Schröder, Lars Koop, and J Agocs. Causality correla-
tion analysis on a cold jet by means of simultaneous piv and microphone mea-
surements. In 6th AIAA/CEAS Aeroacoustics Conference, Stockholm (Swe-
den), vol AiAA-2010-3753, 2010.
[25] Arne Henning, Björn Wrede, and Reinhard Geisler. Aeroacoustic investi-
gation of a high-lift device by means of synchronized piv and microphone
measurements.
[26] Arne Henning, Björn Wrede, and Andreas Schröder. About the ambiguity
of noise source localization based on the causality correlation technique. In
Proceedings of the 17th international symposium on applications of laser tech-
niques to fluid mechanics, Lisbon, Portugal, 2014.
[28] Michael S Howe. Theory of vortex sound, volume 33. Cambridge University
Press, 2003.
[30] MS Howe. A review of the theory of trailing edge noise. Journal of Sound
and Vibration, 61(3):437–465, 1978.
[33] David Jeffrey, Xin Zhang, and David W Hurst. Aerodynamics of gurney flaps
on a single-element high-lift wing. Journal of Aircraft, 37(2):295–301, 2000.
[35] Chisachi Kato, AKIYOSHI Iida, Yasushi Takano, HAJIME Fujita, and
MASAHIRO Ikegawa. Numerical prediction of aerodynamic noise radiated
from low mach number turbulent wake. AIAA paper, (93-0145), 1993.
[36] Valentina Koschatzky, Jerry Westerweel, and Bendiks Jan Boersma. Com-
parison of two acoustic analogies applied to experimental piv data for cavity
sound emission estimation. In Proceedings of the 16th AIAA/CEAS Aeroa-
coustic Conference, pages 7–9, 2010.
[39] YC Li, JJ Wang, GK Tan, and PF Zhang. Effects of gurney flaps on the
lift enhancement of a cropped nonslender delta wing. Experiments in fluids,
32(1):99–105, 2002.
[40] Robert H Liebeck. Design of subsonic airfoils for high lift. Journal of aircraft,
15(9):547–561, 1978.
[42] Xiaofeng Liu and Joseph Katz. Instantaneous pressure and material accelera-
tion measurements using a four-exposure piv system. Experiments in Fluids,
41(2):227–240, 2006.
[45] David Mackenzie. ICAO: a history of the international civil aviation organi-
zation. University of Toronto Press, 2010.
[48] Roy Myose, Michael Papadakis, and Ismael Heron. Gurney flap experiments
on airfoils, wings, and reflection plane model. Journal of Aircraft, 35(2):206–
211, 1998.
[49] Tomomichi Nakamura, Shigehiko Kaneko, Fumio Inada, Minoru Kato, Ku-
nihiko Ishihara, Takashi Nishihara, and Njuki W Mureithi. Flow-induced vi-
brations: Classifications and lessons from practical experiences. Butterworth-
Heinemann, 2013.
[50] F Noca, D Shiels, and D Jeon. A comparison of methods for evaluating time-
dependent fluid dynamic forces on bodies, using only velocity fields and their
derivatives. Journal of Fluids and Structures, 13(5):551–578, 1999.
[51] RM Oreslli, Julio R Meneghini, and Fabio Saltara. Two and three-dimensional
simulation of sound generated by flow around a circular cylinder. In 15th
AIAA/CEAS Aeroacoustics Conference, AIAA, volume 3270, 2009.
[52] G Philippe, D Guy, and L Jean. Gurney flap scaling for optimum lift-to drag
ratio. AIAA j, 35(12):1888–1890, 1997.
[53] Alan Powell. On the aerodynamic noise of a rigid flat plate moving at zero
incidence. The Journal of the Acoustical Society of America, 31(12):1649–
1653, 1959.
[54] Alan Powell. Theory of vortex sound. The journal of the acoustical society of
America, 36(1):177–195, 1964.
[56] Stefan Pröbsting, Fulvio Scarano, Matteo Bernardini, and Sergio Pirozzoli.
On the estimation of wall pressure coherence using time-resolved tomographic
piv. Experiments in fluids, 54(7):1–15, 2013.
[57] Markus Raffel, Christian E Willert, Jürgen Kompenhans, et al. Particle image
velocimetry: a practical guide. Springer, 2013.
[62] Christophe Schram and Lilla Koloszár. Validation and improvement of air-
frame noise prediction tools. Innovation for Sustainable Aviation in a Global
Environment: Proceedings of the Sixth European Aeronautics Days, Madrid,
30 March-1 April, 2011, page 144, 2012.
[63] J Shah. Aeroacoustics and flow dynamics of an airfoil with a gurney flap using
tr-piv. Master of Science Thesis, Delf university of technology, Aerospace
department, 2015.
[64] DR Troolin, EK Longmire, and WT Lai. Time resolved piv analysis of flow
over a naca 0015 airfoil with gurney flap. Experiments in Fluids, 41(2):241–
254, 2006.
[67] Walter G Vincenti. What engineers know and how they know it, volume 141.
Baltimore: Johns Hopkins University Press, 1990.
[68] Daniele Violato, Peter Moore, and Fulvio Scarano. Lagrangian and eulerian
pressure field evaluation of rod-airfoil flow from time-resolved tomographic
piv. Experiments in fluids, 50(4):1057–1070, 2011.
[69] JJ Wang, YC Li, and K-S Choi. Gurney flap—lift enhancement, mechanisms
and applications. Progress in Aerospace Sciences, 44(1):22–47, 2008.
[70] Meng Wang, Jonathan B Freund, and Sanjiva K Lele. Computational predic-
tion of flow-generated sound. Annu. Rev. Fluid Mech., 38:483–512, 2006.
[71] Jerry Westerweel and Fulvio Scarano. Universal outlier detection for piv data.
Experiments in Fluids, 39(6):1096–1100, 2005.
In this case, y takes the values of coherence coefficients calculated basd on the
measurements of stereoscopic PIV, and x the spanwise locations, with z = 0 the
ing a overdetermined system of linear equations, and thus the linear least square
approach can be applied to determine the value of standard deviation σ.
Take log of both side, then Equation A.2 becomes:
x2
ln(y) = − (A.3)
2σ 2
Xueqing Zhang M.Sc. Thesis
119
1
then let Y = ln(y), a = 2σ 2
, X = x2 , the system can be rewritten as linear
functions:
Yi = aXi + εi (A.4)
minQ(a) (A.5)
a
for Q(a) = ε2i = (Yi − aXi )2 . In other word, a is the solution of the following
P P
Figure A.2: Extrapolated Gaussian curve and the predicted length of coherence
• Energy condition: the energy input to the acoustic mode exceeds the
energy dissipation of this mode in the acoustic field.
Figure B.1: Comparison of the power spectra (∆f = 2.4Hz) of the transverse velocity
(u-component) measured by planar PIV and stereoscopic PIV (GF6-AOA4-V20).
In the experimental campaign, at the same incoming flow rate and with the same
model, the experiments with all-plexiglass test section were audibly noisier than
those with plexiglass-Kevlar test section. A possible explanation can be inferred
that the shift in vortex shedding frequency is the result of the effect of sound
on vortex shedding. Figure B.2 shows the schematic of the feedback mechanism
between the flow and acoustic field.
The influence of a transverse sound wave on vortex shedding frequency of a rigid
circular cylinder in a duct has been explored by Blevins(1985) [5], the findings of
which can be summarized as:
2. Sound applied near the vortex shedding frequency shifts the vortex shed-
ding towards the sound frequency. Maximum shifts of 8% in frequency
were achieved with the sound of higher frequency than the vortex shedding.
Blevins further claimed a tendency for the preferred vortex-shedding fre-
quency to decrease when sound is applied and expected a larger shift by
applying sound with frequency lower than the shedding frequency.
From Figure B.1, it can be observed that the width of the frequency peak
measured by stereoscopic PIV is narrower than the one measured by planar
PIV. And the increase in amplitude could be the result of energy input from the
excitation. In addition, the V-tunnel provides incoming flow with turbulent level
below 0.1%, which is very low. However, the cross-section of the test section used
in the current study is 40 × 40cm2 , correspondingly, the frequency of the first
transversal mode is 425Hz, which is larger than and faraway from the natural
shedding frequency 314.9Hz. And thus, the transversal sound wave is not the
reason and there should be other causes.
0.8m, which corresponds to the frequency of the first longitudinal mode of 215Hz.
This longitudinal frequency may have contributed to the shift in vortex shedding
frequency of the Gurney flap.
A local maxima at 215Hz is distinguishable, but not evident on the power spec-
trum of the background noise of the wind tunnel in operation (Figure B.4). How-
ever, the background noise was measured with the installation of the plexiglass-
Kevlar test section, while the background noise measurement on all-plexiglass test
section is missing. If the longitudinal mode of 215Hz is the mode that shifts the
vortex shedding in the all-plexiglass case, a more discernible peak is expected on
the background noise power spectrum of that case. Another issue that remains to
be clarified is what the energy input is, or excitation of the 215Hz longitudinal
mode.
Figure B.4: Power spectra (∆f = 2.4Hz) of the background noise of V-tunnel with empty
plexiglass-Kevlar test section and incoming flow at 20m/s.
Despite the issues to be clarified in the mechanism of frequency shift in this ap-
pendix, a conclusion can still be reached that all-plexiglass test section, although
favoured for PIV set-up, is not as preferable in the acoustic measurements or vor-
tex shedding related measurements.
Figure B.5 offers an example of the longitudinal acoustic pressure distribution in
a cylindrical tube with different wall conditions at different longitudinal locations.
The sound pressure level applied at the inlet of the tube x = −0.203m is 130dB,
with the flow speed at the center of the tube of M a = 0.335. It can be ob-
served that due to the presence of the acoustic liner in the region [0, 0.406]m, the
sound pressure level exhibits a drop up to 10dB (black diamond in Figure B.5),
corresponding to a reduction by 2/3 in sound pressure. While for the solid wall
region [−0.203, 0]m and [0.406, 0.609]m, the sound pressure level almost remains
the same.
It can be inferred that the application of liner can be an effective way of avoiding
resonance by attenuating the energy input of a certain transverse or longitudinal
acoustic mode. It can also be inferred that the acoustic impedance of Kevlar is
larger than that of the plexiglass. Hence, the replacement of Kevlar, at least on
one of the four sides of the test section, would be desirable for avoiding the effect
of sound on vortex shedding.