Download as pdf or txt
Download as pdf or txt
You are on page 1of 204

Aeroacoustics of Turbulent Flow over the Forward-backward

Facing Step

Author:
Ma, Chung-Hao
Publication Date:
2021
DOI:
https://doi.org/10.26190/unsworks/22694
License:
https://creativecommons.org/licenses/by-nc-nd/3.0/au/
Link to license to see what you are allowed to do with this resource.

Downloaded from http://hdl.handle.net/1959.4/71065 in https://


unsworks.unsw.edu.au on 2022-05-19
Aeroacoustics of Turbulent Flow
over the Forward-backward Facing
Step

Chung-Hao Ma

A thesis presented for the degree of


Master of Philosophy

School of Mechanical and Manufacturing Engineering


University of New South Wales
Australia

June 2021
.

Thesis/Dissertation Sheet

Surname/Family Name : Surname


Given Name/s : Given names
Abbreviation for degree as given in the University calendar : PhD
Faculty : Engineering
School : Mechanical and Manufacturing Engineering
Thesis Title : Put your thesis title here

Abstract 350 words maximum

Forward-Backward facing steps (FBS) are often seen in engineering applications, such as the uneven plate joints on
the surfaces of vehicles or surface irregularities on aircraft. The presence of the step enhances the turbulence levels
in the flow-field resulting in large wall pressure fluctuations, which are considered a source of noise, vibration, and
structural fatigue. In this thesis, a combined experimental and numerical study on the aeroacoustics of low Mach
number, turbulent flow over FBS with different aspect ratios is presented.
The wall pressure measurement results indicate that the largest pressure fluctuations occur slightly downstream of
the step leading edge and the downstream flow field is influenced by this upper corner disturbance. The downstream
flow field modification is more severe for the small aspect ratio step when its length is shorter than the reattachment
length.
Large-Eddy Simulation (LES) was also performed to visualize the flow field around the FBS, which were validated
by the wall pressure measurements results. The turbulent kinetic energy levels were found to be higher near the step
leading edge, indicating that the flow was modified by this sharp corner. Meanwhile, spectral proper orthogonal
decomposition (SPOD) was applied and identify the low-rank behaviour in the flow filed. The velocity-pressure
cross-correlation showed that the wall pressure fluctuations were correlated to the wall normal fluctuating velocity
that was associated with the coherent structures. Also, the instantaneous flow field showed that when the step
length was smaller than the shear layer reattachment length, the large scale vortex formed at the leading edge
would shed downstream directly. This was believed due to the lack of the shear layer flapping motion, which was
found to determine the trajectory of the shedding vortex. Without a reattachment point, resulting in a lower level
wall pressure fluctuation on the top surface.
Aeroacoustic beamforming results display the noise source location around the step leading edge and on the
top surface of the step. Coincidentally, this is close to where the highest root-mean-square (r.m.s.) fluctuating
wall pressure was recorded. The acoustic data combined with the wall pressure data suggest that for FBS noise
generation, the leading edge plays a more crucial role compared with the trailing edge. Meanwhile, the integrated
far-field acoustic spectra show that the levels of FBS noise were lower when the step length was smaller than the
shear layer reattachment length. A new scaling law for FBS sound spectra was proposed to address the influence
of the step AR.

Declaration relating to disposition of project thesis/dissertation.

I hereby grant to the University of New South Wales or its agents a non-exclusive licence to archive and to make available (including
to members of the public) my thesis or dissertation in whole or in part in the University libraries in all forms of media, now or here
after known. I acknowledge that I retain all intellectual property rights which subsist in my thesis or dissertation, such as copyright
and patent rights, subject to applicable law. I also retain the right to use all or part of this thesis or dissertation in future works (such
as articles or books).

29/08/2021
Signature Date

The University recognises that there may be exceptional circumstances requiring restrictions on copying or conditions on use. Requests
for restriction for a period of up to 2 years must be made in writing. Requests for a longer period of restriction may be considered in
exceptional circumstances and require the approval of the Dean of Graduate Research
ORIGINALITY STATEMENT

‘I hereby declare that this submission is my own work and to the best of my knowledge
it contains no materials previously published or written by another person, or substantial
proportions of material which have been accepted for the award of any other degree or
diploma at UNSW or any other educational institution, except where due acknowledgment
is made in the thesis. Any contribution made to the research by others, with whom I have
worked at UNSW or elsewhere, is explicitly acknowledged in the thesis. I also declare that
the intellectual content of this thesis is the product of my own work, except to the extent
that assistance from others in the project’s design and conception or in style, presentation
and linguistic expression is acknowledged.’

Signed Date 29/08/2021


COPYRIGHT STATEMENT

‘I hereby grant the University of New South Wales or its agents a non-exclusive licence
to archive and to make available (including to members of the public) my thesis or
dissertation in whole or part in the University libraries in all forms of media, now or here
after known. I acknowledge that I retain all intellectual property rights which subsist in
my thesis or dissertation, such as copyright and patent rights, subject to applicable law.
I also retain the right to use all or part of my thesis or dissertation in future works (such
as articles or books).’

‘For any substantial portions of copyright material used in this thesis, written permission
for use has been obtained, or the copyright material is removed from the final public
version of the thesis.’

Signed ……………………………………………...........................

06/09/2021
Date ……………………………………………..............................

AUTHENTICITY STATEMENT
‘I certify that the Library deposit digital copy is a direct equivalent of the final officially
approved version of my thesis.’

Signed ……………………………………………...........................

06/09/2021
Date ……………………………………………..............................
INCLUSION OF PUBLICATIONS STATEMENT

UNSW is supportive of candidates publishing their research results during their candidature as
detailed in the UNSW Thesis Examination Procedure.

Publications can be used in their thesis in lieu of a Chapter if:

• The student contributed greater than 50% of the content in the publication and is the
‘primary author’, ie. the student was responsible primarily for the planning, execution,
and preparation of the work n the publication.

• The student has approval to include the publication in their thesis in lieu of a Chapter
from their supervisor and Postgraduate Coordinator.

• The publication is not subject to any obligations and contractual agreements with a third
party that would constrain its inclusion in the thesis.

Please indicate whether this thesis contains published material or not.

 This thesis contains no publications, either published or submitted for


publication.

 Some of the work described in this thesis has been published and it has been
documented in the relevant Chapters with acknowledgement.

 This thesis has publications (either published or submitted for publication)


incorporated into it in lieu of a chapter and the details are presented below.

CANDIDATE’S DECLARATION
I delcare that:

• I have complied with the Thesis Examination Procedure.

• Where I have used a publication in lieu of a Chapter, the listed publications(s) below
meet(s) the requirements to be included in the thesis.

Name Signature Date


Chung-Hao Ma
29/08/2021
Given names Surname
Abstract

Forward-Backward facing steps (FBS) are often seen in engineering applications,


such as the uneven plate joints on the surfaces of vehicles or surface irregularities
on aircraft. The presence of the step enhances the turbulence levels in the flow-
field resulting in large wall pressure fluctuations, which are considered a source of
noise, vibration, and structural fatigue. In this thesis, a combined experimental and
numerical study on the aeroacoustics of low Mach number, turbulent flow over FBS
with different aspect ratios is presented.
The wall pressure measurement results indicate that the largest pressure fluc-
tuations occur slightly downstream of the step leading edge and the downstream
flow field is influenced by this upper corner disturbance. The downstream flow field
modification is more severe for the small aspect ratio step when its length is shorter
than the reattachment length.
Large-Eddy Simulation (LES) was also performed to visualize the flow field
around the FBS, which was validated by the wall pressure measurement results.
The turbulent kinetic energy levels were found to be higher near the step leading
edge, indicating that the flow was modified by this sharp corner. Meanwhile, spec-
tral proper orthogonal decomposition (SPOD) was applied and identify the low-rank
behaviour in the flow field. The velocity-pressure cross-correlation showed that the
wall pressure fluctuations were correlated to the wall normal fluctuating velocity
that was associated with the coherent structures. Also, the instantaneous flow field
showed that when the step length was smaller than the shear layer reattachment
length, the large scale vortex formed at the leading edge would shed downstream
directly. This was believed due to the lack of the shear layer flapping motion, which
was found to determine the trajectory of the shedding vortex. Without a reattach-
ment point, resulting in a lower level wall pressure fluctuation on the top surface.
Aeroacoustic beamforming results display the noise source location around the
step leading edge and on the top surface of the step. Coincidentally, this is close to
where the highest root-mean-square (r.m.s.) fluctuating wall pressure was recorded.

i
Abstract ii

The acoustic data combined with the wall pressure data suggest that for FBS noise
generation, the leading edge plays a more crucial role compared with the trailing
edge. Meanwhile, the integrated far-field acoustic spectra show that the levels of FBS
noise were lower when the step length was smaller than the shear layer reattachment
length. A new scaling law for FBS sound spectra was proposed to address the
influence of the step AR.
Acknowledgements

Firstly, I would like to sincerely express my gratitude to my supervisors, Profes-


sor Con Doolan, Dr. Danielle Moreau, and Dr. Manuj Awasthi, for their magnificent
support and guidance. Huge thank you to them for setting a role model as a re-
searcher and always being patient and encouraging to me. It was a great pleasure
to grow and develop under their supervision.
I would secondly like to thank all the people in UNSW Flow Noise Group who
have helped me with my project over the past two years. I would like to thank Dr.
Yendrew Yauwenas for always giving me valuable tips on CFD simulations. Thank
you, Dr. Chaoyang Jiang, Jiawei Tan, and Dr. Angus Wills, for the precious advice
on the experimental setup and data analysis. I would also like to thank Dr. Jeoffrey
Fischer for his help in acoustic measurements and beamforming techniques. A big
thank you to Roman Kisler, Rowena Dixon, Dr. Charitha de Silva, Tingyi Zhang,
Yuchen Ding, Prateek Bahl, Ziao Zhang, and Sean McCreton for always providing
priceless opinions in insightful discussions. I would also like to thank Omear Saeed
for his incredible work in the Aero-Lab that helps me run my experiments smoothly.
Finally, I would like to thank my family. Huge thank you to my parents, Yun-
Hsien Chen and Yu-Long Ma, for the endless love and support when I was far away
from home. This thesis could not have been possible without them.

iii
Contents

Abstract i

Acknowledgements iii

Contents iv

List of Figures vii

List of Tables xv

Nomenclature xvi

1 Introduction 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Aims . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Thesis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2 Literature Review 6
2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2 Noise Generated by Low Mach Number Flow . . . . . . . . . . . . . . 7
2.3 Step Induced Flow Field and Noise . . . . . . . . . . . . . . . . . . . 8
2.3.1 Forward-facing Step . . . . . . . . . . . . . . . . . . . . . . . 8
2.3.2 Backward-facing Step . . . . . . . . . . . . . . . . . . . . . . . 14
2.3.3 Forward-backward Facing Step . . . . . . . . . . . . . . . . . 16
2.4 Summary and Research Gaps . . . . . . . . . . . . . . . . . . . . . . 18

3 Methodology 20
3.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.2 Far-field Noise and Wall Pressure Measurements . . . . . . . . . . . . 21
3.2.1 UNSW Anechoic Wind Tunnel . . . . . . . . . . . . . . . . . . 21

iv
Contents v

3.2.2 Test Parameters and Forward-backward Facing Step Configu-


ration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.2.3 Far-field Acoustic Measurements . . . . . . . . . . . . . . . . . 26
3.2.4 Wall Pressure Measurement . . . . . . . . . . . . . . . . . . . 29
3.3 Oil Flow Visualisation Technique . . . . . . . . . . . . . . . . . . . . 39
3.3.1 UNSW Educational Wind Tunnel . . . . . . . . . . . . . . . . 39
3.3.2 Oil Flow Visualization . . . . . . . . . . . . . . . . . . . . . . 39
3.4 Numerical Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.4.1 Flow Domain and Turbulence Model . . . . . . . . . . . . . . 42
3.4.2 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . 43
3.4.3 Numerical Procedure and Computational Grid . . . . . . . . . 44
3.4.4 Grid Convergence Study . . . . . . . . . . . . . . . . . . . . . 45
3.5 Data Processing and Analysis Methods . . . . . . . . . . . . . . . . . 47
3.5.1 Two-point Correlation/Coherence . . . . . . . . . . . . . . . . 47
3.5.2 Modal Analysis Methods . . . . . . . . . . . . . . . . . . . . . 48
3.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

4 Oil Flow Visualization and Wall Pressure Measurements 52


4.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.2 Oil Flow Visualization Results . . . . . . . . . . . . . . . . . . . . . . 53
4.3 Mean Wall pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.3.1 Smooth Wall . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.3.2 Forward-backward Facing Step Mean Wall Pressure . . . . . . 59
4.4 Fluctuating Wall Pressure . . . . . . . . . . . . . . . . . . . . . . . . 64
4.4.1 R.M.S Wall Pressure . . . . . . . . . . . . . . . . . . . . . . . 64
4.4.2 Wall Pressure Spectrum . . . . . . . . . . . . . . . . . . . . . 70
4.4.3 Two-point Correlation and Coherence Analysis . . . . . . . . . 80
4.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

5 CFD Results 98
5.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
5.2 LES Comparison with Mean and Fluctuating Wall Pressure Measure-
ments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
5.3 Mean Flow Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
5.4 Turbulence Statistics . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
5.5 Instantaneous Vorticity Field . . . . . . . . . . . . . . . . . . . . . . 111
5.6 Spectral Proper Orthogonal Decomposition . . . . . . . . . . . . . . . 113
5.6.1 SPOD Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.6.2 Pressure-Velocity Coherence . . . . . . . . . . . . . . . . . . . 117
5.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
Contents vi

6 Forward-backward Facing Acoustic Characteristics 123


6.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
6.2 Single Microphone Data . . . . . . . . . . . . . . . . . . . . . . . . . 125
6.2.1 Background Noise and Step Induced Noise . . . . . . . . . . . 125
6.2.2 Comparison with Analytical Model . . . . . . . . . . . . . . . 132
6.2.3 Directivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
6.3 Acoustic Beamforming . . . . . . . . . . . . . . . . . . . . . . . . . . 137
6.3.1 Acoustic Beamforming Maps . . . . . . . . . . . . . . . . . . . 137
6.3.2 Integrated Spectra . . . . . . . . . . . . . . . . . . . . . . . . 141
6.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

7 Conclusions 148
7.1 Summary of Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
7.2 Main Findings and Achievement . . . . . . . . . . . . . . . . . . . . . 149
7.3 Recommendations for Future Work . . . . . . . . . . . . . . . . . . . 151

References 153

Appendices 162

A Microphone Details for Wall Pressure Measurements 163


A.1 Undisturbed Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
A.1.1 Streamwise Measurement Locations . . . . . . . . . . . . . . . 164
A.1.2 Spanwise Measurement Locations . . . . . . . . . . . . . . . . 165
A.2 Step1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
A.2.1 Streamwise Measurement locations . . . . . . . . . . . . . . . 167
A.2.2 Spanwise Measurement Locations . . . . . . . . . . . . . . . . 168
A.3 Step2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
A.3.1 Streamwise Measurement Locations . . . . . . . . . . . . . . . 170
A.3.2 Spanwise Measurement Locations . . . . . . . . . . . . . . . . 171
A.4 Step4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
A.4.1 Streamwise Measurement Locations . . . . . . . . . . . . . . . 173
A.4.2 Spanwise Measurement Locations . . . . . . . . . . . . . . . . 175
A.5 Step8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
A.5.1 Streamwise Measurement Locations . . . . . . . . . . . . . . . 177
A.5.2 Spanwise Measurement Locations . . . . . . . . . . . . . . . . 179
List of Figures

1.1 Sketch of FBS flow field. . . . . . . . . . . . . . . . . . . . . . . . . . 2

2.1 Sketch of FFS flow field. . . . . . . . . . . . . . . . . . . . . . . . . . 9


2.2 Schematic representation of a period of vortex shedding. Taken from
Kiya & Sasaki [42]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.3 Step configurations investigated in [66]. Taken from Catlett et al. [66] 13
2.4 Sketch of BFS flow field. . . . . . . . . . . . . . . . . . . . . . . . . . 14

3.1 Schematic diagram of the UNSW anechoic wind tunnel AWT. . . . . 22


3.2 Rectangular steps with 4 different aspect ratios considered in the wall
pressure and acoustic measurements. . . . . . . . . . . . . . . . . . . 22
3.3 Schematic diagram of experimental setup. Red dashed line indicates
the location of the step leading edge (x = 0). . . . . . . . . . . . . . . 24
3.4 Sideview of the experimental setup. . . . . . . . . . . . . . . . . . . . 24
3.5 Boundary layer profile measured at (x, z) = (0, 0) without step for
flow speed = 20, 30, 40 m/s. Solid line, u+ = (1/0.41) ln (y + ) + 5;
Blue ◦,U∞ = 20 m/s; Red 4,U∞ = 30 m/s ;Yellow ,U∞ = 40 m/s. 25
2
3.6 Turbulence stress u02 /U∞ measured at (x, y) = (0, 0) without step for
U∞ = 20, 30, 40 m/s. . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.7 64-microphone phased array. . . . . . . . . . . . . . . . . . . . . . . . 27
3.8 Cross-sectional view of the pressure tap. . . . . . . . . . . . . . . . . 30
3.9 Wall pressure measurement locations for flat plate cases. Red dashed
line indicates step leading edge location. . . . . . . . . . . . . . . . . 30
3.10 Wall pressure measurement locations for AR = 1 step. Red dashed
line indicates step leading edge, black dashed line indicates step trail-
ing edge. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.11 Pressure tap on the vertical faces of the step. . . . . . . . . . . . . . . 32
3.12 Mean wall pressure measurement setup. . . . . . . . . . . . . . . . . . 33
3.13 Remote microphone setup. . . . . . . . . . . . . . . . . . . . . . . . . 34

vii
List of Figures viii

3.14 Sketches of calibration methods in two-step remote microphone cali-


bration [7] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.15 Calibration curves for the pressure tap at (x, y) = (−40, 0) at a free
stream velocity of 20 m/s. (a) Transfer function yielded from both
Cavity calibration and Free-field calibration. (b) Combined calibra-
tion curve, cutoff frequency = 3192 Hz. . . . . . . . . . . . . . . . . . 36
3.16 Calibrated and Uncalibrated pressure spectrum at (x, z) = (−40, 0)
at a free stream velocity of 20 m/s . . . . . . . . . . . . . . . . . . . 37
3.17 Flat plate unsteady pressure spectra for U∞ = 20 m/s at different
streamwise locations compared with Goody’s model [30]. . . . . . . . 38
3.18 Schematic diagram of UNSW EWT and oil flow visualization setup. . 39
3.19 Oil flow visualization result for AR = 4 step at U∞ = 35 m/s. Flow is
coming from left to the right. Yellow solid lines indicate the boundary
of the separation/reattachment regions. Yellow dashed lines indicate
the center of the reattachment regions. Red dashed lines indicate the
step leading edge and trailing edge. . . . . . . . . . . . . . . . . . . . 41
3.20 Schematic diagram of the computational domain and the boundary
condition employed in the simulation. . . . . . . . . . . . . . . . . . . 44
3.21 Computational domain meshing. Side view from the spanwise direc-
tion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

4.1 Separation length xs and reattachment lengths xr1 & xr2 . . . . . . . . 53


4.2 Upstream separation length xs of FBS compared with past FFS studies. 54
4.3 Reattachment length on the top surface (xr1 ) of FBS compared with
past FFS studies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
4.4 Reattachment length downstream of the FBS (xr2 ) compared with
past BFS studies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.5 Reattachment length downstream of FBS compared with past BFS
studies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.6 Streamwise mean wall pressure for undisturbed flow (no step) at z =
0. Blue 3, U∞ =20 m/s; Red ◦, U∞ = 30 m/s; Yellow 4, U∞ = 40
m/s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.7 Spanwise mean wall pressure for undisturbed flow (no step) at two
streamwise locations. (a) At x = −0.5h. (b) At x = 6.4h. Blue 3,
U∞ = 20 m/s; Red ◦, U∞ = 30 m/s; Yellow 4, U∞ = 40 m/s. . . . . 59
4.8 Mean pressure coefficient for four steps configuration at U∞ = 20
m/s. (a) Step1. (b) Step2. (c) Step4. (d) Step8. The red and black
vertical dashed lines represent step leading edge and trailing edge,
respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
List of Figures ix

4.9 Mean pressure coefficient near the step leading edge. Black dashed
line indicates the step leading edge location. (a) Current data com-
pared with past studies. (b) An enlarged view of the leading edge
area. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.10 Mean pressure coefficient downstream of step trailing edge compared
with past backward-facing step studies. . . . . . . . . . . . . . . . . . 63
4.11 Root mean square wall pressure fluctuations for four steps config-
urations at U∞ = 20m/s. The red and black vertical dashed line
represents the step leading edge and trailing edge, respectively. . . . . 65
4.12 Root mean squared wall pressure fluctuations of current data near
the step leading edge compared with past forward-facing step studies. 67
4.13 Dependence of the excess pressure fluctuation intensity, relative to
0
the undisturbed root mean square Cp∞ on x/h. . . . . . . . . . . . . 67
4.14 Root mean squared wall pressure fluctuations of current data near
the step trailing edge compared with past backward-facing step studies. 69
4.15 R.m.s. wall pressure fluctuations at three different Reh . (a) Step1.
(b) Step2. (c) Step4. (d) Step8. The red and black vertical dashed
line represents the step leading edge and trailing edge, respectively.
Blue ◦, Reh = 14, 078 (U∞ =20 m/s) ; Red , Reh = 21, 116 (U∞
=30 m/s); Yellow 3, Reh = 28, 155 (U∞ =40 m/s). . . . . . . . . . . 70
4.16 Upstream wall pressure spectra for four step cases and undisturbed
boundary layer at U∞ = 20 m/s. (a) x = −15h. (b) x = −4h.
(c)x = −2h. (d) x = −0.5h. The black solid line shows the slope of
-5/2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.17 Wall pressure spectra for four step cases on the step top surface and
undisturbed boundary layer at U∞ = 20 m/s. (a) x = 0.5h. (b)
x = 1.5h. (c)x = 2.5h. (d) x = 3.5h. Noted that some of the spectra
are not plotted as the considered location is no longer on the step top
surface. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.18 Wall pressure spectra on the top surface of Step8 at different stream-
wise locations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.19 Downstream wall pressure spectra for four step cases and undisturbed
boundary layer at U∞ = 20 m/s. (a) x ' 0.1xr2 . (b) x ' 0.5xr2 .
(c)x ' xr2 . (d) x ' 3xr2 . The distance is calculated from the step
trailing edge location. . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.20 Streamwise wall pressure fluctuations for four step cases. The vertical
black solid lines are the step edges and the black dashed lines are the
estimated reattachment point. The contour is plotted in dB (ref.
20µPa) (a) Step1. (b)Step2. (c)Step4. (d) Step8. . . . . . . . . . . . 78
List of Figures x

4.21 Wall pressure spectra compared with past studies at two streamwise
locations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
4.22 Space-time correlation coefficient Cpp for four step cases on the top
surface. From left: first column, Step1; second column, Step2; third
column, Step4; fourth column, Step8. From top: first row, x =
2.5h; second row, x = 1.5h; third row, x = 0.5h; fourth row, x =
−0.5h. The black and blue solid lines represent the step leading edge
and trailing edge, respectively. The black dashed lines indicate the
estimated reattachment points. . . . . . . . . . . . . . . . . . . . . . 82
4.23 Space-time correlation coefficient Cpp for four step cases downstream
of the step. From left: first column, Step1; second column, Step2;
third column, Step4; fourth column, Step8. From top: first row,
x ' xr2 ; second row, x ' 0.5xr2 ; third row, x = 0.8h; fourth row,
x = −0.5h. The black and blue solid lines represent the step leading
edge and trailing edge, respectively. The black dotted lines indicate
the estimated reattachment points. . . . . . . . . . . . . . . . . . . . 84
4.24 Cross correlation of the wall pressure fluctuations of Step8 at U∞ =
20 m/s. Reference point x = 0.5h. . . . . . . . . . . . . . . . . . . . 85
4.25 Time averaged convection velocity on the top surface of the step. (a)
Step4. (8) Step8. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
4.26 Time average convection velocity downstream of the step. (a) Step1.
(b) Step2. (c) Step4. (d) Step8. . . . . . . . . . . . . . . . . . . . . . 87
4.27 Spanwise coherence (squared) γ 2 for undisturbed boundary layer at
three different flow speeds. (a) U∞ = 20 m/s. (b) U∞ = 30 m/s. (c)
U∞ = 40 m/s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.28 Spanwise coherence γ 2 for Step8 at x = 2h and the curves. . . . . . . 89
4.29 Spanwise coherence length as the function of frequency for undis-
turbed boundary layer at three flow speeds. . . . . . . . . . . . . . . 90
4.30 Spanwise coherence (squared) γ 2 upstream of the step at x = −0.5h.
(a) Step1. (b) Step2. (c) Step4. (d) Step8. . . . . . . . . . . . . . . 91
4.31 Spanwise coherence length as the function of frequency for U∞ = 20
m/s at x = −0.5h. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.32 Spanwise coherence (squared) γ 2 downstream of the step at x = 2h
from trailing edge. (a) Step1. (b) Step2. (c) Step4. (d) Step8. . . . . 93
4.33 Spanwise coherence length as the function of frequency for U∞ = 20
m/s at x = 2h from trailing edge. . . . . . . . . . . . . . . . . . . . . 93
4.34 Spanwise coherence (squared) γ 2 on the top surface of the step at half
step length x = 0.5l. (a) Step1. (b) Step2. (c) Step4. (d) Step8. . . . 94
List of Figures xi

4.35 Spanwise coherence length as the function of frequency for U∞ = 20


m/s at x = 0.5l from the leading edge (on the top). . . . . . . . . . . 95

5.1 Comparison of LES and measured mean pressure coefficient Cp distri-


bution. Red solid lines, LES; Blue dots, wall pressure measurements.
Red and black dashed lines represent the step leading edge and trail-
ing edge, respectively. Black dotted line in (b) indicates the step
trailing edge in CFD case. (a) LES AR = 1 compared with Step1.
(b) LES AR = 2.4 compared with Step2. (c) LES AR = 4 compared
with Step4. (d) LES AR = 8 compared with Step8. . . . . . . . . . 100
5.2 Comparison of LES and measured r.m.s. wall pressure fluctuations
distribution. Red solid lines, LES; Blue dots, wall pressure measure-
ments. Red and black dashed lines represent the step leading edge
and trailing edge, respectively. Black dotted line in (b) indicates the
step trailing edge in CFD case. (a) LES AR = 1 compared with
Step1. (b) LES AR = 2.4 compared with Step2. (c) LES AR = 4
compared with Step4. (d) LES AR = 8 compared with Step8. . . . . 101
5.3 Mean streamwise velocity contours (U/U∞ ). Black solid lines repre-
sent the mean streamlines. Red triangle and cross mark the locations
of separation points and reattachment points, respectively. The color
bar represent the level of U/U∞ . (a) AR = 1. (b) AR = 2.4. (c)
AR = 4. (d) AR = 8. . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
5.4 Mean ∂u/∂y contours. (a) AR = 1. (b) AR = 2.4. (c) AR = 4. (d)
AR = 8. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
2
5.5 Turbulent kinetic energy (u02 +v 02 +w02 )/U∞ . Black solid lines indicate
∂u/∂y with ten contour levels. (a) AR = 1. (b) AR = 2.4. (c)
AR = 4. (d) AR = 8. . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
5.6 Reynolds stresses component u0 u0 . Black dotted lines indicate ∂u/∂y
with ten contour levels. (a) AR = 1. (b) AR = 2.4. (c) AR = 4. (d)
AR = 8. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5.7 Reynolds stresses component v 0 v 0 . Black dotted lines indicate ∂u/∂y
with ten contour levels. (a) AR = 1. (b) AR = 2.4. (c) AR = 4. (d)
AR = 8. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
5.8 Reynolds stresses component u0 v 0 . (a) AR = 1. (b) AR = 2.4. (c)
AR = 4. (d) AR = 8. . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
5.9 Instantaneous vorticity field for AR = 1 step. The time increment
t∗ = ∆th
U∞
, where ∆t is the sampling time 4.8 × 10−6 s. . . . . . . . . . 112
5.10 Instantaneous vorticity field for AR = 8 step. The time increment
t∗ = ∆th
U∞
, where ∆t is the sampling time 4.8 × 10−6 s. . . . . . . . . . 112
List of Figures xii

5.11 SPOD eigenspectra and first two SPOD modes at three different fre-
quencies for AR = 1. On the left : SPOD eigenspectra for first 12
modes. (b, d, f) First SPOD modes. Black solid lines represent ∂u/∂y
contours; (c, e, g) Second SPOD modes. Black solid lines represent
mean streamlines. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.12 SPOD eigenspectra and first two SPOD modes at three different fre-
quencies for AR = 2.4. On the left : SPOD eigenspectra for first 12
modes. (b, d, f) First SPOD modes. Black solid lines represent ∂u/∂y
contours; (c, e, g) Second SPOD modes. Black solid lines represent
mean streamlines. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.13 SPOD eigenspectra and first two SPOD modes at three different fre-
quencies for AR = 4. On the left : SPOD eigenspectra for first 12
modes. (b, d, f) First SPOD modes. Black solid lines represent ∂u/∂y
contours; (c, e, g) Second SPOD modes. Black solid lines represent
mean streamlines. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
5.14 SPOD eigenspectra and first two SPOD modes at three different fre-
quencies for AR = 8. On the left : SPOD eigenspectra for first 12
modes. (b, d, f) First SPOD modes. Black solid lines represent ∂u/∂y
contours; (c, e, g) Second SPOD modes. Black solid lines represent
mean streamlines. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
5.15 AR = 1 step coherence coefficient Cpv maps for Stδ = 0.116 (top
row) and Stδ = 0.291 (second row). From left: First column, SPOD
mode 1 at the corresponding frequency; Second column, Cpv maps
at x/h ≈ 0.5; Third column, Cpv maps at x/h ≈ 0.95. Black lines:
∂u/∂y with 10 contour levels. . . . . . . . . . . . . . . . . . . . . . . 119
5.16 AR = 2.4 step coherence coefficient Cpv maps for Stδ = 0.233 (top
row) and Stδ = 0.349 (second row). From left: First column, SPOD
mode 1 at the corresponding frequency; Second column, Cpv maps at
x/h ≈ 0.5; Third column, Cpv maps at x/h ≈ 2. Black lines: ∂u/∂y
with 10 contour levels. . . . . . . . . . . . . . . . . . . . . . . . . . . 119
5.17 AR = 4 step coherence coefficient Cpv maps for Stδ = 0.233 (top
row) and Stδ = 0.407 (second row). From left: First column, SPOD
mode 1 at the corresponding frequency; Second column, Cpv maps at
x/h ≈ 0.5; Third column, Cpv maps at x/h ≈ 2. Black lines: ∂u/∂y
with 10 contour levels. . . . . . . . . . . . . . . . . . . . . . . . . . . 120
List of Figures xiii

5.18 AR = 8 step coherence coefficient Cpv maps for Stδ = 0.174 (top
row) and Stδ = 0.407 (second row). From left: First column, SPOD
mode 1 at the corresponding frequency; Second column, Cpv maps at
x/h ≈ 0.5; Third column, Cpv maps at x/h ≈ 2.5. Black lines: ∂u/∂y
with 10 contour levels. . . . . . . . . . . . . . . . . . . . . . . . . . . 120

6.1 Sound spectra for background noise (undisturbed flow) at three dif-
ferent flow speeds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
6.2 Narrow band sound spectra for different AR steps compared with
undisturbed flow at U∞ = 40 m/s. (a) Step1. (b) Step2. (c) Step4.
(d) Step8. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
6.3 SNR for different AR steps (1/12th octave band) at U∞ = 40m/s.
(a) Step1.(b) Step2.(c) Step4.(d) Step8. . . . . . . . . . . . . . . . . 127
6.4 1/12th octave band sound spectra at θ = 83.4◦ . (a) Step1.(b) Step2.(c)
Step4.(d) Step8. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
6.5 Scaled 1/12th octave band sound spectra. (a) Step1. (b) Step2. (c)
Step4. (d) Step8. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
6.6 Comparison of different scaling law. (a) Unscaled 1/12th octave
band sound spectra at U∞ =30m/s. (b) Scaled sound spectra at
U∞ =30m/s using Eq.6.1. (c) Scaled sound spectra at U∞ =30m/s
using Eq.6.2. (d) Scaled sound spectra for all cases using Eq.6.2. . . 132
6.7 Configuration of the forward-facing step. Taken from Awasthi et
al.[9]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
6.8 1/12th octave band sound spectra of the FBS compared with the FFS
analytical model [9]. (a) Step1. (b) Step2. (c) Step4. (d) Step8. . . . 134
6.9 1/12th octave band sound spectra of the FBS compared with the FFS
analytical model [9] . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
6.10 Directivity measurement setup . . . . . . . . . . . . . . . . . . . . . . 135
6.11 Directivity map at 1728 Hz. . . . . . . . . . . . . . . . . . . . . . . . 136
6.12 Directivity map at 7280 Hz. . . . . . . . . . . . . . . . . . . . . . . . 137
6.13 Acoustic beamforming maps for background noise, pre-background-
subtraction, and background subtraction at 5040 Hz. The results are
given for Step8 at Uinf ty = 40 m/s and the level is in dB.. Green
rectangle indicates the location of the step. . . . . . . . . . . . . . . . 137
6.14 Acoustic beamforming maps at 2670 Hz at U∞ =40 m/s. Flow is
form the left and the level is in dB. Green rectangular indicate the
step location. (a) Step1. (b) Step2. (c) Step4. (d) Step8. . . . . . . 139
List of Figures xiv

6.15 Acoustic beamforming maps at 6350 Hz at U∞ =40 m/s. Flow is


form the left and the level is in dB. Green rectangular indicate the
step location. (a) Step1. (b) Step2. (c) Step4. (d) Step8. . . . . . . . 140
6.16 Acoustic beamforming maps at 10079 Hz at U∞ =40 m/s. Flow is
form the left and the level is in dB. Green rectangular indicate the
step location. (a) Step1. (b) Step2. (c) Step4. (d) Step8. . . . . . . . 141
6.17 Sound maps integration region at different frequencies. Frequencies
below 4 kHz using (a), and frequencies above 4k Hz using (b). . . . . 142
6.18 1/12th octave band integrated spectra (Pa2 /Hz) for Step8 at U∞ =40
m/s. (a) Influences of integration region. (b) Comparison integrated
spectra of subtracted, unsubtracted, and background noise. . . . . . . 143
6.19 Integrated spectra (Pa2 /Hz) at different flow speeds. (a) U∞ =20
m/s. (b) U∞ =30 m/s. (c) U∞ =40 m/s. . . . . . . . . . . . . . . . . 144
6.20 Scaled integrated spectra. (a) U∞ =40 m/s. (b) All testing cases. . . 145

A.1 Wall pressure measurement locations for undisturbed flow. Red dashed
line indicates step leading edge location (x = 0). . . . . . . . . . . . . 166
A.2 Wall pressure measurement locations for Step1. Red dashed line in-
dicates step leading edge, black dashed line indicates step trailing
edge. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
A.3 Wall pressure measurement locations for Step2. Red dashed line in-
dicates step leading edge, black dashed line indicates step trailing
edge. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
A.4 Wall pressure measurement locations for Step4. Red dashed line in-
dicates step leading edge, black dashed line indicates step trailing
edge. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
A.5 Wall pressure measurement locations for Step8. Red dashed line in-
dicates step leading edge, black dashed line indicates step trailing
edge. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
List of Tables

3.1 Test cases and step parameters . . . . . . . . . . . . . . . . . . . . . 23


3.2 Undisturbed boundary layer information . . . . . . . . . . . . . . . . 26
3.3 UNSW EWT boundary layer information. . . . . . . . . . . . . . . . 40
3.4 Grid Convergence Index for AR = 4 step . . . . . . . . . . . . . . . . 46

4.1 Oil flow visualization results . . . . . . . . . . . . . . . . . . . . . . . 53


4.2 Estimated downstream reattachment lengths xr1 and xr2 at U∞ =20
m/s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

5.1 Separation and reattachment lengths predicted by LES and compared


with experiment results . . . . . . . . . . . . . . . . . . . . . . . . . . 104

A.1 Undisturbed flow streamwise microphone locations . . . . . . . . . . . 164


A.2 Undisturbed flow spanwise microphone Locations . . . . . . . . . . . 165
A.3 Step1 streamwise microphone locations . . . . . . . . . . . . . . . . . 167
A.4 Step1 spanwise microphone locations . . . . . . . . . . . . . . . . . . 168
A.5 Step2 streamwise microphone locations . . . . . . . . . . . . . . . . . 170
A.6 Step2 spanwise microphone locations . . . . . . . . . . . . . . . . . . 171
A.7 Step4 streamwise microphone locations . . . . . . . . . . . . . . . . . 174
A.8 Step4 spanwise microphone locations . . . . . . . . . . . . . . . . . . 175
A.9 Step8 streamwise microphone locations . . . . . . . . . . . . . . . . . 178
A.10 Step8 spanwise microphone locations . . . . . . . . . . . . . . . . . . 179

xv
Nomenclature
Abbreviations
AR Aspect Ratio
BFS Backward-facing Step
CBF Conventional Beamforming
CFD Computational Fluid Dynamics
CSM Cross Spectal Matrix
EWT UNSW Educational Wind Tunnel
FBS Forward-backward Facing Step
FBS Forward-backward facing step
FFS Forward-facing Step
GCI Grid Convergence Index
KH Kelvin-Helmholtz
LES Large Eddy Simulation
POD Proper Orthogonal Decomposition
PSD Power Spectral Density
SNR Singal-to-Noise Ratio
SPOD Spectral Proper Orthogonal Decomposition
TKE Turbulent Kinetic Energy
UAT UNSW Anechoic Wind Tunnel
VK von Kármán

Greek Symbols

xvi
Nomenclature xvii

γ Coherence coefficient
δ Boundary layer thickness
δ∗ Boundary layer displacement thickness
θ Boundary layer momentum thickness
∆τ Time delay
λ Acoustic wavelength
τw Wall shear sterss
Λ SPOD eigenvalues
Φ SPOD eigenmodes
Θfk SPOD eigenmodes in frequency domain
∆ A finite increment of
µ Coefficient of viscosity
ω Angular frequency
Φpp Pressure spectra
ρ Fluid density

Roman Symbols
M Mach number
Cab Correlation coefficient between quantity a, b
p Pressure
p0 Pressure fluctuation
Rab Correlation between quantity a, b
Cf Skin friction coefficient
Cp Wall pressure coefficient
c0 Speed of sound
f Frequency
He Helmholtz number
q∞ Dynamic pressure
Re Reynolds number
St Strouhal number
t Time
u0 Streamwise velocity fluctuation
v0 wall-normal velocity fluctuation
w0 Spanwise velocity fluctuation
xr1 Reattachment length on the top surface
xr2 Reattachment length downstream of the step
xshif t Sound source shifted distance
xs Upstream separation length
U∞ Free stream velocity

Superscripts and Subscripts

∞ Freestream condition
0
Fluctuating component

max Maximum value

min Minimum value

rms Root mean square value

Operators
(·) Mean
∂ Partial derivative
Σ Summation
E[] Expected value
[]T Transpose

xviii
Chapter 1

Introduction

1.1 Background

The aeroacoustic performance of vehicles is an important issue in the automotive


industry. The flow-induced noise could internally influence passengers’ comfort and
safety and externally cause the noise pollution in the environment. The understand-
ing of the interaction between the flow and the structure is therefore necessary for
the noise reduction.
Surface discontinuities are often found in engineering applications, such as the
uneven plate joints on the surface of vehicles, hatches, doors, or gaps between pan-
els. Noise emission from these surface discontinuities is considered one of the major
sources of the cabin noise. When exposed to fluid flow, these surface irregularities
may separate the incoming boundary layer and modify the flow field. Step configu-
rations are among one of the most common geometries of these surface irregularities.
The presence of the step enhances the turbulence levels in the flow-field resulting
in large wall pressure fluctuations, which can be a source of noise, vibration, and
structural fatigue.
Previous studies [5, 24, 39, 40] have considered the Forward-facing step (FFS)
and Backward-facing step (BFS) immersed in the boundary layer. Awasthi et

1
Chapter 1. Introduction 2

al. [5] concluded that the geometry of the FFS leading edge is the critical factor
in flow field modification as well as far-field noise production. Ji and Wang [39,
40] performed a large-eddy simulation (LES) for both FFS and BFS with different
step height (h) to undisturbed boundary layer thickness (δ) ratio (h/δ) and found
that the sharp corners are critical in flow alteration and sound generation in both
scenarios. Pure FFS or BFS cases (with only one sharp corner involved) have been
widely investigated; however, the flow over a Forward-back facing step (FBS), and
its associated noise production, is not yet fully understood.
The flow configuration for the FBS is different from FFS as the trailing edge of
the step is included in the flow field. Figure 1.1 illustrates the general flow field of
a FBS. Flow separation occurs upstream of the leading edge owing to the adverse
pressure gradient induced by the step. The flow reattaches on the vertical face of the
step and separates again at the sharp leading edge, forming the upstream separation
bubble over the upper surface. The excited separated shear layer will then reattach
on top of the step (if the step is long enough) and forms the upper corner separation
bubble. Further downstream, the existence of the trailing edge causes the backward
separation. The shear layer encompassing these separation bubbles induces intense
turbulence, which is responsible for high-levels of surface pressure fluctuations in
the separation-reattachment region.
Freestream velocity
Upper corner separation bubble

Reattachment point of
Separated shear layer upper corner separation bubble

Time-averaged
dividing streamline
Upper corner
separation point
Re-established

boundary layer

Upstream Backward step Reattachment region


separation bubble separation bubble

Figure 1.1: Sketch of FBS flow field.

The separation and reattachment, caused by the sharp corners in the flow field,
were concluded to be critical factors to the noise production [49]. Doolan and Moreau
Chapter 1. Introduction 3

[19] showed that sound production was affected by the FBS leading edge geometry.
Yet the location of the noise source was found to vary with the step aspect ratio
(AR, step length to step height l/h) [74]. The influences of the FBS aspect ratio on
the overall flow field and noise production require further investigation.
To address this knowledge gap, a combined experimental and numerical study of
the forward-backward facing step pair with different aspect ratios will be presented
in this study.
Chapter 1. Introduction 4

1.2 Aims

The overall aim of this research is to comprehensively understand the flow and noise
generation of the FBS with different aspect ratios placed within low Mach number
turbulent boundary layers.
Specifically, the objectives of this research are:

1. Design and construct a new FBS flow and noise experiment.

2. Measure the steady and unsteady wall pressure for BS with a variety of aspect
ratio.

3. Develop a LES model of FBS flow for a variety of AR.

4. Measure the acoustic radiation from FBS at various AR.

5. Draw together the experimental and computational results to develop new


scaling laws for FBS noise.

1.3 Thesis Outline

A literature review is provided in Chapter 2, summarizing some of the past studies


on FFS, BFS, and FBS. The knowledge gaps are identified and used to develop
the above aims and objectives of this thesis. Chapter 3 describes the experimental
and numerical methodologies employed in this study. The experimental setup for
the wall pressure measurements, acoustic measurements, and oil flow visualizations
are first introduced, followed by the description of the numerical simulation. The
details of the data analysis such as space-time correlation and modal decomposition
methods are presented at the end of this chapter.
The results and discussion are provided in Chapters4, 5, and 6. The oil flow
visualizations and wall pressure measurement results are combined and presented
in Chapter 4, where the turbulent flow characteristics for different AR FBS are
described. In Chapter 5, the flow structures and turbulence statistics calculated by
Chapter 1. Introduction 5

the CFD simulation are presented. The modal decomposition method is applied to
the CFD results further extract the coherent structures in the flow field. Chapter 6
shows the acoustic features of the FBS, including the far-field sound spectra and the
acoustic beamforming results. New scaling laws developed using experimental and
computational results are also presented in this chapter to address the influences of
the step AR.
Finally, conclusions summarizing the findings in this thesis and the recommen-
dations for future work are given in Chapter 7.
Chapter 2

Literature Review

2.1 Overview

This chapter reviews past studies on the flow characteristics and the noise generation
mechanisms of step-shape surface irregularities. The theory of noise induced by low
Mach number flow is first reviewed in Section 2.2. Following this, the discussion of
past studies on the aeroacoustic features of the step configurations are separated into
three sections: forward-facing steps (Section 2.3.1), backward-facing steps (Section
2.3.2), and forward-backward facing step-pairs (Section 2.3.3). Finally, in Section
2.4, the knowledge gaps are summarized.

6
Chapter 2. Literature Review 7

2.2 Noise Generated by Low Mach Number Flow

The theory of sound generated by aerodynamic phenomena was first proposed by


Lighthill [51]. Lighthill identified the non-stationary fluctuating flow effects that
generate the acoustic waves as the source term of the wave equation and derived the
well-known Lighthill’s acoustic analogy. The Lighthill’s equation is given as

∂ 2 ρ0 2 2 0 ∂ 2 Tij
− c 0 ∇ ρ = (2.1)
∂t2 ∂xi ∂xj

where
Tij = ρui uj + (pij − c20 ρ0 )δij (2.2)

is Lighthill’s stress tensor. ρ0 is the density perturbation to the surrounding medium,


c0 is the speed of sound, and pij is the pressure fluctuations. The solution of the
equation can be obtained by applying a Green’s function [32] :

G(x, y, t, τ ) = δ(x − y)δ(t − τ ) (2.3)

where δ is the Dirac delta function. The coordinate x is placed at the observer’s
location (far-field) and y refers to the position in the flow where the sources emerge
(near field).
Lighthill’s analogy was later extended by Curle [18] and Ffowcs Williams &
Hawkings [27], who considered the influence of a solid boundary as a part of the
source term and added a moving surface effect, respectively. The Ffowcs Williams
and Hawkings (FWH) equation is widely applied when the moving surface is included
in the flow field, such as rotating blades. As the step is stationary, the moving surface
term in the FWH equation can be eliminated, and it reduces to Curle’s equation.
Chapter 2. Literature Review 8

The solution to Curle’s equation can be written as

Z T Z  
0 1 ∂G ∂(ρuj )
ρ (x, t) = 2 (pij + ρui uj ) +G nj dS(y)dτ
c0 −T S ∂yi ∂τ
(2.4)
1 T
Z Z  2 
∂ G
+ 2 Tij (y, τ )dV (y)dτ
c0 −T V ∂yi ∂yj

where G is the Green’s function. The right-hand side of the equation categorizes
the sources into two terms, the dipole like term and the quadrupole like term. The
surface integral term creates a dipole-like directivity and is related to the unsteady
surface pressure loading. The volume integral is the quadrupole term, which contains
the sources such as sound radiated by both turbulence and flow distortion. The far-
field acoustic intensity of the dipole term is scaled on the 6th power of the flow speed,
while the intensity of the quadruple term is scaled on the 8th power. This indicates
the ratio of quadrupole intensity to dipole intensity is of the order of O(M 2 ), where
M is the Mach number. Thus, in low Mach number flow, which is the case in
this study, Lighthill’s stress tensor term Tij is less important and the dipole term
is dominant. In the low Mach number flow field, surface discontinuities will alter
the flow structure which will induce pressure fluctuations that are subsequently
converted to acoustic pressure waves.

2.3 Step Induced Flow Field and Noise

2.3.1 Forward-facing Step

Figure 2.1 illustrates the overall flow field about a forward-facing step. The step has
a sharp leading edge and is completely immersed in the boundary layer (h/δ < 1).
Flow separation occurs upstream of the edge owing to the adverse pressure gradient
induced by the step. The flow reattaches on the vertical face of the step and separates
again at the sharp edge. The excited separated shear layer will then reattach at a
couple of step heights downstream of the vertical face at the reattachment point, xr .
Overall, two separation bubbles can be found on each side of the step. The unsteady
Chapter 2. Literature Review 9

loading acting on each surface, as given in Eq. 2.4, is regarded as the major source of
noise generation in this circumstance. Several studies [3, 5, 26, 29, 40, 57, 62, 66, 69]
have been conducted to examine the separation and reattachment phenomena of
a FFS, and some of them also related these flow behaviours to far-field acoustic
radiation.

Separated shear layer


Upstream Downstream
separation bubble separation bubble
Figure 2.1: Sketch of FFS flow field.

Pearson et al. [62] carried out a study that specifically targets to the upstream
recirculation region of a forward-facing step. The step was entirely submerged in a
fully turbulent boundary layer (h/δ=0.68). It was noticed that most of the time,
the upstream separation bubble stayed in closed form, which means the reattach-
ment point can be found on the vertical face of the step. Yet occasionally, the
reattachment point would rise over the step edge and result in growth of the down-
stream separation bubble. The dominant mechanism of this sudden propulsion of
the upstream bubble was attributed to the spanwise flow along the edge, which was
influenced by upstream flow condition. Therefore the upstream bubble was weakly
affected by the step geometry. The study observed out-of-plane mass flux in the
closed separation bubble, which has also been noticed by Awasthi [3]. Awasthi sug-
gested that the upstream bubble lifted small scale structures away from the wall,
leading to the amplification of the low frequency wall pressure spectral level in this
region.
Downstream of the FFS, the separated shear layer that forms at the sharp edge
Chapter 2. Literature Review 10

largely influences the downstream flow. Sigurdson [71] and Kiya & Sasaki [41, 42]
concluded that the unsteadiness of this separated shear layer was mainly determined
by two mechanisms; Kelvin–Helmholtz (KH) vorticity roll-up & shedding and shear
layer flapping motion. The latter is associated with the oscillation of the reattach-
ment point. Figure 2.2 [42] sketches a period of vortex shedding. It can be seen that
the vortex shedding motion corresponds with the mean streamline moving back and
forth, which is referred to as the shear layer flapping motion.

Figure 2.2: Schematic representation of a period of vortex shedding. Taken from Kiya &
Sasaki [42].

Sherry et al. [69] conducted a particle image velocimetry (PIV) experiment to


understand the flow configuration of the recirculating region downstream of the
edge. The outcome demonstrated that the reattachment length is highly dependent
on Reynolds number based on step height (Reh ) especially when the step is immersed
in the boundary layer. The reattachment length was found to be most sensitive to
Reh when Reh was less than 8500. The author suggested that the turbulent mixing
behaviour in the separated shear layer would influence the reattachment length.
Nematollahia & Tachie [57] considered h/δ = 0.23 & 0.15 FFS in different up-
Chapter 2. Literature Review 11

stream wall roughness conditions and found that the downstream reattachment
length decreases with increasing upstream roughness condition. The results sug-
gested that the separated shear layer flapping motion is affected by the upstream
wall roughness.
Awasthi et al. [5] presented an experimental study of three different step sizes
h/δ = 0.038, 0.15, and 0.6. The reattachment length was measured by flow oil vi-
sualization, and the near field pressure fluctuations were measured by microphones
placed underneath the wall (on the horizontal and vertical surface). The results show
that the upstream reattachment length was weakly dependent on the step size, and
the separation bubble in this region amplifies low-frequency pressure oscillations.
The author suggested that the high-frequency structures (small scale eddies) were
convected away from the wall by the recirculation bubble, leaving the larger eddies
and their asociated low-frequency pressure signatures acting on the surface. Fur-
ther, the downstream reattachment length was found to be strongly affected by the
step geometry. The reattachment length is about 1.6h for the smallest step and
increases to 4.2 h for the largest step. Also, opposite to the upstream flow regime,
the downstream separation bubble elevates the pressure spectra across the entire
frequency range.
Awasthi et al. [5] found that the downstream bubble is more sensitive to step
dimensions. The observation agrees with Farabee and Casarella [26] in that the
features of the wall pressure fluctuations in the two separation bubbles were different.
In Ref. [26], the adverse pressure gradient was concluded to be the main factor that
affected the upstream separation bubble, while the pressure fluctuations within the
downstream separation bubble were mostly influenced by the separated shear layer.
An early investigation of the far-field sound radiated from the FFS was performed
by Howe [36], who attributed the acoustic sources to the diffraction of the flat, rigid-
wall pressure field and derived an equation for far-field sound radiation that demon-
strates dipole directivity in the streamwise direction. Ji & Wang [39] performed the
numerical simulations of both FFS and BFS cases. The step height to boundary
Chapter 2. Literature Review 12

layer thickness ratio was varied from 0.0083 to 0.53. The study provides a clear
understanding of noise generation of both the forward-facing and backward-facing
step. The authors suggested that although the largest wall pressure fluctuations
were found to be near the reattachment point, the sources of the noise were located
at the vicinity of the sharp edge in both steps. It was concluded that the FFS noise
is determined by both diffraction and turbulence modification around the corner.
For larger size steps, noise created by the FFS is louder than by the BFS due to the
strong vertical velocity components created by the leading edge that increases the
turbulence level. The contribution of this vertical velocity is negligible in the small
step configurations.
Glegg et al. [29] proposed an analytical model to physically address the flow
mechanism and sound radiation over a sharp corner step. A tailored Green’s function
was applied to derive the model. The model shows that the acoustic source is
dominated by unsteady loading created by the downstream separation bubble and
the shed vorticity is a second order effect compared to the pressure loading. The
prediction showed excellent agreement with previous experimental results [3, 66] and
showed a dipole like radiation pattern. The study paints a clear picture of the flow
physics around the forward-facing step and provides a helpful way to understand the
associated noise radiation. However, the limitation of this model is the scale of the
step. If the step height is close to the acoustic wavelength or the spanwise length is
considered to be finite, the acoustically compact assumption will no longer be valid.
The effect of scattering and three-dimensionality should be included to obtain a more
comprehensive model. Schram et al. [45] suggested that non-compact geometry
would alter the far-field acoustic pressure spectra and found a better prediction by
implementing Curle’s analogy [18] on non-compact bodies.
Catlett [66] conducted an acoustic measurement of FFS, BFS , and gaps, which
were formed by placing a forward-facing step behind a BFS (Fig.2.3). The step
heights ranged from 0.1 ∼ 1 boundary layer thickness. The results showed that
noise generated from the FFS is 10 dB louder than the BFS, which concurs with
Chapter 2. Literature Review 13

the numerical prediction by Ji & Wang [39]. Also, the forward-facing step noise
was found to be dependent on step height, while BFS noise is not. For the gap
configurations, forward step sound was dominant if the edge of the rear step stayed
outside of the wake region (upstream of the reattachment point). When the rear
step was exposed to the reattaching shear layer, both forward and backward facing
step noise appeared but forward facing step noise was more obvious. If the edge
of the rear step was immersed in the wake region, the backward facing step noise
dominated.

Figure 2.3: Step configurations investigated in [66]. Taken from Catlett et al. [66]

The edge of the upper corner of the FFS is considered to be responsible for noise
generation. Awasthi et al. [6, 8] presented a series of studies that investigated the
near-field (wall pressure) and far-field (sound) behaviours of the rounded FFS. The
step height to the boundary layer thickness ratio h/δ was 0.26. Corner rounding
radius to step height ratios (r/h) of 0%, 6.25%, 12.5%, and 25% were studied. It was
observed that the downstream reattachment length decreased with an increase in
rounding, while the upstream separation was barely affected. The authors suggested
that the rounded corner directed the separated shear layer closer to the surface. At
a free stream velocity of 30 m/s, the sound measurement results showed that for
6.25% rounding, the sound power at frequencies around 3 kHz to 4 kHz was reduced
by 3.7 dB. While under the same flow speed, for large rounding, attenuation can be
found throughout the entire frequency range with largest reduction of 16.6 dB at
Chapter 2. Literature Review 14

4.5 kHz. The acoustic beamforming results also showed that the noise sources were
mainly located at the edge of the FFS.
Awasthi et al. [9] proposed an analytical model for FFS far-field sound prediction.
The predicted result showed a quantitatively good agreement with the experimental
data and presented a dipole like directivity when the step is acoustically compact.
The model indicates that the far-field sound is mostly determined by the incoming
flow speed and the front face geometry.

2.3.2 Backward-facing Step

Flow over a backward-facing step (BFS) is commonly founded in engineering ap-


plications, such as at the inlet of the combustion chamber or in pipe flow. The
flow structure of the BFS has been widely investigated in the past few decades.
Figure 2.4 illustrates the flow field around the step. Flow separation occurs at the
edge of the step, and the separated shear layer is formed due to the velocity differ-
ence inside and outside of the layer. The flow reattachment occurs at several step
heights downstream of the edge, where the shear layer strikes the wall. A separation
bubble downstream of the edge is formed between the reattachment point and the
step. Downstream of the reattachment point, the boundary layer will progressively
redevelop.

Separated shear layer

Re-established
boundary layer

Backward step
separation bubble Reattachment region

Figure 2.4: Sketch of BFS flow field.

The downstream reattachment length is weakly dependent on Reh as the values


are mostly around 5h ∼ 7h across a wide range of Reh [14, 26, 40, 48, 73]. The
Chapter 2. Literature Review 15

separated shear layer formed at the backward-facing edge exhibits the shear layer
flapping motion [41] which is associated with vortex shedding events. The similarity
between the separated shear layer and the mixing layer was pointed out by Eaton
& Johnston [22] and Le et al. [48]. Eaton & Johnston [22] experimentally measured
the growth rate of the shear layer and found that it is akin to a mixing layer.
Farabee and Casarella [26] conducted wall pressure measurements to investigate
the unsteady loading on the surfaces of both forward-facing steps and backward-
facing steps. For the BFS, it was noticed that at the reattachment point, the
spectra exhibit high spectral levels at low frequencies but have lower spectral level
at higher frequencies. This indicates that reattached flow impinging on the wall
mostly consists of large scale structures formed in the shear layer and features a lack
of small-scale velocity fluctuations. Ji & Wang [40] also noticed a similar pattern
in the BFS unsteady wall pressure spectra obtained from CFD calculations. They
pointed out that the reattachment length is able to reasonably scale the downstream
r.m.s pressure distributions and collapse the peaks at x ≈ xr for different step height
cases.
Jacob et al. [38] carried out an experimental study of the sound induced by a
BFS. Laser Doppler Anemometry (LDA) was employed for the aerodynamic mea-
surement and acoustic data were collected using a circular array of nine microphones.
The results showed that the major acoustic source was located at 2∼3 step heights
downstream of the edge, which was near the reattachment region. The sound source
location is consistent with where the highest turbulent kinetic energy (T.K.E.) was
measured. This finding agrees with Catlett et al. [66] but conflicts to Ji & Wang
[39], who argued that the main noise source of a BFS is concentrated at the edge.
Ji & Wang [39] claimed that in the BFS scenario, the diffraction of the boundary
layer source field is the main mechanism of noise generation and is near the edge.
They also found that the noise created by the FFS is generally louder than that by
the BFS as the the dominant mechanism of the FFS is larger in magnitude. Note
that both studies did not take the upstream flow disturbances into account where
Chapter 2. Literature Review 16

in FBS flow in the downstream area was affected by the upper corner disturbance.

2.3.3 Forward-backward Facing Step

Figure 1.1 illustrates the general flow field of a forward-backward facing step. Due
to the adverse pressure gradient, the boundary layer will separate and reattach
upstream of the step, which is similar to the FFS configuration. Downstream of
the step leading edge, for the large AR step, the separated shear layer exhibits two
reattachments with one on the top surface and the other one downstream of the
step. Meanwhile for the small AR step where the step length is shorter than the
reattachment length, the separated shear layer will only reattach downstream of
the step. The position of the downstream reattachment point depends on the step
height and the length. After the reattachment, the turbulent boundary layer will
re-develop in downstream flow field.
Fang & Tachie [23] conducted a particle image velocimetry (PIV) experiment to
investigate the flow field downstream of the FBS leading edge. An AR = 2.36 with
h/δ = 0.2 step was considered. In this case, the step AR is slightly larger than the
reattachment length on the step top surface, therefore two downstream separation
bubbles are visible. They observed the mixing behaviour of two separated shear
layers (emanating from the leading edge and the trailing edge, respectively) and
noticed that the rear part of the first separation bubble (on the step) is influenced by
the second separation bubble (downstream of the step). The unsteady features of the
first separation bubble were studied using proper orthogonal decomposition (POD)
and its first mode was attributed to the separated shear layer flapping motion, which
is in charge of the separation bubble enlargement and shrinkage. The results also
provided evidence of the temporal correlation between the two separation bubbles,
indicating the influence of the shear layer flapping motion on the downstream bubble.
Moore et al. [54] conducted PIV experiment with AR = 1 and 5 two-dimensional
squares placed in free stream flow. The study showed that the for AR = 1 square
where the reattachment point was not observed, the leading edge separated shear
Chapter 2. Literature Review 17

layer flapping motion bled into the downstream wake and enhanced the downstream
von Kármán instability. As in [54] the square was placed in a free stream, the
effect of the enhanced downstream instability on the wall pressure fluctuations is
still unclear.
Compared to the FFS, the acoustic characteristics of the FBS have only been
considered in a few studies. Leclercq et al. [49] conducted an experimental campaign
for a FBS with h/δ > 1 (142%) and AR = 10. The experimental data was later
validated through numerical simulation by Addad et al. [1]. The step was long
enough to keep the reattachment point on the top surface of the step. The backward
reattachment length was affected by upstream events and reduced by 3h compared to
typical backward flow. The acoustic results revealed that the leading edge dominated
noise production and that the source was located at the leading edge separation
region, where high velocity and pressure fluctuations were measured. They also
show that the shear layer generated by flow separation significantly contributed to
the wall pressure levels.
Doolan & Moreau [19] performed an experimental study of noise generated by
sub-boundary layer FBS (h/δ < 1). Flow structure modification around the step
was investigated and the far-field sound measurement was carried out. The study
considered various geometries of the step, such as triangular and rectangular steps
with steps height to boundary layer thickness ratio h/δ = 0.26, 0.526 and 0.79, and
AR = 8, 4, and 2.67. Hot wire anemometry was used to collect boundary layer
information. The measurements indicated that the upstream separation length is
reduced for triangular steps as the adverse pressure gradient was smaller. This
was not found by Awasthi et al. [8] since the adverse pressure gradient associated
with the rounded corner is not large enough to influence the pressure gradient.
The slanted leading edge reduces the strong adverse pressure gradient that occurs
in front of the rectangular step. The energy extraction from the triangular step
is therefore smaller, which implies lower fluctuating pressure levels. Meanwhile,
the upstream separation is found to be coupled with the downstream perturbed
Chapter 2. Literature Review 18

region. Yet Leclercq et al. [49] observed no coherent information was transported
downstream of the edge. The difference may have originated from the step height.
The steps in Leclercq’s [49] experiment were higher than the incoming boundary
layer thickness. From the far-field results of [19], it can be noticed that for the
larger size step, the triangular step created less noise. It can be concluded that the
front surface (leading edge) in a forward-backward facing step dominates the noise
production.
Springer et al. [74] performed both numerical simulations and experimental
measurements of a forward-backward facing step with AR = 1. Although h/δ is
not clear in this study, it provides a general understanding of the small aspect ratio
FBS. The numerical calculation indicated that the source term was the shear layer
behind the leading edge.

2.4 Summary and Research Gaps

The flow structures of a FFS and BFS have been widely studied. Separation and
reattachment were concluded to be critical factors controlling flow modification. The
downstream separation bubble of a FFS is highly sensitive to the step geometry, such
as the shape of the edge or the step height. The separated shear layer can be found
in both forward and backward steps, and is responsible for the high level of wall
pressure fluctuations. In terms of sound production, the FFS was found to create
more noise compared with the BFS, and the difference becomes more obvious as the
step size increases. The noise source of the FFS was consistently claimed to be at
the vicinity of the edge, while descriptions of the noise mechanism of the BFS are
divergent. This requires further examination.
For forward-backward facing steps, the separation and reattachment, which are
affected by the sharp corners in the flow field, were concluded to be critical factors
concerning the flow. The sound production of the FBS is mostly affected by the
leading edge geometry; however, the location of the noise source was found to vary
with the step aspect ratio. If the step was long enough to keep the reattachment
Chapter 2. Literature Review 19

point on the top surface, the noise production is determined by the step leading edge.
While for the small AR step where the reattachment point is not observed on the top
surface, the noise source term seems to move a little downstream of the step. The
influence of step aspect ratio of the FBS on the the far-field noise generation remains
unclear. Therefore, the present study aims to address the following knowledge gaps:

1. The influence of the step aspect ratio of the FBS on the flow field. How does
the step length alter the unsteady features of the flow?

2. The exact noise source location(s) of the FBS with different aspect ratios.

3. The influences of the FBS aspect ratio on the far-field acoustic characteristics,
i.e. the spectral level and shape. And how can the acoustic level be scaled?
Chapter 3

Methodology

3.1 Overview

This chapter describes the experimental and numerical methodology employed to in-
vestigate the aeroacoustic and aerodynamic behaviours of flow over forward-backward
l
facing steps with different aspect ratio (AR = h
). First, the experimental setup,
step model geometry, and the instrumentation will be introduced. The far-field
acoustic measurement and wall pressure measurement were both conducted with
the same setup in the UNSW anechoic wind tunnel UAT (details can be found in
Section 3.2). Details of the acoustic data acquisition procedure and beamforming
are given in Section 3.2.3. The remote microphone two-step calibration methods for
the unsteady wall pressure measurements are introduced in Section 3.2.4. The Oil
flow visualization was carried out in the UNSW educational wind tunnel (EWT) to
investigate the flow separation and reattachment length (Section 3.3). Large Eddy
simulation was used to compute the turbulent flow field. Computational details
are given in Section 3.4. Finally, the data analysis methods such as space-time
correlation and modal decomposition methods are presented in Section 3.5.

20
Chapter 3. Methodology 21

3.2 Far-field Noise and Wall Pressure Measure-

ments

3.2.1 UNSW Anechoic Wind Tunnel

The far-field acoustic measurement and surface pressure measurements were con-
ducted in the UNSW Anechoic Wind Tunnel (UAT). The facility is an open-jet
type wind tunnel with a 0.455 m × 0.455 m test-section enclosed by a 3 m × 3.2
m × 2.15 m anechoic chamber. The UAT is a suction type tunnel where a centrifu-
gal fan is placed downstream of the test-section and is connected to the anechoic
chamber through a diffuser. The maximum test speed of the wind tunnel is around
70 m/s with the free stream turbulence intensity ≈ 0.3% at 20 m/s. The cham-
ber is acoustically treated with the 0.54 m thick Melamine foam attached on the
surrounding walls that provides an anechoic condition above 300 Hz.
Figure 3.1 is a schematic diagram of the UAT. Before entering the anechoic
chamber, the flow is accelerated through a bell-mouth inlet consisting of turbulence
reduction screens and a honeycomb mesh. The flow then enters a 5.5:1 contraction
upstream of the test-section. An endplate was flush mounted on the side of the inlet
and attached to this plate were the steps investigated in this study. After passing
over the step and the end plate, the jet entering the anechoic chamber through
the inlet is captured by a collector (also known as the jet-catcher) placed 1.7 m
downstream of the inlet. The fan is connected to a U-bend shape muffler-diffuser
downstream of the chamber which attenuates the noise generated by the fan so that
background noise in the chamber is low. More details of the facility can be found in
Ref. [20]
Chapter 3. Methodology 22

Figure 3.1: Schematic diagram of the UNSW anechoic wind tunnel AWT.

3.2.2 Test Parameters and Forward-backward Facing Step

Configuration

Four different step AR were considered in the experiments (Fig. 3.2). The step
height for all testing cases is 10 mm while the step length varies from 10, 20, 40
to 80 mm, resulting in aspect ratios of 1, 2, 4, and 8. Note that for measurements
conducted in the UAT (wall pressure measurements and acoustic measurements),
the terms Step1, Step2, Step4, and Step8 are used to represent test cases AR = 1,
2, 4, and 8, respectively.

Unrounded
rectanguler step
Step1, AR = 1

y Step2, AR = 2
x

Step4, AR = 4

Step8, AR = 8

Figure 3.2: Rectangular steps with 4 different aspect ratios considered in the wall pressure
and acoustic measurements.
Chapter 3. Methodology 23

The FBS test model was mounted on an endplate as shown in Fig. 3.3. The
coordinate system is shown with the origin set at the leading edge and at the middle
of the step span. The step width (in the z direction) is 470 mm, which is longer
than the free jet width (455 mm), ensuring flow two-dimensionality.
The grey area in Fig. 3.3 (downstream of the step) represents a separate plate
witch is slidable in the x-direction. This movable part allows the pressure taps that
are mounted in this area to adapt to different step lengths. This guarantees the
same distance from the trailing edge to the testing points are the same for each
case. Details of wall pressure measurement locations are given in AppendixA.

Table 3.1: Test cases and step parameters

Test
h (mm) l (mm) AR U∞ (m/s) Reh h/δ
case
20 14,078 0.526
Step1 10 10 1 30 21,116 0.526
40 28,155 0.5
20 14,078 0.526
Step2 10 20 2 30 21,116 0.526
40 28,155 0.5
20 14,078 0.526
Step4 10 40 4 30 21,116 0.526
40 28,155 0.5
20 14,078 0.526
Step8 10 80 8 30 21,116 0.526
40 28,155 0.5

Free stream velocities in these measurements were set to 20, 30 and 40 m/s. The
corresponding Reynolds number based on step height (Reh ) varies from 1.4 × 105
to 2.8 × 105 (see Table 3.1). The step leading edge was located 400 mm away
from the jet exit plane, where the coordinate system origin is placed (see Fig. 3.3).
The undisturbed boundary layer thickness was measured in the absence of the step
at (x, z) = (0, 0) using hot wire anemometry. A single-sensor Dantec Dynamics®
miniature hot-wire probe was used, and the probe was controlled by an IFA300
Constant Temperature Anemometer (CTA) system. The velocity data was collected
at 52100 Hz for 32 seconds at each position. The boundary layer profiles at U∞ = 20,
30 ,40 m/s are given in Fig. 3.5. A good comparison for velocity profiles and the law
Chapter 3. Methodology 24

of wall can be seen in Fig. 3.5, which indicates the flow is turbulent. The turbulence
stress in the streamwise direction (x), u02 , is given in Fig. 3.6. The undisturbed
boundary layer thickness at step location ((x, z) = (0, 0)) is given in Table 3.2. Note
that the skin friction coefficient Cf for each flow speed was estimated using Clauser’s
method [15]. The boundary layer thickness was 20 mm at U∞ = 40 m/s and the step
height to boundary layer ratio (h / δ) was around 0.5 for all testing cases (boundary
layer information refers to Table 3.2).

Figure 3.3: Schematic diagram of experimental setup. Red dashed line indicates the
location of the step leading edge (x = 0).

Figure 3.4: Sideview of the experimental setup.


Chapter 3. Methodology 25

28

26

24

22

20

18

16

14

12
102 103 104

Figure 3.5: Boundary layer profile measured at (x, z) = (0, 0) without step for flow speed
= 20, 30, 40 m/s. Solid line, u+ = (1/0.41) ln (y + ) + 5; Blue ◦,U∞ = 20 m/s; Red 4,U∞
= 30 m/s ;Yellow ,U∞ = 40 m/s.

1.8

1.6

1.4

1.2

0.8

0.6

0.4

0.2

0
0 1 2 3 4 5 6
10-3

2 measured at (x, y) = (0, 0) without step for U


Figure 3.6: Turbulence stress u02 /U∞ ∞ =
20, 30, 40 m/s.
Chapter 3. Methodology 26

Table 3.2: Undisturbed boundary layer information

U∞ δ δ∗ θ
(x, y) H Reδ Reδ∗ Reθ Cf
(m/s) (mm) (mm) (mm)

20 (0, 0) 19 1.98 1.57 1.27 25851 2691 2127 0.0035

30 (0, 0) 19 1.82 1.47 1.24 38776 3697 2988 0.0034

40 (0, 0) 20 1.77 1.45 1.23 54422 4793 3923 0.0033

3.2.3 Far-field Acoustic Measurements

Acoustic measurements were carried out using a 64-microphone phased array con-
sisting of 1/4” GRAS 40PH phase matched microphones with flat frequency response
range between 50 Hz and 20 kHz. The nominal sensitivity of each microphone at
250 Hz is 50 mV/Pa (±20dB). The microphones are mounted on a 2 m × 2 m per-
forated steel grid, which is parallel with the endplate and located 1.3 meters away
from it (y direction). The center microphone is 150 mm downstream (x direction)
of the leading edge of the step which results in an 83.4◦ observation angle θ (Fig.
3.3 & 3.4). The array microphones are arranged in spiral manner with 9 spiral
arms (7 microphones in each arm) and a center microphone (Fig. 3.7). The center
microphone is aligned with the step spanwise centerline (z = 0).
The signal to noise ratio (SNR) between the step noise level and the background
noise level was calculated using Eq.3.1 [4]:

 
Φpp (f )
SN R = 10 log10 (3.1)
Φpp, no step (f )

where Φpp is the spectral density measured with a FBS and Φpp,no step is the spectral
density measured at the same flow velocity but without the step. The SNR ratio
was used in the present study to determine the frequency range that is of interest
for the FBS scenario.
The microphone signals were acquired using a National Instruments PXI data
acquisiton system. The sampling rate was 65,536 Hz with a duration of 32 seconds
Chapter 3. Methodology 27

Figure 3.7: 64-microphone phased array.

for each test. The time series signal obtained from each microphone was projected
to the frequency domain using the Fourier transform with record length = 8192,
overlap = 50% and a Hanning window was used on each window. The array has an
optimal design for beamforming applications [63].

Aeroacoustic Beamforming

The far-field acoustic signals are beamformed using conventional beamforming (CBF)
[56]. The cross-spectrum matrix C of the signals collected by each microphone in
the microphone array is steered to pre-determined points in a scanning plane, where
the source at the grid point is assumed to be monopole. For the signal from a specific
grid point k, the phase-shift between each microphone is considered as the steering
vector. This steering vector is based on the monopole free-field Green’s function.
The steering vector hk is written as

hk = [w1 , w2 , . . . , wn ]T (3.2)
Chapter 3. Methodology 28

where wj is the weighting factor of the jth microphone. It is defined as

wj = exp(2πik0 (rj,k − r0,k )) (3.3)

where k0 is the acoustic wavenumber, rj,k is the distance between grid point k on
the scanning plane and jth microphone in the array and r0,k is the distance between
grid point k and the reference microphone, which in this measurement is the center
microphone of the array. The beamforming output at different frequency for a n-
microphone array is calculated as

1 T
B(f ) = h Ch (3.4)
n2

where n = 64 in the present measurement, C is the cross-spectrum matrix of the


signal, and h is the steering vector given in Eq. 3.2. The diagonal components are
generally removed from C to obtain the higher signal-to-noise ratio Cdiag.=0 . The
output expression is then given as

1
B(f ) = hT Cdiag.=0 h. (3.5)
n2 −n

Also, the background far-field noise (measured in the absence of the step) was
pre-subtracted from the step noise cross-spectra matrix (CSM) to eliminate the
acoustic contamination from the tunnel and the CSM diagonal was removed to re-
duce the self-noise before beamforming. The beamforming was performed in 1/12th
octave bands. A shear layer correction [61] was used in the beamforming algorithm
to correct the effect of sound convection due to the flow in the open-jet wind tunnel.
The model suggests that the shifted distance of the sound source xshif t in the flow
can be determined by

xshif t = M d, (3.6)

where M is the Mach number of mean flow and d is the distance from the source to
Chapter 3. Methodology 29

the shear layer edge formed at the wind tunnel inlet.


Area integrated spectra for each step will be considered in this study to eliminate
noise contamination from other parts of the facility. The integration was applied fol-
lowing the procedure described in Mueller [56], where the integrated power spectral
level at each frequency band IB (f ) is calculated as

ΣS B∆S (f )
IB (f ) = (3.7)
ΣS PSFc∆S (f )

where S is the area of interest (determined by the beamformed map) and ∆S is


the pixel of the beamformed map. The beamforming output at each pixel B∆ S(f )
is normalized by the point spread function (PSFc ) with the source placed at the
center of each pixel. The integration region only includes the noise source, which is
identified by the beamforming sound map. The integration area at higher frequency
is reduced in size to eliminate the irrelevant side lobe. The details of the integration
region will be given in Section 6.3.2.

3.2.4 Wall Pressure Measurement

Mean and fluctuating wall pressure data were obtained with 0.5 mm I.D. pressure
taps. For the flat plate cases (no step), 28 pressure taps were used in streamwise
direction with the first located at x = −15h and the last located at x = 50.4h.
Figure 3.8 shows a cross section of the pressure tap arrangement. Countersunk 0.9
mm holes were drilled on both the endplate and steps, and a thin metal tube with
0.8 mm O.D. and 0.5 mm I.D. was flush inserted into each hole. Cynoarylate was
applied to ensure the gap between the metal tube and the countersunk hole on the
wall is sealed. A flexible tube (0.81 mm I.D. and 1.2 mm O.D.) was connected to the
other side of the metal tube. Where this tube is connected depends on whether mean
or unsteady pressure is measured. These techniques will be explained in Section 4.3
and Section 4.4, respectively.
Chapter 3. Methodology 30

Metal probe I.D. = 0.5 mm

0.9 mm countersunk

Wall Wall

O.D. = 0.8 mm
Cynoarylate

100 ~200 mm long


flexible tube

Figure 3.8: Cross-sectional view of the pressure tap.

Flat Plate Testing Cases

Figure 3.9 shows the distribution of the pressure taps used for the flat plate cases.
The streamwise measurement points are located on the step spanwise centerline
(z = 0) and the measurements were taken between -15 h to 50.4 h. In the spanwise
direction, the mean wall pressure measurement considered pressure taps from z = 0
to z = 6h to ensure the flow two-dimensionality and the fluctuating pressure was
only measured from z = 0 to z = 2h since the spacing for z = 6h is too large to
resolve the spanwise structure.

Spanwise test locations

Streamwise test locations


(at the center of the step z=0)

Step leading edge


location

Figure 3.9: Wall pressure measurement locations for flat plate cases. Red dashed line
indicates step leading edge location.
Chapter 3. Methodology 31

Step Testing Cases

Figure 3.10 displays the pressure tap distribution for Step1 case. The blue area is
the step top surface, and the leading edges always align with x = 0 for all aspect
ratios. The grey area here and in Fig. 3.9 is the slidable area shown in Fig. 3.3
which moves in the streamwise direction when the different steps were installed so
that the first pressure taps in this area always stay the same distance to the step
trailing edge (black dashed line). Thus, the upstream testing points (white area)
and the step downstream testing points (grey area) remain the same distance to the
leading edge (red dashed line) and the trailing edge (black dashed line), respectively,
regardless of the different lengths of each step.
For all step cases the streamwise measurement points are distributed from -15
h upstream of the step leading edge (Fig. 3.10 red line) to 50 h downstream of the
step trailing edge (Fig 3.10 black line) with different numbers of the pressure taps
on the top surfaces of the various aspect ratio steps. Spanwise unsteady pressure
measurements were only considered within 2 step heights as previous studies [8, 39]
suggested that the spanwise coherent length scale is within 1-2 step heights. Mean
wall pressure measurements in the spanwise direction were only taken for the flat
plate case to ensure the flow two-dimensionality. Note that in both the streamwise
and spanwise direction, the testing locations are not equally spaced. There are
greater number of measurement points inside the separation/reattachment region as
the flow features change rapidly in these regions and require higher spatial resolution.
The test stations on the front face and back face of the steps (one pressure tap
on each vertical face, for setup see Fig. 3.11) are located at z/h = 6, y/h = 0.5.
Three other testing points are positioned at x = −0.5h, x = 0.5l (half step length),
and x = 6h from the trailing edge, respectively. Since all of the testing points
downstream of the black dashed line in Fig. 3.10 are inside the grey movable area
in Fig. 3.3, the tap locations remain the same for all cases. Details of pressure taps
locations for all step cases can be found in Appendix A.
Chapter 3. Methodology 32

Figure 3.10: Wall pressure measurement locations for AR = 1 step. Red dashed line
indicates step leading edge, black dashed line indicates step trailing edge.

x
0.5 ℎ
Metal probe
2 mm
Wall

Flexible tube
Figure 3.11: Pressure tap on the vertical faces of the step.

Mean Wall Pressure

The mean wall pressure was measured for all step in order to understand the pressure
distribution across the flow field. In this measurement, the flexible tubes were
connected to a 64-port (1.6 mm O.D.) DTC Initium Pressure Scanning System as
shown in Fig. 3.12. One of the ports was used to measure the anechoic chamber
ambient pressure p∞ so that the pressure coefficient Cp = (p − p∞ )/0.5ρV 2 can be
calculated.
The mean pressure measurements were taken at both streamwise pressure tap
locations and on the vertical faces of the step (Fig. 3.11). The spanwaise mean pres-
sure was only measured in flat plate (no step) case to ensure flow two-dimensionality.
Chapter 3. Methodology 33

0.8 mm to 1.6 mm
Adaptor
1.6 mm
0.8 mm
Flexible tube
Flexible tube
Measure ambient
pressure

DTC Initium Pressure Scanning System (64-port)

Figure 3.12: Mean wall pressure measurement setup.

Fluctuating Wall Pressure Measurement

A common way to measure fluctuating wall pressure is via transducers directly flush
mounted into the wall. However, the physical spacing between two measurement-
points will be restricted by the size of transducers and the smallest length scale
thatcan be resolved is limited. To increase spatial resolution, the remote micro-
phone method [4, 33] was used. Instead of mounting the transducers directly under
the wall, the transducers are remotely located and connected to plastic flexible tubes
through which the pressure fluctuations transmit. Since the pressure fluctuations
are attenuated while travelling inside the flexible tube, the length of the tubes are
generally set to be around 10 ∼ 20 mm long. This is long enough to reach the
transducer and provides reasonable resolution in the high frequency region.
Figure 3.13 shows the setup of the remote microphone measurement. The stiff
metal probe with 0.5 mm I.D. is flush mounted to the endplate/steps surface and
is connected to the remote microphone housing using the flexible tubing. Inside
the remote housing, a small channel branches into two, one directs the pressure
fluctuations to a 1/4” GRAS 40PH microphone where they are measured and the
other leads to a 2 m long flexible tube. The long flexible tube is seen as an acoustic
termination since the pressure wave will dissipate as it travels down the tube such
Chapter 3. Methodology 34

that reflection inside the housing can be minimised.

Stiff metal probe


Remote microphone housing

2 m Long acoustic
Flexible tube termination

1/4” GRAS 40PH phase


matched microphone

Figure 3.13: Remote microphone setup.

Since the acoustic wave will attenuate or resonate when it is transferred inside
the tube, each pressure port needs to be calibrated before the measurement. In this
study, a two-step hybrid calibration method proposed by Awasthi et al.[7] was used.
The calibration process calculates the frequency response of each tube by comparing
the remote microphones to a factory-calibrated reference microphone at the surface
that is exposed to the same noise source. The transfer function (in Hz) of each tube
can be calculated as
φSM (f )
Hcal (f ) = (3.8)
φSR (f )SRef

where φSM (f ) is the cross-spectrum between the remote microphone output voltage
and the speaker input voltage signal, φSR (f ) is the cross-spectrum between the
reference microphone output voltage and the speaker input voltage signal, and SRef
is the sensitivity (V/Pa) of reference microphone.
Chapter 3. Methodology 35

Cavity Calibration Free-Field Calibration


Speaker

Clibration Reference
Coupler Speaker
Mic.
Reference
Mic.

Cavity
2 mm

Metal tube

(a) Cavity calibration setup (b) Free-field calibration setup

Figure 3.14: Sketches of calibration methods in two-step remote microphone calibration


[7]

The two-step calibration method proposed by Awasthi et al. [7] includes both
cavity calibration and free-field calibration(Fig. 3.14 [7]). In the cavity calibration,
a calibration coupler was used to insulate the ambient noise. The calibration coupler
is a hollow cylinder where the speaker (noise source) is placed on one end and the
pressure tap on the other end, and the reference microphone is mounted on the side
(see Fig. 3.14(a)). With this calibration coupler, the cavity calibration performs well
at low frequency region; however, resonance will occur when the acoustic wavelength
is associated with the cavity geometry (Fig. 3.15 (a)). Whereas for the free-field
calibration, the reference microphone and the pressure tap are directly exposed
to the speaker, no resonance is seen. Yet the low frequency region is generally
contaminated due to scattering effects. A hybrid calibration curve was therefore
created by combining the calibration curves obtained from two methods. In Fig. 3.15
(a) there is an overlap frequency region where both two methods are applicable. A
frequency in that region is chosen such that the calibration curves can be smoothly
combined. The free-field calibration curve is adopted above the cutoff frequency,
while below it, the cavity calibration curve is selected. The comparison between the
raw data, the calibrated data for the turbulent boundary layer on the flat plate (no
step) at a flow speed of 20 m/s is shown in Fig. 3.16.
Chapter 3. Methodology 36

Low frequency
contamination Overlap
region

Geometry associated
resonance
(a)

10-4

10-5 Free- field


, V /Pa2

10-6
2 2
|H(f)|

10-7

10-8 Cavity

10-9
101 102 103 104
Frequency, Hz
Cutoff
Frequency
(b)

Figure 3.15: Calibration curves for the pressure tap at (x, y) = (−40, 0) at a free stream
velocity of 20 m/s. (a) Transfer function yielded from both Cavity calibration and Free-
field calibration. (b) Combined calibration curve, cutoff frequency = 3192 Hz.
Chapter 3. Methodology 37

Frequency, Hz

Figure 3.16: Calibrated and Uncalibrated pressure spectrum at (x, z) = (−40, 0) at a free
stream velocity of 20 m/s

An empirical model of surface pressure fluctuations underneath a zero pressure


gradient turbulent boundary layer was proposed by Goody [30]. The model is given
by :  2
ωδ
C2
φ(ω)U∞ U∞
= " #3.7 #7 (3.9)
τw2 δ 3/4 " 
ωδ ωδ
+ C1 + C3
U∞ U∞

where φ(ω) is the power spectral density of surface pressure fluctuations (Pa2 /Hz)
τw is the wall shear stress ( 12 ρU∞
2
Cf ), C1 = 0.5, C2 = 3.0 and C3 = 1.1RT−0.57 . RT is
the function of Reynolds number based on displacement thickness θ, defined as

 3/4
U∞ θ
RT = 0.11 . (3.10)
ν

The calibrated pressure spectra of the undisturbed boundary layer at different


streamwise locations on the flat plate (without step) are plotted in Fig. 3.17. The
measurement shows a good agreement with Goody’s model (calculated with flow
parameters in Table 3.2), which shows that the remote microphone technique can
accurately measure the unsteady wall pressure fluctuations.
Chapter 3. Methodology 38

100

10-1

10-2

10-3

10-4

10-5

100 101 102

Figure 3.17: Flat plate unsteady pressure spectra for U∞ = 20 m/s at different streamwise
locations compared with Goody’s model [30].
Chapter 3. Methodology 39

3.3 Oil Flow Visualisation Technique

The step upstream separation length xs , top-surface reattachment length xr1 , and
downstream reattachment length xr2 were measured in the oil flow visualization
experiment.

3.3.1 UNSW Educational Wind Tunnel

Oil flow visualization was conducted in the EWT. Figure 3.18 is a schematic diagram
of EWT. This facility is an open-loop wind tunnel that consists of a 0.3 × 0.3 ×
0.61 m test section with a maximum flow speed of 65 m/s. The air flow is driven by
an axial suction fan and is rectified by a honeycomb flow straightener at the inlet
of the tunnel. More details about this facility can be found in [16].

Honeycomb Flow Straightener

Axial Suction Fan


Forward-backward 0.3 x 0.3 x 0.61m
facing step Test Section Diffuser

Oil mixture(white)
Contact paper (dark)

4.6 m
Contraction (bell mouth) Section
Figure 3.18: Schematic diagram of UNSW EWT and oil flow visualization setup.

3.3.2 Oil Flow Visualization

The undisturbed boundary layer information in UNSW Educational Wind Tunnel


was first acquired from a pitot tube experiment and the boundary layer thickness at
the step leading edge is given in Table 3.3. The steps used in this experiment were
different from those used in the UAT experimental campaign (described in Section
Chapter 3. Methodology 40

3.2.3) as the undisturbed boundary layer thickness is different in the two facilities.
Step of AR = 3, 4, 6, and 8 were considered at a free stream velocity U∞ = 35
m/s in this experiment. Sherry et al. [69] and Awasthi et al. [5] suggested that
the reattachment length for step geometry is dependent on the Reynolds number
(Reh ) and the step height to boundary layer thickness ratio (h/δ) when the step
is immersed in the boundary layer. The undisturbed boundary layer thickness is
9.5 mm in EWT at U∞ = 35 m/s, therefore in this measurement the steps height
is selected as 5 mm, which results in a comparable Reh (1.25 × 104 ) and h/δ ratio
(52.6%) to all the other cases in present study.

Table 3.3: UNSW EWT boundary layer information.

U δ δ∗ θ
Reh h/δ H Reδ Reδ∗ Reθ
(m/s) (mm) (mm) (mm)
35 12318 9.5 0.526 0.82 0.71 1.17 23405 2022 1736

A mixture of vegetable oil, Kerosene, and talcum powder with 1:13:8 volume
mixing ratio was applied to the black contact paper that was used to cover the step
and the region upstream and downstream of the step. After the flow is turned on, the
mixture takes about 180 seconds to dry and display the flow features. Figure 3.19
shows the flow visualization image for the AR = 4 step. The red dashed lines indicate
the leading edge and the trailing edge of the step and xs is the upstream separation
length. Due to the unsteady nature of the flow, the reattachment locations were
labelled as regions (∆xr1 and ∆xr2 ) where splashing patterns can be seen at their
bounds and the time-mean reattachment lengths were measured at the center of
these regions. The tests were performed multiple times for each case and the results
were similar. Note that xs and xr1 were measured from the leading edge while xr2
was measured from the trailing edge as the step length varies in the different cases.
Chapter 3. Methodology 41

Flow

xs xr2

xr1 xr1 xr2


Figure 3.19: Oil flow visualization result for AR = 4 step at U∞ = 35 m/s. Flow is coming
from left to the right. Yellow solid lines indicate the boundary of the separation/reattach-
ment regions. Yellow dashed lines indicate the center of the reattachment regions. Red
dashed lines indicate the step leading edge and trailing edge.
Chapter 3. Methodology 42

3.4 Numerical Simulation

Computational Fluid Dynamics (CFD) simulation was applied to understand the


velocity field of the flow. Instantaneous flow quantities such as the fluctuating
velocity and the fluctuating wall pressure were provided by the CFD calculation.

3.4.1 Flow Domain and Turbulence Model

Step aspect ratios AR = 1, 2.4, 4 and 8 were considered in the CFD simulation in
order to make comparisons with experimental results. The step heights were set to
be 5 mm and the free stream velocity was 35 m/s in every cases which leads to Reh
= 1.2 ×104 . As the corresponding Mach number is low (M =0.103), the flow can be
considered as incompressible. Large-eddy simulation (LES) was adopted to resolve
the incompressible turbulent flow field. In LES, the governing equations are derived
by applying a spatial filter for small eddies instead of time averaging for fluctuating
terms as is done in RANS, which makes the instantaneous quantities available above
the grid cut-off frequency [10].
The filtered flow velocity components ūi in an incompressible turbulent flow are
governed by the continuity equation (Eq. 3.11) and the Navier-Stokes equation (Eq.
3.12) :
∂ ūi
= 0, (3.11)
∂xi

and
∂ ūi ∂ ∂σij p̄ ∂τij
ρ +ρ (ūi u¯j ) = − − (3.12)
∂t ∂xj ∂xj ∂xi ∂xj

where τij is the subgrid-scale stress and σij is the molecule viscosity stress tensor.
They are defined by
τij ≡ ρui¯uj − ρūi u¯j , (3.13)

and "  #
∂ ūi ∂ u¯j 2 ∂ ūl
σij ≡ µ + − µt δij . (3.14)
∂xj ∂xi 3 ∂xl
Chapter 3. Methodology 43

The eddy-viscosity µt in Eq. 3.14 remains unknown and needs to be modeled. The
Wall-Adapting Local Eddy-Viscosity (WALE) Model proposed by Nicoud[58] was
applied in the present study as it is found returning better prediction for wall-
bounded flow. In WALE model the eddy-viscosity is modeled by

(Sijd Sijd )(3/2)


µt = ρL2s (3.15)
(S¯ij S¯ij )5/2 + (Sijd Sijd )5/4

where Sij is the rate-of strain tensor, given by

 
1 ∂ ūi ∂ u¯j
Sij ≡ + . (3.16)
2 ∂xj ∂xj

Sijd is the traceless symmetric part of the square of the velocity gradient tensor[58]
and Ls is the mixing length for subgrid scales determined by

Ls = min κd, Cw V 1/3



(3.17)

where κ is Karman constant, d is the distance to the closest wall, Cw is the WALE
constant, and V is volume of a cell in the computational domain.

3.4.2 Boundary Conditions

Figure 3.20 shows computational domain of the step cases. A separate RANS simu-
lation of the flow over the flat plate (without step) was first performed by adopting
the k − ω SST turbulence model. This was done to generate the turbulent boundary
layer velocity profile, which is then imported as the inlet velocity profile in the step
cases. A constant total pressure is applied at the outlet boundary. A periodical
boundary condition is used in the spanwise direction and a no-slip boundary condi-
tion is used on the walls. A symmetry boundary condition is employed at the top
of the domain.
Chapter 3. Methodology 44

Figure 3.20: Schematic diagram of the computational domain and the boundary condition
employed in the simulation.

3.4.3 Numerical Procedure and Computational Grid

The governing equations were solved using Ansys Fluent 19.1 finite-volume solver.
Second-order schemes were used for both time and space discretisation. The velocity-
pressure coupling was conducted with the SIMPLE method. The calculations were
performed on the Australia National Computational Infrastructure (NCI)’s super-
computer, Gadi.
The computational domain size is Lx = 60h, Ly = 30h, Lz = 5h in the stream-
wise (x), wall normal (y), and spanwise (z) directions, respectively. A similar domain
size has been found to be adequate in previous step simulations [2, 39, 48]. The step
leading edge locates at 20h downstream from the inlet. A structured mesh was em-
ployed in the simulation. The meshing is shown in Fig. 3.21. Mesh refinement was
employed at around the step and wall. The grid size in each dimension is determined
by ∆x+ , ∆y + , and ∆z + . The ∆x+ in streamwise direction varied from 5 to 50 with
the smallest mesh created at the vertical faces of the step. The ∆y + is set to be 1
Chapter 3. Methodology 45

on all horizontal surfaces and grows to 40 at y / h ∼ 3. The ∆z + value is a constant


at 25 in the spanwise direction. The number of total cells is around 16.5 million. A
grid convergence study was performed to confirm the mesh refinement is sufficient
(details are given in Section 3.4.4). The time step size was 8 × 10−7 seconds with the
corresponding CFL number < 1 throughout the entire computational domain. The
data sampling time was 4.8 × 10−6 second with the total sampled length equivalent
to 117.6 flow through time after the flow reached qusai-steady state.

Figure 3.21: Computational domain meshing. Side view from the spanwise direction.

3.4.4 Grid Convergence Study

The grid independence study was performed to AR = 4 step to demonstrate grid


convergence. Since the step geometries are similar, the grid convergence study result
were applied to all testing cases. The grid convergence index (GCI) was calculated to
estimate the spatial discretization error [65]. Three different grid resolutions with
mesh refinement concentrated on the wake region and in the vicinity of the step
were considered. The aforementioned ∆y + and ∆z + value remain the same in all
three cases. The grid refinement ratio was estimated by r = (ki /ki+1 )1/3 , where k is
the number of cells in each case (k = 6.2, 13.5, 24 million) and i = 1, 2, 3 represents
the coarse mesh, medium mesh and fine mesh, respectively.
Chapter 3. Methodology 46

The r.m.s. wall pressure at x = 3h downstream of the leading edge and the mean
velocity magnitude in the wake region at x = 6h and y = 0.5h were chosen to eval-
uate grid convergence. Both values are averaged along the spanwise direction. The
results are shown in Table 3.4. The parameter GCI3,2 and GCI2,1 are the grid con-
vergence indices for medium-to-fine and coarse-to-medium refinements, respectively.
The Richardson error estimator  was calculated to quantify the discretization error
between the value predicted by the Richardson extrapolation method [64] and by the
medium resolution grid when the grid spacing approaches zero. The error estimator
 for the medium resolution grid was 1.1% and 2.47% for r.m.s. wall pressure and
mean velocity magnitude, respectively. The value of GCI3,2 was lower than GCI2,1
for both variables, which implies the flow parameters are indeed converging as the
computational cell size reduced. The medium resolution grid was adopted in this
study as it provides reasonable estimation of the flow properties, while requiring
lower computational cost.

Table 3.4: Grid Convergence Index for AR = 4 step

GCI2,1 GCI3,2

(coarse-to-medium) (medium-to-fine)
Velocity
34.19% 3.21% 2.47%
Magnitude
r.m.s. Wall
11.8% 1.36% 1.1%
Pressure
Chapter 3. Methodology 47

3.5 Data Processing and Analysis Methods

3.5.1 Two-point Correlation/Coherence

Two-point statistics are commonly used to understand the length scale and time
scale of turbulent structures. The spatial two-point correlation is defined as

Raa (xref , xref + ∆x) = E[a(xref )a(xref + ∆x)], (3.18)

where a is the flow variable, xref indicates the reference point, and E[ ] denotes the
P
expected value operator and is defined as E[X, Y ] = xi yi , where X = [x1 , x2 , ...xi ]
and Y = [y1 , y2 , ..., yi ]. The correlation coefficient can be calculated as

Raa (xref , xref + ∆x)


Caa (xref , xref + ∆x) = . (3.19)
Raa (xref , xref )0.5 Raa (xref + ∆x, xref + ∆x)0.5

The combination of spatial and temporal correlation provides information about the
convection of the flow. This space-time correlation was applied to the unsteady wall
pressure measurements data to understand the near-wall flow behaviours. Since the
fluctuating pressure p0 was measured remotely, the space-time correlation coefficients
were computed taking the inverse Fourier transform of the calibrated spectra [5].
The Fourier transform of Eq.3.18 yields the cross spectrum Φaa , where

Z ∞
Φaa (xref , xref + ∆x; f ) = Raa (xref , xref + ∆x; t)e−i2πf t dt. (3.20)
−∞

The coherence between two points xref and xref + ∆x in frequency domain γ can
then be calculated as the magnitudes of cross spectrum normalized by the square
root of the respective auto spectra.

|Φaa (xref , xref + ∆x, f )|


γ(xref , xref + ∆x, f ) = .
|Φaa (xref , xref , f )|0.5 |Φaa (xref + ∆x, xref + ∆x, f )|0.5
(3.21)
Chapter 3. Methodology 48

3.5.2 Modal Analysis Methods

Modal analysis methods were applied in the present study to extract the energetic
coherent structures from the flow data. These techniques are widely used in reduced-
order modelling, flow control and data compression/reconstruction, where the most
dominant modes are identified and selected to represent the stochastic flow field.
Proper Orthogonal Decomposition (POD), proposed by Lumley [52], is among
the first modal analysis for fluid dynamics. Standard (space-only) POD seeks a set
of orthogonal basis functions that best approximates the stochastic terms (mean-
subtracted) based on the sampled data set. On each axis of the POD basis, the
variance is minimized and the covariance between axes should ideally be zero (or-
thogonal). Consider a time series of flow quantity data Q at various spatial locations.
Here the flow variable can be expended as


X
Q(x, t) = Q(x) + q, q= aj (t)φj (x) (3.22)
j=1

where Q(x) is the time-mean value and q is a stochastic term. These stochastic
terms can be decomposed into modes (φj ) and the associated coefficient aj , where
j is the mode number.
In this study, the snapshot-based method was used to calculate the POD modes.
The data matrix q (mean removed) at a time instant tk can be written as

 
 q(x1,1 , tk ) q(x1,2 , tk ) . . . q(x1,n , tk ) 
 
 q(x2,1 , tk ) q(x2,2 , tk ) . . . q(x2,n , tk ) 
q(x, tk ) =  (3.23)
 
.. .. ... .. 

 . . . 

 
q(xm,1 , tk ) q(xm,2 , tk ) . . . q(xm,n , tk )
Chapter 3. Methodology 49

, where this can be reshaped into a long single column vector:


 
 q(x1,1 , tk ) 
 
 q(x , t ) 
 1,2 k 
..
 
.
 
 
q(x, tk ) = 
 
 (3.24)
 q(x2,1 , tk ) 
 
 .. 

 . 

 
q(xm,n , tk )

The long column vector in Eq. 3.24 is called a snapshot of the data at tk . The
flow quantity data q for entire time series can therefore be written as an M × N
matrix (Eq.3.25) where M is the total number of the spatial sampling points and
N is the number of time steps.

 
 q x1 
 
.. .. ..
. . .  
 qx 

   2
qM ×N q1 q2 . . . qN  ,
=  q1 =  .  (3.25)
. .  .. 
.. .. ..   
.  
qxM

The POD modes are obtained by performing an eigendecomposition of the


correlation matrix qT q:
qT qΦ = λΦ. (3.26)

The eigenvector Φ = [φ1 , φ2 , . . . , φj ] contains the POD modes (orthogonal basis)


q. The descending-order sorted eigenvalues λ = [a1 , a2 , . . . , aj ] correspond to the
energy that each mode contains.
The standard POD method provides modes which optimally represent spatial
correlations within the sampled region in the flow field. The higher energy of the
mode, the better correlated and more dominant the structure is.
The standard (space-only) POD sees each snapshot as an individual data set,
therefore the sequence of these snapshots is irrelevant. Although it extracts the
Chapter 3. Methodology 50

dominant spatial coherent structures, the evolution of these structures in time is


missing. To capture the temporal dynamics of the modes, Spectral Proper Orthog-
onal Decomposition (SPOD) was introduced [70]. Instead of directly computing the
eigendecomposition in Eq.3.26, SPOD decomposes the cross-spectral density matrix
(CSM) of q at each frequency.
By taking the cross-Fourier transform of each component of q (Eq. 3.27), the
CSM at frequency fk can be obtained by using the complex conjugate and a spatial
weighting function W (Eq3.28).

   
 q1 (t1 ) . . . q1 (tN )   q̂11 (fk ) . . . q̂1M (fk ) 
 . .. ..  FFT  . .. .. 
q= . . ===⇒ q̂fk = . . , (3.27)
 . .    . . 
   
qM (t1 ) . . . qM (tN ) q̂M 1 (fk ) . . . q̂M M (fk )

Cfk = q̂∗fk W q̂fk . (3.28)

The eigendecomposition of the CSM at frequency fk is,

Cfk Θfk = Λfk Θfk (3.29)

where the eigenvector Θfk are the SPOD modes and the eigenvalue Λfk are the corre-
sponding modal energies at discrete frequency fk . These modes usually demonstrate
the dynamic physics of the flow field such as vortex shedding, flow instability, and
so on.
Chapter 3. Methodology 51

3.6 Summary

The experimantal and numerical methods were discussed in this chapter. The step
geometries considered in these measurements are AR = 1, 2, 4, 8 with the step height
to boundary layer thickness ratio (h/δ) ∼ 0.5. The far-field acoustic data were
collected by 64-microphone phased array.
Surface pressure measurements were carried out using pressure taps on the sur-
face of the wall and the steps. The mean wall pressure were measured using a DTC
Initium Pressure Scanning System. The unsteady wall pressure data were obtained
by the remote microphone technique[33, 44]. The two-step remote microphone cal-
ibration method [7] was applied to the measured pressure spectrum to remove the
attenuation and resonance caused by the extended flexible tubes.
Oil flow visualization was conducted in the UNSW educational wind tunnel.
Steps with aspect ratios = 3, 4, 6, 8 and (h/δ) ∼ 0.526 were considered in this
experiment. An oil-kerosene-talcum powder-mixture was applied about the step that
immersed in the flow. The averaged upstream separation lengths and downstream
reattachment lengths were obtained by examining the pattern displayed after the
oil mixture drys out.
The CFD simulations were performed in the present study to further understand
the flow field features of the FBS. Steps with AR = 1, 2.4, 4, 8 and (h/δ) ∼ 0.5 were
implemented in the numerical calculation. LES was used to compute the flow field
and the WALE model was used to take account of the unresolved small scale struc-
tures. The numerical model were validated using the experimental results obtained
from the aforementioned wall pressure measurements in this study. Modal decom-
position methods such as POD, SPOD were introduced to analyze the stochastic
turbulent flow field. The technique were applied to the velocity field computed by
CFD simulations to capture the coherent structures existing the turbulent flow.
Chapter 4

Oil Flow Visualization and Wall Pres-


sure Measurements

4.1 Overview

This chapter provides the results obtained from the oil flow visualization and the
wall pressure measurements. In Section 4.2, the reattachment lengths measured
from the oil flow visualization experiments are compared with past studies. The
steady and unsteady wall pressure measurement results are presented in Section
4.3 and 4.4, respectively. Section 4.4.2 shows the fluctuating wall pressure spectra
measured at different streamwise locations and indicates the interactions between
the flow and the step. The data were further processed to attain the turbulent flow
information, such as convection velocity (Section 4.4.3) and the spanwise coherence
length (Section 4.4.3).

52
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 53

4.2 Oil Flow Visualization Results

The upstream separation and downstream reattachment lengths (see Fig. 4.1) are
given in Table 4.1. The upstream separation lengths xs are −1.5h to −1.6h for all 4
steps. These values are comparable to −1.5h ∼ −1.9h [5], −1.27h [31] and −1.57h ∼
−1.96h [40], but slightly longer than −0.8h to −1h reported by Ref. [13, 24, 47, 49]
for flow over a FFS or FBS with different h/δ ratio and Reh (see Fig. 4.2). It is
worth noting that regardless the different h/δ ratio and Reh considered in previous
studies, the upstream separation lengths, xs , are generally between -2 to -0.8h for
FFS and FBS scenario. Pearson et al. [62] suggest that the upstream separation
bubble characteristics are modulated by upstream incoming flow structures.

xr1
xs xr2
Figure 4.1: Separation length xs and reattachment lengths xr1 & xr2 .

Table 4.1: Oil flow visualization results

Step
Step U∞
Height Reh h/δ xs /h xr1 /h xr2 /h
AR (m/s)
h (mm)
3 35 5 12,318 0.526 -1.5 1.4 6.26
4 35 5 12,318 0.526 -1.6 1.65 5.86
6 35 5 12,318 0.526 -1.6 1.6 5.95
8 35 5 12,318 0.526 -1.6 1.8 5.9
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 54

3.5

2.5

1.5

0.5

0
10 2 10 3 10 4 10 5

Figure 4.2: Upstream separation length xs of FBS compared with past FFS studies.

In general, two recirculating regions can be observed downstream of the FBS


leading edge, one of which locates on top of the step and the other settles downstream
of the step. The lengths of upstream and downstream recirculating regions are
denoted as xs , xr1 and xr2 , respectively (Fig. 4.1). Due to the unsteady nature
of turbulence, these lengths were determined using the center of the reattachment
region ∆xr . An example image can be found in Section 3.3.
For the recirculating region on top on the FBS, the flow reattached 1.4h-1.8h
away from the leading edge in this measurement. This is comparable to 1.6h found by
Fang and Tachie [24] who considered a FBS with 2.36 l/h and 0.154 h/δ. Although
the h/δ of 0.154 in [24] is smaller than 0.526 in present measurement, the Reynolds
numbers based on step height are close (Reh =13200 in [24] and 12318 in current
study) and results in a similar reattachment length. Some previous studies [5, 40, 69]
investigated the sub-boundary layer (h/δ < 1) FFS and found that the reattachment
lengths vary with Reh . Figure 4.3 shows the comparison of current results and other
studies (h/δ < 1) [5, 24, 26, 40, 69]. It can be seen that in studies [69] and [5] the
reattachment length from the leading edge increase with Reh and the value roughly
settles between 3-4h when Reh >15000. It can also be observed that for the FBS
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 55

cases (current data and Fang and Tachie [24]) the reattachment lengths on the step
top surface (xr1 ) is around 50% of the FFS cases under the similar Reh . The reason
for these different values of xr1 in FFS and FBS is unclear.

0
10 3 10 4 10 5 10 6

Figure 4.3: Reattachment length on the top surface (xr1 ) of FBS compared with past FFS
studies.

Downstream of the step, the second reattachment length xr2 was measured to be
5.9 - 6.26 h behind the step trailing edge in this experiment. As the aspect ratios
of the step (AR = 3, 4, 6, 8) regarded in this measurement are all larger than the
first reattachment length xr1 , the downstream flow preserves some BFS flow field
characteristics. Figure 4.4 compares the present FBS data with previous studies of
BFS [14, 26, 40, 48, 73]. It can be seen that the xr2 in the current measurement is
comparable to other BFS studies despite the leading edge disturbance. Figure 4.4
also shows that the reattachment length is weakly dependant on Reh in the BFS
step scenario.
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 56

0
10 3 10 4 10 5

Figure 4.4: Reattachment length downstream of the FBS (xr2 ) compared with past BFS
studies.

As some shorter steps (AR = 1, 2) are considered on wall pressure and acoustics
measurement in the present study, it is important to understand how the down-
stream reattachment length xr2 varies with the aspect ratio. Figure 4.5 shows the
second reattachment lengths xr2 observed for different aspect ratio FBS. It can be
noticed that at the small AR (AR < 3 − 4), xr2 varies nearly linearly with the step
aspect ratio where the shorter steps tend to have a larger downstream recirculating
region. On the contrary, for AR > 3 − 4, the value of xr2 plateaus at 4 − 6h and
does not change drastically with aspect ratio. The different behaviours in two area
is likely due to the recirculating region on top of the FBS. For the longer steps
(AR > xr1 ), the separated flow from the leading edge reattaches to the top surface
of the FBS and re-develops into a new boundary layer; therefore the downstream
flow is able to retain BFS features which stabilizes the second reattachment length
xr2 . On the other hand for the AR < xr1 steps, since the recirculating region can not
form on top of the step, the suction from this negative pressure area is smaller (see
Section 4.3). This allows the separated flow to travel further downstream. Bergeles
& Athanassiadis [12] remarks that the second reattachment length (xr2 ) varies lin-
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 57

early with the step length l when step length is shorter than the first reattachment
length. They found that after AR > 4, xr2 remains constant. It should be noted
that this critical aspect ratio is determined by the first reattachment length xr1 ,
which may differ with different flow parameters. In the present measurement, xr2 is
similar in all testing cases as the smallest step (AR=3) is still long enough to keep
the first reattachment point (xr1 /h=1.4-1.8) on the top of the step.

20

18

16

14

12

10

0
0 2 4 6 8 10 12

Figure 4.5: Reattachment length downstream of FBS compared with past BFS studies.

4.3 Mean Wall pressure

4.3.1 Smooth Wall

Mean wall pressure statistics were collected by the pressure taps installed on the wall.
The instrumentation of the measurement is given in Section 3.2.4. The wall pressure
coefficient Cp is calculated to understand the mean wall pressure distribution of the
flow field. It is given as
p − p∞
Cp = 1 2
(4.1)
2
ρU∞
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 58

1 2
where 2
ρU∞ is the dynamic pressure, p is the mean pressure measured at each
pressure taps and p∞ is the ambient pressure in the anechoic chamber. The mean
wall pressure of the undisturbed flow (no step) in streamwise and spanwise directions
at three different flow speeds (20, 30, 40 m/s) are given in Fig. 4.6 and Fig. 4.7. In
streamwise direction, the Cp values are mostly between 0 to 0.01 with the largest
value of 0.019 measured at the first pressure tap downstream of the leading edge
location (x = 0). This is due to the inevitable minor discontinuity between panels.
The result shows that a good zero pressure gradient was achieved under different
flow speed throughout the testing section.
Figure 4.7 shows the results in spanwise direction at two streamwise locations
(x = −0.5h andx = 6.4h, h = 10 mm is the step height ). The mean wall pressure
was measured in the spanwise direction to establish the flow two-dimensionality.
The nearly identical value of pressure coefficient across those spanwise measuring
points indicates the spanwise homogeneity and the flow is two dimensional within
the testing section.

0.5

-0.5
-20 -10 0 10 20 30 40 50 60

Figure 4.6: Streamwise mean wall pressure for undisturbed flow (no step) at z = 0. Blue
3, U∞ =20 m/s; Red ◦, U∞ = 30 m/s; Yellow 4, U∞ = 40 m/s.
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 59

0.5 0.5

0 0

-0.5 -0.5
-1 0 1 2 3 4 5 6 7 -1 0 1 2 3 4 5 6 7

(a) x = −0.5h (b) x = 6.4h

Figure 4.7: Spanwise mean wall pressure for undisturbed flow (no step) at two streamwise
locations. (a) At x = −0.5h. (b) At x = 6.4h. Blue 3, U∞ = 20 m/s; Red ◦, U∞ = 30
m/s; Yellow 4, U∞ = 40 m/s.

4.3.2 Forward-backward Facing Step Mean Wall Pressure

The mean wall pressure was measured for four different aspect ratios FBS (AR =
1, 2, 4, 8) in the streamwise direction. The wall pressure coefficient Cp distributions
are given in Fig. 4.8. Red dashed lines and Black dashed lines in the figure indicate
the leading edge and trailing edge of the step, respectively. Noted that all of the
horizontal measurements in streamwise direction were taken at the center of the
step span (z = 0); while on the vertical faces, the measuring locations at (y, z) =
(0.5h, 6h) were not aligned with other pressure taps due to the physical limitation of
the setup. In upstream region, the Cp curves follow the nearly identical pattern in all
four cases where a roll-up can be observed and the similar highest level of ∼0.36 are
found on the vertical face. The curves plunge immediately after the upper corner,
coincides with the first downstream recirculating region. Downstream of the step
trailing edge, the Cp value drops again in the second recirculation region and then
slowly returns to near zero region.
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 60

1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6
-0.8 -0.8
-1 -1
-20 0 20 40 60 -20 0 20 40 60

1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6
-0.8 -0.8
-1 -1
-20 0 20 40 60 -20 0 20 40 60

Figure 4.8: Mean pressure coefficient for four steps configuration at U∞ = 20 m/s. (a)
Step1. (b) Step2. (c) Step4. (d) Step8. The red and black vertical dashed lines represent
step leading edge and trailing edge, respectively.

For Step4 and Step8, the same peak negative value of -0.8 occurs at the first
downstream testing point (x = 0.5h). While for Step1 and Step2, the peak negative
value appears at the same location but with lower value of -0.4 and -0.65, respec-
tively. The difference is likely due to the different reattachment positions of the first
downstream separation bubble. Bergeles & Athanassiadis [12] suggest that with the
lower suction pressure on the top surface, the separating shear layer would have
higher separation angle and therefore results in a longer downstream reattachment
length. It should be noted that despite the similar Reh and h/δ in the oil flow vi-
sualization experiment and in the wall pressure measurement, they were conducted
in different facilities. Therefore the separation lengths and reattachment lengths
obtained previously are only used as an approximated reference in the analysis of
this measurement.
Figure 4.9 compares some FFS studies (h/δ between 0.41-0.6) [5, 26, 40] with the
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 61

Cp curves obtained from the current measurement. The comparison only considers
measurement points upstream of the step trailing edge, where the step geometry
resembles a FFS. It can be seen that in general the Cp data of FFS and FBS follow
a similar pattern when h/δ is comparable. A closer look at the data near the leading
edge is shown in Fig. 4.9b. The positive peak values upstream are all around 0.4
and observed at the closest testing point to step leading edge. Downstream of the
leading edge in the negative pressure area, Step4 and Step8 agree well with the
CFD results reported by Ji & Wang [40] who considered FFS with h/δ = 0.53. For
Step1 and Step2, with the incompleteness of the recirculation region, the weaker
suction pressure is found on the top surface of the step. Although Step2 is longer
than the averaged xr1 estimated from the oil flow visualization, the instantaneous
reattachment point may move further downstream occasionally [23] and therefore
present a differing pattern compared with FFS.
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 62

0.8

0.6

0.4

0.2

-0.2

-0.4

-0.6

-0.8

-1
-20 -15 -10 -5 0 5 10 15 20

(a)

0.4

0.2

-0.2

-0.4

-0.6

-0.8

-1
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5 3

(b)

Figure 4.9: Mean pressure coefficient near the step leading edge. Black dashed line indi-
cates the step leading edge location. (a) Current data compared with past studies. (b)
An enlarged view of the leading edge area.

Figure 4.10 shows the downstream Cp comparison between different FBS and
some previous studies [26, 40] who consider the BFS. It can be seen that although
the mean wall pressure distributions for Step4 and Step8 show good agreement with
the FFS scenario (upstream of trailing edge), they do not agree well with the BFS
scenorio downstream of the step trailing edge. This is due to the FBS leading
edge disturbance where the separated shear layer was formed. Despite the fact
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 63

that step lengths are too short for the re-developed boundary layer to return to
the smooth wall condition before meeting the step trailing edge, the difference can
still be observed between Step4 and Step8. This suggests that the longer step allows
more space for the reattached boundary layer relaxation process to take place, which
further influences the downstream flow features. For Step1 and Step2, because the
on-step reattachment does not occur, the upstream separated shear layer is able to
travel further and results in the longer negative Cp area. Among all of the testing
cases, the largest and smallest peak negative Cp value are spotted at Step1 and
Step8, respectively. This is, again, due to the reattachment event of the upstream
separated shear layer. If the step is longer, the Cp returns to zero within a shorter
distance downstream.

0.4

0.2

-0.2

-0.4

-0.6

-0.8

-1
0 5 10 15 20 25 30

Figure 4.10: Mean pressure coefficient downstream of step trailing edge compared with
past backward-facing step studies.
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 64

4.4 Fluctuating Wall Pressure

4.4.1 R.M.S Wall Pressure

The root mean square value of the wall pressure fluctuations were computed in
the frequency domain as the remote microphone setup may alter the time domain
signals. The r.m.s. value were calculated as

sZ
f2
Prms = Φpp df (4.2)
f1

where Φpp is the single sided wall pressure power spectrum and f1 ,f2 are the lower
and upper limit frequencies of the integral. Due to the low frequency contamination
and high frequency resolution limited by the remote microphone technique and pres-
sure tap diameter [7, 17], the frequency range presented here is 0.056 < f δ/U∞ < 14
(56 Hz to 13992 Hz). The Prms distributions at Reh = 14078 for steps are given in
Fig. 4.11. The r.m.s. pressure presented here was normalised on dynamic pressure
2
q∞ = 1/2ρU∞ .
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 65

0.12 0.12

0.1 0.1

0.08 0.08

0.06 0.06

0.04 0.04

0.02 0.02

0 0
-20 -10 0 10 20 30 40 50 60 -20 -10 0 10 20 30 40 50 60

0.12 0.12

0.1 0.1

0.08 0.08

0.06 0.06

0.04 0.04

0.02 0.02

0 0
-20 -10 0 10 20 30 40 50 60 -20 -10 0 10 20 30 40 50 60

Figure 4.11: Root mean square wall pressure fluctuations for four steps configurations at
U∞ = 20m/s. The red and black vertical dashed line represents the step leading edge and
trailing edge, respectively.

Upstream of the steps in Fig. 4.11, Prms shows an identical pattern for all step
cases where a sharp increase in r.m.s. pressure occurs in the separation region
x = −2h. Downstream of the step leading edge, the peak Prms values are found
at different locations for each step. The peak value of Prms = 0.0825 is measured
at x = 0.5h for Step1. Values around 20% higher were observed for the rest of the
cases. The peak occurs at x = 1h for Step2 and Step4 while for Step8 it is seen at
x = 1.5h.
Figure 4.12 shows the Prms of the present FBS cases compared with h/δ < 1
FFS cases [5, 26, 39]. Note that the comparison only considers the measurements
upstream of the trailing edge where the flow field resembles the FFS scenario. It can
be seen that the current data generally agrees with past experimental and numerical
studies. The offset of the peak value probably originates from the different Reh
and h/δ. Meanwhile, the location of the peaks are noticed further downstream for
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 66

the longer steps. For the FFS, the peaks are mostly occur at 2-3h dowstream of
the leading edge whereas they are found to be 1.5h for Step8 and 1h for Step2
and Step4. This is likely related to the reattachment point of the separated shear
layer where the peak is often seen slightly upstream of the reattachment point.
After the reattachment point, because of the boundary layer relaxation process, the
Prms decreases with the downstream distance. Awasthi et al. [5] suggested that the
disturbances (plotted in log-log scale) produced by the step would develop a near-
constant slope downstream of the reattachment region. The disturbances, quantified
q
as Cp02 − Cp∞02 are given in Fig. 4.13. C 0 is the root mean square pressure coefficient
p

0
Prms /q∞ and Cp∞ is the undisturbed r.m.s. pressure coefficient which in this case
is measured at the farthest upstream location x = −15h. It can be seen that the
disturbances start to decay from x ' 2h for both Step4 and Step8. The slope
of -1 is consistent with Ji & Wang [40] and close to -7/6 reported by Awasthi et
al. [5]. Assuming the decline starts soon after the reattachment point, the top
surface reattachment length (xr1 ) in this measurement can be approximated as 2h
for Step4 and Step8 ,which is somewhat longer than 1.65h and 1.8h measured for
AR = 4 and 8 steps in Section 4.2. Since the decay is not clear for Step2, the
reattachment length remains unknown. Normalising by xr1 , the peak locations for
Step4 and Step8 are estimated to be x = 0.5xr1 and 0.6xr1 , respectively. These are
similar to 0.6xr reported by Ji & Wang [40].
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 67

0.12

0.1

0.08

0.06

0.04

0.02

0
-20 -15 -10 -5 0 5 10 15

Figure 4.12: Root mean squared wall pressure fluctuations of current data near the step
leading edge compared with past forward-facing step studies.

10-1

Step2
Step4
Step8

100 101

Figure 4.13: Dependence of the excess pressure fluctuation intensity, relative to the undis-
0
turbed root mean square Cp∞ on x/h.

Downstream of the step (see Fig. 4.11), the r.m.s. pressure drops immediately
after the trailing edge and rises again reaching the peak region at the reattachment
area. Prior studies [26, 40, 48, 50] suggest that for the BFS flow field, the peak
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 68

of r.m.s. pressure fluctuations appears just slightly upstream of the reattachment


point. Due to the unsteady manner of the turbulent shear layer reattachment, the
reattachment location should be considered as a region. In figure 4.11, downstream
of the step clear peaks are hard to identify for some of the FBS steps. Instead,
the Prms is found around a peak area. This peak area can be regarded as the
reattachment region where the separated flow unsteadily impinges upon the wall.
The mean reattachment length xr2 for this measurement can then be estimated
as the average of this region to step trailing edge. The downstream reattachment
lengths are given in Table 4.2. The second downstream reattachment lengths xr2
of 5.35h and 5.1h are comparable to 5.8-5.9h measured in Oil flow visualization
(Section 4.2) for Step4 and Step8.
For Step1 the xr2 of 11.75h is similar to the values between 10-12h that are
commonly measured for a square obstacle (AR = 1) [12, 55, 68]. It can be noticed
that the for smaller steps (Step1 and Step2), xr2 is longer. This aligns with Berge-
les & Athanassiadis [12] who claimed that when the step aspect ratio < xr1 , the
downstream reattachment length varies with step aspect ratio in a negative linear
relationship. For Step4 and Step8, xr2 of 5.35h and 5.1h are comparable to 5.8-5.9h
measured in Oil flow visualization (Section 4.2) and again shows that xr2 does not
change a lot with the step AR once the flow reattaches on the top surface. Note that
only the results of Reh = 14078 is presented here as will be shown in Section 4.4.2
the estimated reattachment lengths do not seem to vary with Reynolds number.

Table 4.2: Estimated downstream reattachment lengths xr1 and xr2 at U∞ =20 m/s

U test
h/δ Reθ xr1 xr2
(m/s) case
Step1 0.5 14,078 - 11.75
Step2 0.5 14,078 - 6.5
20
Step4 0.5 14,078 2 5.35
Step8 0.5 14,078 2.5 5.1

Figure 4.14 shows the downstream Prms variations of the FBS and the pure BFS
(no upstream disturbances) [26, 40] scaled on reattachment length xr . It can be seen
that xr is a reasonable scaling which produces similar Prms curves despite the level
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 69

difference. The pressure fluctuations are in general larger for Step1 and Step2 in the
downstream region with the peak values around 65% higher than that of FFS, Step4
and Step8. Fang and Tachie [23] spotted the upstream separated shear layer (from
leading edge) and downstream separated shear layer (from trailing edge) mixing
behaviours for a AR = 2.36 FBS. They noticed that the vortex shedding from the
leading edge can directly alter the separation bubble formed downstream and even
dominate the vortex shedding from the trailing edge. This is potentially the reason
for shorter steps having higher Prms values downstream. For large aspect ratio
steps, the upstream separated shear layer is allowed to reattach on the top surface
before reaching the downstream separated layer. The relaxation of the reattached
flow enables the downstream flow field to retain the BFS flow features to a certain
degree. The Prms of Step4 and Step8 shows a similar pattern to the BFS but with
a higher value downstream of the reattachment region.

0.09

0.08

0.07

0.06

0.05

0.04

0.03

0.02

0.01

0
0 2 4 6 8 10

Figure 4.14: Root mean squared wall pressure fluctuations of current data near the step
trailing edge compared with past backward-facing step studies.
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 70

0.12 0.12

0.1 0.1

0.08 0.08

0.06 0.06

0.04 0.04

0.02 0.02

0 0
-20 -10 0 10 20 30 40 50 60 -20 -10 0 10 20 30 40 50 60

0.12 0.12

0.1 0.1

0.08 0.08

0.06 0.06

0.04 0.04

0.02 0.02

0 0
-20 -10 0 10 20 30 40 50 60 -20 -10 0 10 20 30 40 50 60

Figure 4.15: R.m.s. wall pressure fluctuations at three different Reh . (a) Step1. (b)
Step2. (c) Step4. (d) Step8. The red and black vertical dashed line represents the step
leading edge and trailing edge, respectively. Blue ◦, Reh = 14, 078 (U∞ =20 m/s) ; Red
, Reh = 21, 116 (U∞ =30 m/s); Yellow 3, Reh = 28, 155 (U∞ =40 m/s).

Figure 4.15 compares the r.m.s. pressure measured at 3 different Reh (14,078,
21,116, and 28,155). It can be seen that for all step cases the normalized values
at different Reh are nearly identical, which indicates the wall pressure fluctuations
demonstrate similar features at different incoming flow speed. For this reason, the
following discussion will be mainly focused on Reh = 14, 078 cases.

4.4.2 Wall Pressure Spectrum

Upstream

The unsteady wall pressure spectra in the step upstream region compared with the
undisturbed boundary layer in Fig. 4.16. The plots are shown in dimensionless scale
where pressure spectra Φpp and frequency were non-dimensionalised by free stream
velocity and undisturbed boundary layer thickness at x = 0. Figure 4.16 shows the
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 71

wall pressure spectra at 4 different streamwise locations upstream of the step leading
edge from x = −15h to x = 0.5h.
At x=-15h, the spectra are identical to that found in an undisturbed boundary
layer. Moving downstream toward the step at x = −4h (Fig. 4.16(b)), it can be seen
that the effect of the adverse pressure gradient induced by the steps appears. The
spectral level for all step cases experience a decrease at high frequency compared
with the undisturbed boundary layer. This is due to the lifting up of the boundary
layer where the small flow structures that account for the high frequency compo-
nents move away from the wall. Near the separation location x = −2h, an increase
in level in the low-frequency region (< 1 kHz) can be observed. The separation of
the boundary layer introduces large turbulent components and enhances the lower
frequency components. Inside the upstream separation bubble at x = −0.5h, the
spectral level at low frequency region is further increased. It is speculated that the
slow-moving recirculating bubble enclosed by the separated layer mainly contributes
to this low-frequency level enhancement. This large scale structure raises the spec-
tral levels at lower frequencies around 30 to 100 times that of the undisturbed
boundary layer.
In Fig. 4.16(d), a decline with the slope of -5/2 in mid-frequency (f δ/U∞ >
0.256) can be observed for all steps cases. The same slope value was also found
inside the FFS upstream recirculating region by Awasthi et al. [5].
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 72

10 -2 10 -2

10 -4 10 -4

10 -6 10 -6
Step1
Step2
-8
10 Step4 10 -8
Step8
Undisturbed

10 -10 10 -10
10 -2 10 -1 10 0 10 1 10 -2 10 -1 10 0 10 1
-2 -2
10 10

10 -4 10 -4

10 -6 10 -6

10 -8 10 -8

10 -10 10 -10
10 -2 10 -1 10 0 10 1 10 -2 10 -1 10 0 10 1

Figure 4.16: Upstream wall pressure spectra for four step cases and undisturbed boundary
layer at U∞ = 20 m/s. (a) x = −15h. (b) x = −4h. (c)x = −2h. (d) x = −0.5h. The
black solid line shows the slope of -5/2.

Top Surface

Figure 4.17 shows the pressure spectra measured on the top surface of the FBS. The
first measurement point downstream of the step leading edge is located inside the
separation bubble at x = 0.5h (Fig. 4.17(a)). The spectra are elevated throughout
the entire frequency range of interest due to the separated shear layer formed at the
sharp upper corner. The separated shear layer induces turbulence that generates
large wall pressure fluctuations. The levels at low frequencies are raised by 50 to
60 dB (∆dB = 10 log10 (∆Φpp /pref ), where pref = 20 × 10−6 Pa) compared with the
undisturbed boundary layer. Spectral roll-off occurs at f δ/U∞ ≈ 0.5 for Step1 and
at f δ/U∞ ≈ 0.68 for the remaining steps with a slope of -3. George et al. [28] showed
that a -11/3 slope occured in the turbulence-mean-shear interaction pressure spectra
and a -7/3 region in turbulence-turbulence interaction pressure spectra. The -3 slope
noticed here suggests that the pressure fluctuations in this area are produced by both
turbulence-mean-shear interactions and turbulence-turbulence interactions. For the
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 73

smallest AR Step1, the spectral level is generally 10 dB lower than the rest of the
step cases in the mid-frequency region because the roll-off starts at lower frequency.
This is likely due to the lack of the reattachment on the step top surface, which
alters the trajectory of the separated shear layer, lifts the flow structures away and
triggers spectral roll-off at a lower frequency compared with the rest cases. Moving
downstream to x = 1.5h, x = 2.5h, and x = 3.5h (Fig. 4.17 (b), (c), & (d), the
pressure spectra remain similar for all step cases near the reattachment region. The
roll off of the spectra at these locations were observed with -7/3 ∼ -8/3 slopes, which
is close to the -7/3 slope associated with the turbulence-turbulence interactions.

10-2 10-2

10-4 10-4

10-6 10-6
Step1
Step2
Step4 10-8
10-8
Step8
Undisturbed

10-10 10-10
10-2 10-1 100 101 10-2 10-1 100 101

10-2 10-2

10-4 10-4

10-6 10-6

10-8 10-8

10-10 10-10
10-2 10-1 100 101 10-2 10-1 100 101

Figure 4.17: Wall pressure spectra for four step cases on the step top surface and undis-
turbed boundary layer at U∞ = 20 m/s. (a) x = 0.5h. (b) x = 1.5h. (c)x = 2.5h. (d)
x = 3.5h. Noted that some of the spectra are not plotted as the considered location is no
longer on the step top surface.

Figure 4.18 shows the pressure spectra for Step8 at different top surface locations.
A spectral dip in mid-frequency region can be seen at the first measuring point
downstream of the leading edge x = 0.5h. A similar spectral dip was also noticed
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 74

by Farabee & Casarella [26] at the leading edge of a FFS. Note that in Fig. 4.17(a),
the spectral dip is not observed for the Step1 case. Although the reason of this mid-
frequency dip remains unknown, it is suspected to be associated with the separation
bubble on the top surface as the spectral dips is only seen in the cases where the this
separation bubble is complete (reattachment point is observed on the top surface).
Figure 4.18 also shows that the spectral level reduces with downstream distance.
The dashed line in Fig. 4.18 labels the spectral roll-off point for each measurement
position except for x = 0.5. It can be seen that the low frequency plateau regions
become shorter when moving away from the leading edge. This suggests that the
large scale structures created by the leading edge decay along with downstream
distance. For all testing points on the top surface of Step8, the spectral levels over
the entire frequency range are higher than that of undisturbed boundary layer.

x = 0.5h
x = 1.5h
-2
10 x = 2.5h
x = 3.5h
x = 6h
x = 7h
10-4 Undisturbed

10-6

10-8

10-10
10-1 100 101

Figure 4.18: Wall pressure spectra on the top surface of Step8 at different streamwise
locations.

Downstream

Now consider the wall pressure fluctuations downstream of the steps. Figure 4.19
shows the pressure spectra for all step cases compared with the undisturbed bound-
ary layer at 4 different downstream locations. Since the length of the steps are dif-
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 75

ferent, the downstream locations discussed in this section are the distance from the
step trailing edge and normalised on the second reattachment length xr2 estimated
in Section 4.4.1. As Mabey [53] and Ji & Wang [40] suggested, the reattachment
length is a good scale for BFS flow.
Inside the recirculating region at the front part of the separation bubble x/xr2 '
0.1 (Fig. 4.19 (a)), the pressure spectra present a similar pattern with that in the
upstream separation area where the spectral level increases at low frequency but
decreases at high frequency (Fig. 4.16(d)). The downstream separation bubbles are
in general larger than the upstream separation bubble and produce more uplift of
flow structures. Therefore, a lower level at high frequency region can be seen in Fig.
4.19a. Also, the shape of the spectra inside the separation bubble (Fig. 4.19(a)&(b))
for each step are not consistent. This is potentially due to the effect of step aspect
ratio. As discussed earlier for small aspect ratio steps, the upstream separated shear
layer may not reattach on the top surface for the steps that are not long enough.
Despite being subjected to different upstream flow conditions (different AR), the
pressure spectra at x/xr2 ' 0.1 for all step cases generally display a -11/3 decay
slope, which indicates the turbulence-mean-shear interactions at the upstream part
of this recirculating region.
At the middle part of the recirculating region x/xr2 ' 0.5, it can be seen that the
pressure spectra are roughly divided into two groups. The spectral levels for Step4
and Step8 are mostly lower than Step1 and Step2. For the shorter steps, large struc-
tures with formed at the leading edge may shed downstream without reattaching on
the top surface and therefore elevates the spectral level more. It should be noted
that although the length of Step2 is close to xr1 , due to the oscillating nature of
reattachment event, the reattachment point does not always stay on its top surface.
Moving to the reattachment region x ≈ xr2 in Fig. 4.19c, the shape of the spec-
tra for all steps are more consistent. It can be observed that the shorter steps tend
to have higher spectral level at low frequency region; however, the difference soon
becomes inconspicuous after the roll-off begins at f δ/U∞ ≈ 0.15. The flow reattach-
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 76

ment mechanism for different steps is similar and dominates the reattachment area,
thus results in similar wall pressure spectra.
The spectra for all steps remain similar further downstream in the redeveloped
boundary layer. Figure 4.19d shows the pressure spectra at x ' 3xr2 . Although the
wall pressure spectra are similar at, the spectral level of Step1 and Step2 are 5.5
dB higher than that of Step4 and Step8 in the low frequency region (f δ/U∞ < 0.2),
which indicates the large scale structures formed in the mixing shear layer are sus-
tained in the redeveloped boundary layer. Also, compared to the undisturbed bound-
ary layer, the spectra are tilted and a -7/3 spectral roll-off slope can be observed.
This implies the new boundary layer has not yet returned to the undisturbed bound-
ary layer at x ' 3xr2 . Farabee & Casarella [26] reported that even at 12xr after
the BFS, the redeveloped boundary layer still did not completely return to the
equlibrium state.

10 -2
10 -2

10 -4
10 -4

10 -6
10 -6
Step1
10 -8 Step2
Step4
10 -8
Step8
Undisturbed
10 -10
10 -10
10 -1 10 0 10 1 10 -2 10 -1 10 0 10 1

10 -2 10 -2

10 -4 10 -4

10 -6 10 -6

10 -8 10 -8

10 -10 10 -10
10 -2 10 -1 10 0 10 1 10 -2 10 -1 10 0 10 1

Figure 4.19: Downstream wall pressure spectra for four step cases and undisturbed bound-
ary layer at U∞ = 20 m/s. (a) x ' 0.1xr2 . (b) x ' 0.5xr2 . (c)x ' xr2 . (d) x ' 3xr2 . The
distance is calculated from the step trailing edge location.

Figure 4.20 shows spectral maps, which display the frequency spectral levels
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 77

measured in the streamwise direction, for all step cases. The contours are presented
in dB ref. 2 × 10−5 Pa. The dashed lines indicate the estimated reattachment points
(mentioned in Section 4.4.1). It can be seen that upstream of the step, the upstream
separation bubble energizes the low frequency contents and the up-lifted motions of
small structures caused by the separated layer lower the level at high frequencies.
Now compare the two separation bubbles downstream of the leading edge, the
one on the step top surface (SBA) and the one behind the step (SBB). For Step4
and Step8, the pressure spectrum is rich in mid-low frequency (<0.4) region inside
SBA. This high energy area is not clearly seen outside SBA. For SBB, this high
level fluctuating pressure domain is found nearby the reattachment point rather
than inside the separation bubble. This indicates that the extensive wall pressure
fluctuations on top of the step are mostly associated with the separated shear layer
encompassing the separation bubble, whereas downstream of the step the large fluc-
tuating pressure is generated by the flow reattachment events. Also for Step1 and
Step2, the downstream high pressure levels extend further than for Step4 and Step8.
As mentioned in Section 4.4.1, vortex shedding from the leading edge may alter the
trailing edge separated shear layer. The energized low frequency components seen
here are potentially the large scale structures shedding from the leading edge and
mixing with the downstream separated layer.
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 78

Figure 4.20: Streamwise wall pressure fluctuations for four step cases. The vertical black
solid lines are the step edges and the black dashed lines are the estimated reattachment
point. The contour is plotted in dB (ref. 20µPa) (a) Step1. (b)Step2. (c)Step4. (d) Step8.
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 79

Pressure Spectra Compared with Previous Studies

The pressure spectra for the current FBS are compared with some past FFS and
BFS studies where only one sharp edge is involved. Figure 4.21a considers the
pressure spectra at x = 2 ∼ 3h downstream of the step forward-facing edge. The
current data display a reasonable agreement with submerged-boundary layer steps
(0.41 < h/δ < 0.6) across a wide range of Reh [5, 26, 40] but deviate from the larger
step (h/δ > 1). This indicates that compared to Reh , the step height to boundary
layer thickness ratio (h/δ) plays a more critical role in FFS flow field modification.
Figure 4.21b shows the comparison with BFS [26, 40] downstream of the step at the
reattachment point. It can be seen that the pressure spectra for the larger aspect
ratio FBS present a more similar pattern with the BFS. This suggests that the FBS
aspect ratio indeed affects the downstream flow field and for a larger AR, where
the flow has longer relaxation space on the top surface, the downstream flow field is
close to the BFS scenario.

100

10-2

10-4

10-6

10-8

10-10
10-2 10-1 100 101

(a) x = 2 ∼ 3h.
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 80

100

10-2

10-4

10-6

10-8

10-10
10-2 10-1 100 101

(b) x = xr2 from step trailing edge.

Figure 4.21: Wall pressure spectra compared with past studies at two streamwise locations

4.4.3 Two-point Correlation and Coherence Analysis

Correlation

In this section, the two-point statistics of the wall pressure will be used to identify
flow structures that are associated with the pressure fluctuations. The two-point
correlation of steps flow in streamwise direction is computed using streamwise pres-
sure taps (z = 0). The step leading and trailing edges affect the flow structures
the most; therefore, the following discussion will mainly focus on the streamwise
space-time correlation near the step edges.
Figure 4.22 shows the step space-time correlation contour for Reh = 14078 near
the step leading edge (upstream of the leading edge and on the top surface). Each
column illustrates the space-time correlation for one step at 4 different streamwise
locations. Shown from the left to the right columns are Step1 to Step8. And
from bottom to top are the reference locations x/δ = 0 moving from upstream to
downstream (at xref = −0.5h, xref = 0.5h, xref = 1.5h, and xref = 2.5h). The black
solid lines and the blue solid lines represent the step leading edge and trailing edge,
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 81

respectively, and the black dashed lines seen in Step4 and Step8 are the estimated
reattachment point on the top surface. Note that Fig. 4.22 (a), (b) and (e) are
blank because the reference points at x = 1.5h and x = 2.5h no longer stay on the
top surface for Step1 and Step2.
At a short distance upstream of the step leading edge (x = −0.5h), wide, shallow
correlations are seen in the bottom row of Fig. 4.22. This correlation is caused by
the large, slow moving upstream separation bubble. The correlation is blocked by
the black solid line, which indicates that the step leading edge modifies the flow
field so that the pressure fluctuations on both sides of the leading edge are nearly
decorrelated.
Downstream of the leading edge, negative lobes can be observed for Step4 and
Step8 (indistinctly in Step2). These negative lobes imply that the new flow struc-
tures are formed. The staggered nature of the correlation lobes might be associated
with the structures in the separated shear layer. The length scale of these lobes is
around 1h (0.5δ), which is about the same length of Step1. Therefore these shed-
ding structures could potentially bypass the step top surface and directly travel
downstream without impinging on the top surface. This might explain the lower
r.m.s. wall pressure measured on the top surface for Step1 (see Fig. 4.4.1). In the
reattachment region, since the wall pressure fluctuations are still dominated by this
separated shear layer structure, the reattachment events are affecting the correla-
tion. This is also noticed by Ji & Wang [40]. Also, the slope of the main lobe is seen
to slightly increase with downstream distance, implying that the convection velocity
is increasing. This will be discussed in Section 4.4.3.
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 82

Figure 4.22: Space-time correlation coefficient Cpp for four step cases on the top surface.
From left: first column, Step1; second column, Step2; third column, Step4; fourth column,
Step8. From top: first row, x = 2.5h; second row, x = 1.5h; third row, x = 0.5h; fourth
row, x = −0.5h. The black and blue solid lines represent the step leading edge and trailing
edge, respectively. The black dashed lines indicate the estimated reattachment points.

Figure 4.23 provides the step space-time correlation contour downstream of the
step. Again, shown from left to right are the step with smallest to largest aspect
ratio (Step1 to Step8) and from bottom to top are streamwise locations moving
downstream (reference points at x = −0.5h, x = 0.8h, x ' 0.5xr2 , and x ' xr2
computed from step trailing edge). Here the black dotted lines indicate the estimated
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 83

downstream reattachment locations. For the first and second row from the bottom
of Fig. 4.23, the reference points are selected just upstream and downstream of the
step trailing edge (blue solid line) at x = −0.5h and x = 0.8h, respectively. It can
be seen that the correlation contour is continuous across the step trailing edge for all
step cases, indicating that the trailing edge does not alter the flow field as severely
as the leading edge does. This disagrees with Ji & Wang [40] who considered the
pure BFS where only one sharp corner is involved in the flow field. They found that
the BFS edge indeed modified the flow albeit not as much as the FFS edge did.
The difference is perhaps due to the leading edge corner disturbance which occurres
in the FBS scenario. In the present study, the FBS leading edge creates the large
separated shear layer structures which are carried across the trailing edge. It can
also be observed that the main lobes in the downstream recirculating area are much
wider. This indicates the large, slow-moving structure of the downstream separation
bubble. Similar to Fig. 4.22, the negative side lobes seen here implies the shedding
structures in the separated shear layer are periodically striking the wall. The length
scale of these structures are around 3 times larger than that observed on the top
surface.
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 84

Figure 4.23: Space-time correlation coefficient Cpp for four step cases downstream of the
step. From left: first column, Step1; second column, Step2; third column, Step4; fourth
column, Step8. From top: first row, x ' xr2 ; second row, x ' 0.5xr2 ; third row, x = 0.8h;
fourth row, x = −0.5h. The black and blue solid lines represent the step leading edge and
trailing edge, respectively. The black dotted lines indicate the estimated reattachment
points.

Convection Velocity

The cross-correlation of wall pressure fluctuations between two points can be used to
understand the speed at which the flow structures are convected. The time-averaged
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 85

convection velocity Uc is defined as

Uc = ∆x/∆τmax (4.3)

where ∆x is the spatial separation between two testing points and ∆τmax is the time
delay of the peak in the cross-correlation (Fig. 4.24). The normalized convection
velocity Uc /U for the undisturbed flow found in this measurement for U∞ = 20, 30,
and 40 m/s is 0.7, 0.68, and 0.69, respectively. The values are comparable to 0.6-0.8
that are generally recorded in undisturbed flow measurements [37, 43].

0.8

0.6

0.4

0.2

-0.2

-0.4
-0.01 -0.008 -0.006 -0.004 -0.002 0 0.002 0.004 0.006 0.008 0.01

Figure 4.24: Cross correlation of the wall pressure fluctuations of Step8 at U∞ = 20 m/s.
Reference point x = 0.5h.

Figure 4.25 shows the convection velocity under 3 different flow speeds at different
streamwise locations on the top surface of Step4 and Step8. The current data show
a fairly good agreement with Camussi et al. [13] who studied a FFS. The convection
velocity Uc /U∞ varies from around 0.3 to 0.5 with a nearly linear increase along the
streamwise direction for Step4. A similar trend can be observed for the Step8 case.
The low convection velocity close to step leading edge may be associated with the
separation bubble on the top surface, where the coherent structures are slower. The
convection velocity is accelerated to between 0.6 and 0.7 of free stream velocity in
the re-developed boundary layer on Step8. It can be noticed that the flow speed
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 86

does not significantly influence the results.


1 1

0.9 0.9

0.8 0.8

0.7 0.7

0.6 0.6

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7

(a) (b)

Figure 4.25: Time averaged convection velocity on the top surface of the step. (a) Step4.
(8) Step8.

The convection velocity at different streamwise locations downstream of the step


are plotted in Fig. 4.26. It can be seen that the results reasonably agree with
previous BFS studies [35, 37] regardless of the step aspect ratio. In the BFS sce-
nario, the convection velocity firstly decelerates and then re-accelerates from 3-4
h downstream of the backward-facing edge. Hudy et al. [37] suggest that this re-
acceleration is associated with the separated shear layer shedding mechanism. It
should be noted that the plots do not scale well with the downstream reattachment
length xr2 , indicating that the convection velocity downstream might be more re-
lated to the distance to the trailing edge rather than the downstream reattachment
length xr2 . The lowest Uc recorded here is around 0.3 of the free stream speed and
the value returns to undisturbed flow level (0.7U∞ ) in the redeveloped boundary
layer. Again, the Reynolds number based on step height seems to have marginal
influence on the convection velocity variation.
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 87

1 1

0.9 0.9

0.8 0.8

0.7 0.7

0.6 0.6

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30

(a) (b)
1 1

0.9 0.9

0.8 0.8

0.7 0.7

0.6 0.6

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30

(c) (d)

Figure 4.26: Time average convection velocity downstream of the step. (a) Step1. (b)
Step2. (c) Step4. (d) Step8.

Spanwise Coherence

As introduced in Section 3.5.1, here the cross spectrum of the fluctuating wall pres-
sure at two different points in spanwise direction was used to quantify the spanwise
coherence. The pressure taps at z = 0 were selected as the reference point and the
spatial separation between two points ∆z are not equal (see Section 3.2.4). The
contour map of the coherence (squared) γ 2 are presented to illustrate the relations
of space and frequency. Figure 4.27 (a)-(c) show the spanwise coherence (γ 2 ) for
undisturbed boundary layer at 20,30 and 40 m/s, respectively. In this figure the
horizontal axis represents the normalized frequency in log scale and the vertical
axis represents the distance from the reference point to different spanwise testing
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 88

locations. As the pressure taps were only placed on one side of the reference point
(z = [0 2h] see Fig. 3.9), the contour maps were presented only single-sided in
space. The contour levels are shown in 0.01 increments and the corresponding color
is specified at the top of the figure.

Figure 4.27: Spanwise coherence (squared) γ 2 for undisturbed boundary layer at three
different flow speeds. (a) U∞ = 20 m/s. (b) U∞ = 30 m/s. (c) U∞ = 40 m/s.

For the undisturbed boundary layer in Fig. 4.27, the coherence maps show
similar features at different flow speeds. Low levels of coherence are seen in the
spanwise direction at low frequency region, and the coherence drops away after
around f δ/U > 1. Palumbo [60] suggested that the coherence length at a particular
frequency can be obtained by fitting an exponential curve to the measured γ at that
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 89

frequency. The spanwise coherence length Lfi at the frequency fj is then calculated
as
|z0 −zi |

γ(z0 , z1 , fj ) ∼
=e Lf
i (4.4)

where z0 and z1 are two separated points in spanwise direction and γ(z0 , z1 , fj ) is the
measured coherence. Figure 4.28 illustrates the measured data and the exponential
fit curves at 4 different frequencies for Step8 at x = 2h downstream from the trailing
edge (U∞ = 20m/s).

1
232 Hz
0.9 392 Hz
488 Hz
0.8 872 Hz

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2

Figure 4.28: Spanwise coherence γ 2 for Step8 at x = 2h and the curves.

The curve fitting can be applied to each frequency and the estimated spanwise
coherence lengths can then be presented as a function of frequency. Figure 4.29
shows the coherence length for three different flow speeds. At f δ/U less than 0.2,
the spanwise coherence length scales are generally 0.3 ∼ 0.5 of the boundary layer
thickness with the value slightly larger for higher incoming flow speed. This indicates
the wall pressure fluctuations for undisturbed flow are influenced by structures that
are limited by 0.5 boundary layer thickness in spanwise direction. Note that Fig.
4.29 excludes content of f δ/U > 2.4 since the coherence lengths above this frequency
are too small and become difficult to resolve for the smallest spatial spacing ( ∆z =
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 90

0.2h).

2
u = 20m/s
u = 30 m/s
1.5 u = 40 m/s

0.5

0
10-1 100

Figure 4.29: Spanwise coherence length as the function of frequency for undisturbed
boundary layer at three flow speeds.

Figure 4.30 shows the spanwise coherence contour at x = −0.5h for 4 step cases
at U∞ = 20 m/s. The contour maps for all of the step cases are nearly identical,
where the higher coherence can be observed at lower frequency region within half
step height (∆z < 0.5h) compared to the undisturbed boundary layer (Fig. 4.27).
Note that vertical lines at f δ/U =10 and small low level coherence bump at f δ/U
around 8 are regarded as electrical noise since at that frequency the microphone
response to the wall pressure is greatly attenuated. The vertical line noise is also
seen in Fig. 4.16(d) where the measurements were conducted at the same streamwise
location.
The spanwise coherence length scale at different frequencies for 4 steps at x =
−0.5h are given in Fig. 4.31. The coherence length scale are weakly dependent
on step aspect ratios as this is at the upstream location where the flow conditions
are similar. It can be noticed that a roll-off occurs at f δ/U ' 0.25 and in the low
frequency region (f δ/U < 0.25). The spanwise coherence length for all steps are
somewhat larger than the undisturbed boundary layer and then drops near the zero
value at f δ/U ' 1. At f /δ > 0.7, the coherence lengths become smaller than the
undisturbed boundary layer. This implies that the flow structures in spanwise direc-
tion become wider at low frequency region and narrower at higher frequency region.
Note that at x = −0.5h the pressure taps are immersed in the upstream separation
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 91

bubble where the low frequency structures are enhanced and high frequency struc-
tures are lifted. Similar feature are also seen in the wall pressure auto-spectrum in
Fig. 4.16.

Figure 4.30: Spanwise coherence (squared) γ 2 upstream of the step at x = −0.5h. (a)
Step1. (b) Step2. (c) Step4. (d) Step8.

2
Step1
Step2
1.5 Step4
Step8
Undisturbed

0.5

0
10-1 100

Figure 4.31: Spanwise coherence length as the function of frequency for U∞ = 20 m/s at
x = −0.5h.

Now consider the separation bubble downstream of the step. Figure 4.32 shows
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 92

the spanwise coherence at x = 2h downstream of trailing edge. When compared with


the upstream separation bubble in Fig. 4.31, the spanwise coherence of downstream
separation bubble is larger. The high level of coherence across the spanwise direction
indicates the large, slow-moving structure of the separation bubble. The coherence
lengths of each step at different frequencies are shown in Fig. 4.33. At the lower
frequency region when f δ/U is below 0.2, the spanwise coherence length scales are
affected by the step aspect ratios: the coherence length of Step1 and Step2 are larger
than that of Step 4 and Step8. In this frequency range, for Step1 and Step2, the
spanwise coherence length scales are generally between 3∼5 times of the step height.
While for Step4 and Step8 the spanwise coherence length are mostly between 1.5 ∼
3 step height and the length scales increase with frequency. Above f δ/U =0.2, the
coherence length to frequency plots of different steps collapse to a single data set
and gradually reduce to near zero.
The large scale coherent structures seen in Step1 and Step2 at lower frequency
region were associated with the large structures induced by leading edge shed-
ding event. Large spanwise coherence is retained since the structures are mixing
with the downstream separation bubble without reattaching. The critical frequency
f δ/U =0.2 seen here is very close to 0.2 ∼ 0.3 which is showed in Fig. 4.19 to be
the frequency where the spectral level starts to decay. This suggests that regardless
of the aspect ratio, the turbulence and shear flow interaction mechanism dominates
the flow structures when f δ/U =0.2. The coherence length scales are approximately
2.8 for all step cases.
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 93

Figure 4.32: Spanwise coherence (squared) γ 2 downstream of the step at x = 2h from


trailing edge. (a) Step1. (b) Step2. (c) Step4. (d) Step8.

5
Step1
Step2
4
Step4
Step8
Undisturbed
3

0
10-1 100

Figure 4.33: Spanwise coherence length as the function of frequency for U∞ = 20 m/s at
x = 2h from trailing edge.

On the top surface of the step, due to the physical limitation of the experiment,
the spanwise measurements were only made at half of the step length locations.
Therefore, the streamwise locations of these pressure taps are x = 0.5h (Step1),
x = h (Step2), x = 2h (Step4) and x = 4h (Step8). For Step1 and Step2, where the
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 94

testing locations are regarded inside the separation bubble formed by the leading
edge, the spanwise coherence contour (Fig. 4.34 (a) & (b)) shows a wider coherence
length scale at f δ/U =1 compared with those measured at upstream and downstream
locations (Fig. 4.30 & Fig. 4.32). The spanwise coherence length scale inside the
top surface separation bubble is larger than the undisturbed boundary layer (Fig.
4.35). At x = 2h (Step4) where the testing points are considered immersed in the
reattachment region, the spectral roll-off has shifted to a higher frequency at f δ/U '
0.5 ∼ 0.6 (was found to occur at around 0.25 in the undisturbed boundary layer
and 0.2 in downstream separation bubble). Outside of the recirculating area at x =
4h(Step8) in the redeveloped boundary layer, the spanwise coherence length scale
returns to the undisturbed boundary layer level except for in the lower frequency
region which might be due to the effect of large structures remaining in the flow.

Figure 4.34: Spanwise coherence (squared) γ 2 on the top surface of the step at half step
length x = 0.5l. (a) Step1. (b) Step2. (c) Step4. (d) Step8.
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 95

2
Step1
Step2
1.5 Step4
Step8
Undisturbed

0.5

0
10-1 100

Figure 4.35: Spanwise coherence length as the function of frequency for U∞ = 20 m/s at
x = 0.5l from the leading edge (on the top).
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 96

4.5 Summary

Features of the reattachment length and the wall pressure fluctuations for different
aspect ratio steps have been investigated in this chapter. In the oil flow visualiza-
tion experiment, the reattachment length after the step trailing edge is found to
be affected by the step aspect ratio. The reattachment points are found further
downstream in the small aspect ratio step. For the longer steps, the values return
to that generally found in BFS scenarios.
In the wall pressure measurement, the peak negative value of Cp measured on
the top surface of the step is found to be lower for the smaller aspect ratio step.
This is related to the lack of reattachment on the top surface. And due to the upper
corner influences, the Cp distribution downstream of the FBS does not agree well
with BFS studies.
The wall pressure auto-spectra at different streamwise locations are compared
with measurements taken in an undisturbed boundary layer. Inside the upstream
and downstream recirculating areas, the spectral level for steps are elevated at low
frequency (f δ/U < 1) due to the large, slow-moving structures associated with sep-
aration bubble. The spectral levels decrease at high frequency because the small
scale structures are lifted away from the wall. Different spectral roll-off slope indi-
cate the different turbulence mechanism dominating the pressure fluctuations at that
region. Near the sharp edges (step leading edge and trailing edge) where the sep-
arated shear layers are formed, the fluctuating wall pressure is associated with the
turbulence-mean-shear interactions whereas downstream of the reattachment, the
pressure fluctuations are believed to be overshadowed by the turbulence-turbulence
interactions. Also, the streamwise frequency spectrum reveals that the wall pres-
sure fluctuations generated by the step leading edge are generally larger than that
created by the trailing edge across the whole frequency region.
The two-point statistics of the wall pressure shows the influence of the upper cor-
ner disturbance to the downstream flow field. The space-time correlation contours
Chapter 4. Oil Flow Visualization and Wall Pressure Measurements 97

show that the step trailing edge effect of flow modification becomes less important
because high momentum structures created by the leading edge are convected down-
stream. This again, is consistent with the previous studies that the leading edge
of the FBS plays the more important role in the flow field. The spanwise coher-
ence length for different aspect ratio steps shows that for the smaller AR step, the
spanwise coherence lengths are no longer at low frequency as the large structures
produced by the step leading edge blends into the downstream flow without losing
spanwise coherence by the reattachment events on the top surface.
Chapter 5

CFD Results

5.1 Overview

In this chapter, the numerical simulation results obtained using LES are presented.
The CFD results are compared with experimental data presented in Section 5.2 and
show a reasonable agreement with both steady and unsteady wall pressure.
The flow field around different steps is presented in Section 5.3. The mean veloc-
ity field identifies the separation and reattachment regions as well as the separated
shear layer structures near the step. The location of high level turbulent kinetic
energy (TKE) and Reynolds stress components are found consistent with the sepa-
rated shear layer structure. The highest levels are seen near the step leading edge,
indicating that the flow fields are mainly modified by the upper step corner.
In Section 5.6, spectral proper orthogonal decomposition (SPOD) is performed on
the wall-normal fluctuating velocity v 0 . The flow field presents low-rank behaviour in
low frequency region (Stδ < 1), implying dominance of the first SPOD modes. The
SPOD modes successfully capture the coherent structures in the separated shear
layer. Finally, the wall pressure fluctuation and wall-normal fluctuating velocity
coherence maps shows the contribution of vortex shedding events induced by the
step leading edge on surface pressure.

98
Chapter 5. CFD Results 99

5.2 LES Comparison with Mean and Fluctuating

Wall Pressure Measurements

The CFD simulations were carried out at Reh = 12, 318 for AR = 1, 2.4, 4, and 8
steps. Note that Step1, Step2, Step4, and Step8 were only used to refer to testing
cases with AR = 1, 2, 4, and 8, respectively, in the wall pressure and acoustic
measurements. Figures 5.1 & 5.2 compare the wall pressure coefficient and r.m.s.
fluctuating wall pressure predicted by LES with the wall pressure data collected
in section 4 at a similar Reynolds number of Reh = 14, 078. For the mean wall
pressure (Fig.5.1), the overall agreement between the LES and experiments is good
for AR = 2, 4, and 8 steps. The LES model over-predicts the Cp value inside of the
downstream recirculation bubble for AR = 1 step. This is probably because the size
of the computational grid inside the separation region is not ideal. Nonetheless, the
overall agreement between the CFD prediction and experiment is reasonable.
For the unsteady wall pressure (see Fig.5.2), the LES results overall agree very
well with experimental data. At the location of the upstream separation bubble,
there is some disagreement for all step cases. This is because the inlet velocity
profile for the LES was pre-calculated using a steady RANS model; therefore it has
no fluctuating velocity components with turbulence developing only in the short
fetch upstream of the step. This results in the LES being unable to capture the
undisturbed turbulent flow statistics accurately. However, the most dominant flow
modification in the present configuration occurs at step leading edge; and the in-
stantaneous velocity components are mostly determined by this edge and are barely
affected by the upstream conditions. It can be seen that despite an offset of the
peak value, the LES predictions downstream of the step leading edge are a reason-
able match with the experimental results for Step2, Step4, and Step8. For AR = 1,
although the LES tends to over predict the r.m.s. value, the overall trend is simi-
lar. It should be noted that the purpose of the CFD simulation is to visualize the
flow alterations caused by the FBS. The flow field investigation mainly focuses on
Chapter 5. CFD Results 100

the region near and downstream of the leading edge, where the LES prediction are
reasonable.
1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6
-0.8 -0.8
-1 -1
-15 -10 -5 0 5 10 15 20 25 30 35 -15 -10 -5 0 5 10 15 20 25 30 35

1 1
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
-0.2 -0.2
-0.4 -0.4
-0.6 -0.6
-0.8 -0.8
-1 -1
-15 -10 -5 0 5 10 15 20 25 30 35 -15 -10 -5 0 5 10 15 20 25 30 35

Figure 5.1: Comparison of LES and measured mean pressure coefficient Cp distribution.
Red solid lines, LES; Blue dots, wall pressure measurements. Red and black dashed lines
represent the step leading edge and trailing edge, respectively. Black dotted line in (b)
indicates the step trailing edge in CFD case. (a) LES AR = 1 compared with Step1. (b)
LES AR = 2.4 compared with Step2. (c) LES AR = 4 compared with Step4. (d) LES
AR = 8 compared with Step8.
Chapter 5. CFD Results 101

0.2 0.2

0.15 0.15

0.1 0.1

0.05 0.05

0 0
-15 -10 -5 0 5 10 15 20 25 30 35 -15 -10 -5 0 5 10 15 20 25 30 35

0.2 0.2

0.15 0.15

0.1 0.1

0.05 0.05

0 0
-15 -10 -5 0 5 10 15 20 25 30 35 -15 -10 -5 0 5 10 15 20 25 30 35

Figure 5.2: Comparison of LES and measured r.m.s. wall pressure fluctuations distribu-
tion. Red solid lines, LES; Blue dots, wall pressure measurements. Red and black dashed
lines represent the step leading edge and trailing edge, respectively. Black dotted line in
(b) indicates the step trailing edge in CFD case. (a) LES AR = 1 compared with Step1.
(b) LES AR = 2.4 compared with Step2. (c) LES AR = 4 compared with Step4. (d) LES
AR = 8 compared with Step8.

5.3 Mean Flow Field

The mean velocity field near the FBS is given in Fig.5.3. In the figure, the mean
streamwise velocity U / U∞ is shown with mean streamlines overlayed. For AR
= 2.4, 4, and 8 steps, three separation bubbles can be observed in the flow field
and only two are seen for the AR = 1 step. In general, the flow deflects upward
ahead of the step, separates at the leading edge and then reattaches downstream
of the leading edge. After reattaching on the step, the boundary layer detaches
again at the trailing edge to form a large separation bubble. The downstream mean
reattachment lengths xr1 and xr2 were determined by the zero-crossing location of
the wall shear stress. Table 5.1 compares the mean reattachment lengths predicted
by LES for different aspect ratio steps with experiment. Note that the zero-crossing
Chapter 5. CFD Results 102

location of the wall shear stress is ambiguous in the upstream area, therefore the
separation points are estimated from the mean streamlines.

Figure 5.3: Mean streamwise velocity contours (U/U∞ ). Black solid lines represent the
mean streamlines. Red triangle and cross mark the locations of separation points and
reattachment points, respectively. The color bar represent the level of U/U∞ . (a) AR = 1.
(b) AR = 2.4. (c) AR = 4. (d) AR = 8.

The estimated upstream separation lengths from the LES are all approximately
-0.7∼-0.8 h regardless of the step aspect ratio. This value is around 50% below
Chapter 5. CFD Results 103

that observed in the oil flow visualisation (Section 4.2) but is comparable to −0.8h
to −1h observed in previous studies [13, 24, 47, 49]. As suggested by Pearson et
al. [62], the upstream separation bubble characteristics are modulated by upstream
incoming flow structures; therefore, the separation length seems to be independent
of the step geometry and h/δ.
The separated flow on top of the step reattaches at a position of 2h from the
leading edge for steps with AR = 2.4, 4, and 8. The reattachment point is not
found on the AR = 1 step. The first reattachment length xr1 is comparable to
2h estimated in wall pressure measurements for step4 and step8 and 1.65h to 1.8h
measured by the oil flow visualization experiment (Chapter. 4). This value is also
similar to 1.6h reported by Fang & Tachie[23] for a AR = 2.36 FBS. It can be seen
that the step aspect ratio has a small influence on the first reattachment length
when the step is longer than xr1 .
The second reattachment lengths that occur downstream of the step, xr2 , are
found to be 11.5h, 6.1h, 4.8h, and 4.5h from step trailing edge for AR = 1, 2.4, 4,
and 8 step, respectively. These values are very close to that estimated in section 4.4.1
(see Table 5.1). This demonstrates that using r.m.s. wall pressure is a reasonable
way to estimate the downstream reattachment length. The xr2 predicted by LES
also agrees with Bergeles & Athanassiadis[12] who suggested that xr2 is larger for
smaller AR FBS, and the value remains nearly constant when AR > 4. It should be
noted that although the flow conditions are similar for CFD simulations and the oil
flow visualization measurements, xr2 are slightly different. This is potentially due to
the unsteady nature of the turbulence where the reattachment lengths are estimated
by the center of the reattachment region in the oil flow visualization measurements.
From the mean flow field contours (Fig.5.3), it can be clearly seen that the
downstream separation bubble is sensitive to step aspect ratio. For a longer FBS
xr1
(AR > h
), as the step length is long enough to accommodate the first downstream
separation bubble, the backward-facing step nature is preserved as the reattachment
lengths are nearly constant and close to BFS value. Whereas for AR = 1 step, the
Chapter 5. CFD Results 104

separated layer flows over the step without reattaching on top surface, resulting in
the large recirculation region.
Figure 5.4 shows the contour of mean ∂u/∂y, which is used to approximate the
shear layer structure. The figure shows that the main shear layer structures are
encompassing the separation bubbles and the highest values are all found near the
step leading edge. The mean ∂u/∂y contour for AR = 1 step shows only one large
shear layer structure and it is slightly tilted up compared to the rest cases. This
results in a higher angle trajectory and is one of the reasons for the longer down-
stream reattachment length. For AR =2.4, 4, and 8 step, where the reattachment
locations occur on the step top surface, the shear layer structures are seen emanat-
ing from both the leading edge and trailing edge. The two separated shear layers
merge into one structure further downstream. Fang & Tachie[23] suggested that the
step downstream separation bubble would be altered by the top surface separation
bubble via a mixing process.

Table 5.1: Separation and reattachment lengths predicted by LES and compared with
experiment results

Methods AR h/δ Reh xs /h xr1 /h xr2 /h


1 0.526 12,318 -0.8∼-0.7 - 11.5
2.4 0.526 12,318 -0.8∼-0.7 2 6.1
LES
4 0.526 12,318 -0.8∼-0.7 2 4.8
8 0.526 12,318 -0.8∼-0.7 2 4.5
1 0.5 14,078 - - 11.75
2 0.5 14,078 - - 6.5
Wall Pressure
4 0.5 14,078 - 2 5.35
8 0.5 14,078 - 2 5.1
3 0.526 12,318 -1.5 1.4 6.26
Oil Flow 4 0.526 12,318 -1.6 1.65 5.86
Visualization 6 0.526 12,318 -1.6 1.6 5.95
8 0.526 12,318 -1.6 1.8 5.9
Chapter 5. CFD Results 105

Figure 5.4: Mean ∂u/∂y contours. (a) AR = 1. (b) AR = 2.4. (c) AR = 4. (d) AR = 8.

5.4 Turbulence Statistics

Figure 5.5 shows the turbulent kinetic energy (TKE) distributions for the different
aspect ratio steps. The black solid lines are the contours of ∂u/∂y shown in Fig. 5.4.
It can be observed that the turbulent kinetic energy induced by the step overlaps
with the shear layer structure with the highest levels found just downstream of
the step leading edge. The high level TKE is associated with the formation of the
separated shear layer at the step leading edge, where the flow is highly turbulent.
Chapter 5. CFD Results 106

Downstream of the step, the mixing of two shear layers enhances the TKE level
(e.g.10 < x/h < 15 in Fig.5.5 (c)) and the TKE levels for smaller aspect ratio steps
are in general higher. The higer TKE levels in smaller AR steps are due to the large
structures created at leading edge carrying high level turbulent energy downstream.
Also, the TKE associated with the first separation bubble is in general two to three
times higher than that with the second separation bubble, which indicates that the
leading edge plays a critical role in turbulence generation.

2 . Black solid lines indicate ∂u/∂y


Figure 5.5: Turbulent kinetic energy (u02 + v 02 + w02 )/U∞
with ten contour levels. (a) AR = 1. (b) AR = 2.4. (c) AR = 4. (d) AR = 8.

The Reynolds stress components in the x − y plane ((u0 u0 ) , (v 0 v 0 ) and (u0 v 0 )) are
Chapter 5. CFD Results 107

shown in Fig. 5.6, 5.7 & 5.8 for all step cases. The dot-dashed lines represent mean
∂u/∂y. The overall contours for each Reynolds stress component are similar to the
turbulent kinetic energy spatial distribution, which is associated with the separated
shear layers. For every step case, the highest magnitude velocity fluctuations value
occur right downstream of the leading edge with the second highest region located
downstream of the trailing edge. The velocity fluctuations in the streamwise direc-
tion u0 u0 (Fig. 5.6) are approximately twice of that in the wall-normal direction
v 0 v 0 (Fig. 5.7). On the top of the step, for AR =2.4, 4, and 8 steps, the most
intense region for v 0 v 0 is located at the rear of the first separation bubble, which is
observed slightly downstream of the maximum region for u0 u0 . This is perhaps due
to the reattachment event that occurs at the rear part of the separation bubble that
involves flow fluctuations in the wall-normal direction. Similar phenomena can be
observed downstream of the step where the wall-normal fluctuating velocity v 0 v 0 is
higher near the reattachment point for all step cases.
The Reynolds stress u0 v 0 (Fig. 5.8) exhibits small positive values near the leading
edge regardless of the step aspect ratio. This observation has also been reported
by several previous studies [13, 23, 34]. Hattori and Nagano [34] believed this was
caused by the counter-gradient diffusion phenomenon (CDP), which occurs when the
Reynolds shear stress and the mean velocity gradient are opposite in sign. Whereas
Nematollahi and Tachie [57] explained that this is due to the interaction between
the high-momentum convection in the streamwise direction and the upwash along
the vertical face.
In general, the step without flow reattaching on the top surface (AR = 1) has
higher TKE than the other cases. While for AR =2.4, 4, and 8 steps, although
the flow reattaches on the step top surface, the downstream turbulence statistics
are affected by step aspect ratio. With smaller AR, the downstream flow fields
are more easily influenced by the separated shear layer emanating from the leading
edge. Also, it can be noticed that the area covered by the first separation bubble
shear layer contains the strongest velocity fluctuations and highest turbulent kinetic
Chapter 5. CFD Results 108

energy. This region has also been identified as the location of the highest wall
pressure fluctuations (see Fig.5.2).

Figure 5.6: Reynolds stresses component u0 u0 . Black dotted lines indicate ∂u/∂y with ten
contour levels. (a) AR = 1. (b) AR = 2.4. (c) AR = 4. (d) AR = 8.
Chapter 5. CFD Results 109

Figure 5.7: Reynolds stresses component v 0 v 0 . Black dotted lines indicate ∂u/∂y with ten
contour levels. (a) AR = 1. (b) AR = 2.4. (c) AR = 4. (d) AR = 8.
Chapter 5. CFD Results 110

Figure 5.8: Reynolds stresses component u0 v 0 . (a) AR = 1. (b) AR = 2.4. (c) AR = 4.


(d) AR = 8.
Chapter 5. CFD Results 111

5.5 Instantaneous Vorticity Field

Previous literature [46, 54, 71] has examined the shear layer vorticity formation
mechanism of turbulent flow over a bluff body. Just downstream of the leading
edge at the front of the separation bubble, the shear layer rolls up due to the
Kelvin–Helmholtz (KH) instability. These KH-roll up vortices are convected down-
stream and pair with other vortices, growing into larger scale structures. The large
scale vortex shedding away from the leading edge is associated with large scale
momentum transport.
Fig.5.9 and 5.10 show the instantaneous vorticity fields for steps with AR = 1
and 8, respectively. The figures are presented at a constant time increments 0.67t∗ ,
where t∗ = ∆th
U∞
. In both step cases the KH-roll up and pairing of the vortices can
be clearly seen at the leading edge. However, the trajectory of the vortex shedding
for two cases are different. For AR = 1 step, the large scale vortex is shed to the
downstream wake region, while for AR = 8 step, it is shed toward the wall.
Previous studies [21, 57, 59] point out the two main mechanisms that control
the unsteadiness of the flow over the FFS step, include KH vorticity roll up and
shear layer flapping motion. The latter is claimed as the factor that leads to os-
cillating reattachment point [21, 42, 59]. Moore et al. [54] suggests that the lack
of reattachment allows the leading edge vortex to be shed into the wake region di-
rectly, and the flapping motion of the separated shear layer combines with the wake
resulting in the formation of von Kármán (VK) vortices. The large VK structures
correspond with larger scale momentum transport, which leads to higher levels of
Reynolds stress and TKE in the downstream area for the smaller aspect ratio step
(Figures 5.5 to 5.8).
xr1
For AR = 8 step (and other AR > h
steps), due to the shear layer flapping
motion, the large scale vortex formed at the leading edge are able to impinge on the
wall, and therefore a higher level of r.m.s. wall pressure fluctuation was recorded on
the top surface (Fig. 5.2).
Chapter 5. CFD Results 112

KH-roll up and vortex pairing Vortex shedding to downstream directly

Figure 5.9: Instantaneous vorticity field for AR = 1 step. The time increment t∗ = ∆th
U∞ ,
where ∆t is the sampling time 4.8 × 10−6 s.

Separated shear layer flapping motion


KH-roll up and vortex pairing (reattachment point oscillating)

Figure 5.10: Instantaneous vorticity field for AR = 8 step. The time increment t∗ = ∆th
U∞ ,
where ∆t is the sampling time 4.8 × 10−6 s.
Chapter 5. CFD Results 113

5.6 Spectral Proper Orthogonal Decomposition

5.6.1 SPOD Modes

The wall pressure fluctuations essentially correspond to the unsteady loading acting
on the surface, and the wall-normal fluctuating velocity v 0 is involved in this motion.
In this section, Spectral Proper Orthogonal Decomposition (SPOD) will be applied
to the v 0 velocity field. As the step leading edge is regarded as the main source
of flow modification, the analysis mainly concentrates on the flow field around the
leading edge.
Figures 5.11 to 5.14 shows the SPOD results of the v 0 velocity field around the
step for all cases. On the left of these figures are the SPOD spectrum of each eigen-
mode. The corresponding eigenvalue is regarded as the energy that the eigenmode
contains and is normalized by the maximum value of the first mode. The first two
modes at three different frequencies are plotted on the right to identify the coherent
structures in this flow separation area.
It can be seen that for all step cases, at Stδ between 0.06 and 1, the separation
of the modal energy (spectral level) between the first and the rest of the eigenmodes
is large. This indicates that the first mode contains the largest portion of the total
energy and this is referred to as low-rank behaviour [67] where the leading mode
predominates the flow field of interest. Above Stδ ≈1, the low-rank behaviour is not
as pronounced. Here three frequencies are chosen to present the flow field of v 0 , the
two local peak frequencies in the spectra and Stδ ≈ 1. For Figures 5.11 to 5.14 the
middle column shows the first modes and in the third column from the left are the
second modes.
This discussion will begin by examining the first modes for all step cases. The
black solid lines are the contours of mean ∂u/∂y, representing the approximate
shear layer structure. It can be observed that the periodically appearing coherent
structures are consistent with the ∂u/∂y contours, identifying vorticity shedding
behaviours in the separated shear layer structures. This is believed to be driven
Chapter 5. CFD Results 114

by the KH-roll up and pairing mechanism, which is one of the main mechanism
controlling the unsteadiness of the separated shear layer. The spatial scale of the
coherent structures of the first mode are seen to be larger moving downstream from
the leading edge due to the growing and pairing of these vortices.
For the second modes, as the energy is in general lower, the coherent structure
shapes are not as distinct as for the first mode. Yet at the second peak frequency,
xr1
for AR > h
steps ( Figures 5.12 to 5.14 (d)), the coherent modes are seen more
clearly at 0 < x/h < 2. These concentrated structures can be referred to as a
Kelvin–Helmholtz (KH) type wavepacket[67], which originally evolve through the
Kelvin–Helmholtz mechanism, reach a peak, and then decay downstream. The
mean streamlines are plotted with the SPOD second modes contours, demonstrating
the on-step separation bubble. It can be noticed that this KH-type wavepacket
approximately overlapped with the separation bubble and the mode shape becomes
unclear downstream of the reattachment point (Figures 5.12 to 5.14 (d)). It is
speculated that this KH-type wavepacket is related to the flapping motion of the
separated shear layer on the step top surface as the wavepacket structures seem to
be enclosed by the reattachment event. The KH-type wavepacket is not seen for
AR = 1 since the step is too short for the reattachment.
To briefly summarise, the SPOD eigenspectra indicates that the low-rank be-
haviour of flow field around the step leading edge at frequency between 0.06 < St <
1. The first SPOD mode captures the structure of the separated shear layer created
by the step leading edge, and the KH-roll up and vortex pairing is considered the
dominant mechanism in this separated shear layer. While the second mode is not
as energetic as the first mode, the organized KH-type wavepackets are observed for
xr1
AR > h
steps. It is assumed to be associated with the shear layer flapping motion,
which is the other main mechanism responsible for the unsteadiness of the separated
shear layer.
Chapter 5. CFD Results 115

Figure 5.11: SPOD eigenspectra and first two SPOD modes at three different frequencies
for AR = 1. On the left : SPOD eigenspectra for first 12 modes. (b, d, f) First SPOD
modes. Black solid lines represent ∂u/∂y contours; (c, e, g) Second SPOD modes. Black
solid lines represent mean streamlines.

1 2 2

1 1

0.8 0 0
-1 0 1 2 3 -1 0 1 2 3

0.6 2 2

1 1

0.4 0 0
-1 0 1 2 3 -1 0 1 2 3

0.2 2 2

1 1

0 0 0
100 -1 0 1 2 3 -1 0 1 2 3

Figure 5.12: SPOD eigenspectra and first two SPOD modes at three different frequencies
for AR = 2.4. On the left : SPOD eigenspectra for first 12 modes. (b, d, f) First SPOD
modes. Black solid lines represent ∂u/∂y contours; (c, e, g) Second SPOD modes. Black
solid lines represent mean streamlines.
Chapter 5. CFD Results 116

Figure 5.13: SPOD eigenspectra and first two SPOD modes at three different frequencies
for AR = 4. On the left : SPOD eigenspectra for first 12 modes. (b, d, f) First SPOD
modes. Black solid lines represent ∂u/∂y contours; (c, e, g) Second SPOD modes. Black
solid lines represent mean streamlines.

Figure 5.14: SPOD eigenspectra and first two SPOD modes at three different frequencies
for AR = 8. On the left : SPOD eigenspectra for first 12 modes. (b, d, f) First SPOD
modes. Black solid lines represent ∂u/∂y contours; (c, e, g) Second SPOD modes. Black
solid lines represent mean streamlines.
Chapter 5. CFD Results 117

5.6.2 Pressure-Velocity Coherence

The highest levels of r.m.s. fluctuating wall pressure were measured on the top
surface of the step (see Fig.5.2). The same location also recorded extensive TKE
and Reynolds stress contributions (see Fig.5.5 to 5.8). In order to address the
relationship between the wall pressure fluctuations and the turbulent velocity field,
the coherence between the fluctuating velocity in the wall-normal direction v 0 and
the fluctuating wall pressure p0 has been calculated. Figures 5.15 to 5.18 shows the
coherence coefficient (squared) contour Cpv between p0 on the reference wall location
(marked by red triangles) and the v 0 flow field above the step. The coherence was
only computed for the flow field above the step as this region is considered more
important for the wall pressure on the step top surface. Here the two local peak
frequencies seen in the SPOD eigenspectra (Figures 5.11 to 5.14) were chosen for
the discussion.
At the first reference point inside the separation bubble at x/h ≈ 0.5, the high
coherence level region is seen to generally coincide with the separated shear layer
for all steps, indicating its significant contribution to the wall pressure fluctuations.
The highly coherent region in Fig.5.17 (b) shows a similar shape to the first SPOD
mode at Stδ = 0.233, meaning that the surface pressure is mainly caused by this
coherent structure at this frequency. Similar features are observed at the second
frequency peak for all step cases (subplot e in figs. 5.11 to 5.14), where the shape
of the highly coherent region resembles the first SPOD mode at the corresponding
frequencies, demonstrating the influences of the shear layer on the wall pressure.
The effect of the shear layer has also been seen in time domain correlation analysis
[13, 40, 47] where the shedding structures are identified as the main contributor to
the wall pressure fluctuations.
While Fig.5.17 (b) shows a good correspondence for Cpv and the SPOD mode,
in Fig. 5.15, 5.16, and 5.18 (b), an extended coherent area can be observed. This
extended coherent area can be attributed to the large scale vortex of the shear layer
interacting with the turbulence structures in the mean flow. Therefore, the wall
Chapter 5. CFD Results 118

pressure fluctuations at this reference point (x/h = 0.5) can be considered to be


influenced by the combined effects of the shear layer structures and the large scale
turbulence interaction. These CFD results support the argument made in Section
4.4.2. As shown in Fig.4.17 (a), the wall pressure spectra measured at x/h = 0.5 dis-
play a power -3 decay rate, suggesting that the pressure fluctuations were produced
by both turbulence-mean-shear interactions and turbulence-turbulence interactions.
Figures 5.12 to 5.14 (c), (f) show Cpv at the second reference point, close to
the mean reattachment point at x/h ≈ xr1 for AR =2.4, 4, and 8 steps. It can be
seen that the high coherence level region mostly shifts outside of the separated shear
layer, indicating that the main contribution to the wall pressure at the reattachment
region comes from the large scale turbulence interaction. Again, this confirms the
-8/3 decay rate observed in Section 4.4.2 Fig.4.17 (b)&(c), which suggests that the
turbulence-turbulence interaction dominates the fluctuating wall pressure produc-
tion.
From the p0 ,v 0 coherence analysis, it can be seen that the wall pressure fluctu-
ations near the FBS leading edge, regarded as the main source of far-field noise,
are highly associated with the turbulent flow structures. The wall pressure fre-
quency spectrum given in Fig.4.20 shows the high spectral level at low-frequency
content (Stδ < 0.8) on the step top surface. At the same frequency range, the
flow field exhibits low-rank behaviour where the first SPOD mode is predominant.
The low-frequency pressure fluctuations are believed to be induced by the dominant
eigenmodes, which are identified as the separated shear layer structures, given the
evidence shown in the Cpv contours.
Chapter 5. CFD Results 119

Figure 5.15: AR = 1 step coherence coefficient Cpv maps for Stδ = 0.116 (top row) and
Stδ = 0.291 (second row). From left: First column, SPOD mode 1 at the corresponding
frequency; Second column, Cpv maps at x/h ≈ 0.5; Third column, Cpv maps at x/h ≈ 0.95.
Black lines: ∂u/∂y with 10 contour levels.

Figure 5.16: AR = 2.4 step coherence coefficient Cpv maps for Stδ = 0.233 (top row) and
Stδ = 0.349 (second row). From left: First column, SPOD mode 1 at the corresponding
frequency; Second column, Cpv maps at x/h ≈ 0.5; Third column, Cpv maps at x/h ≈ 2.
Black lines: ∂u/∂y with 10 contour levels.
Chapter 5. CFD Results 120

3 3 3
0.8

2 2 0.7 2

1 1 0.6 1

0.5
0 0 0
0 1 2 3 4 0 1 2 3 4 0 1 2 3 4
0.4
3 3 3
0.3

2 2 0.2 2

1 1 0.1 1

0
0 0 0
0 1 2 3 4 0 1 2 3 4 0 1 2 3 4

Figure 5.17: AR = 4 step coherence coefficient Cpv maps for Stδ = 0.233 (top row) and
Stδ = 0.407 (second row). From left: First column, SPOD mode 1 at the corresponding
frequency; Second column, Cpv maps at x/h ≈ 0.5; Third column, Cpv maps at x/h ≈ 2.
Black lines: ∂u/∂y with 10 contour levels.

Figure 5.18: AR = 8 step coherence coefficient Cpv maps for Stδ = 0.174 (top row) and
Stδ = 0.407 (second row). From left: First column, SPOD mode 1 at the corresponding
frequency; Second column, Cpv maps at x/h ≈ 0.5; Third column, Cpv maps at x/h ≈ 2.5.
Black lines: ∂u/∂y with 10 contour levels.
Chapter 5. CFD Results 121

5.7 Summary

This chapter has presented the velocity field obtained from numerical simulation
by applying LES modelling at Reh = 1.3 × 104 . The CFD results show reasonable
agreement with experimental steady and unsteady wall pressure at similar Reh . The
mean flow field around the steps is presented, displaying recirculating flow upstream,
on top and downstream of the step. The flow mean reattachment lengths identified
in the simulation are also comparable to the experimental results.
The separated shear layer was approximately indicated by the ∂u/∂y contour.
Two separated shear layers, emanating from the leading edge and the trailing edge,
are seen in AR = 2.4, 4, and 8 steps, while only one shear layer induced by the
step leading edge is observed for AR = 1 step. The two shear layers merging
phenomenon is more obvious for the small aspect ratio steps, indicating that the
downstream separation bubbles are affected by the upper corner separated shear
layer. Also, the turbulence statistics show that the high level TKE and Reynolds
stress components are coincident with the separated shear layer structures for all
step cases. This explains the critical role of the separated shear layer generated by
the sharp edges in flow modification and turbulence generation.
xr1
The instantaneous vorticity fields for AR = 1 and 8 steps show that for h
>
AR steps, the vortex formed at the leading edge will shed to the downstream flow
xr1
directly. While for h
< AR steps, the separated shear layer flapping motion can
be observed on the top surface.
The spectral proper orthogonal decomposition unveils the low-rank dynamics of
the flow field around the step. The most energetic eigenmodes are similar in shape
to the separated shear layer. The first SPOD modes are considered to be related
to the KH-roll up and vortex pairing, and the second SPOD modes show the KH-
type wavepackets are linked to the shear layer flapping motion. These two physical
interpretations for the SPOD modes are the main mechanisms that are responsible
to the large scale unsteadiness of the flow over the step with sharp edge.
Chapter 5. CFD Results 122

Finally, the coherence between the wall pressure fluctuations p0 and wall-normal
fluctuating velocity v 0 was calculated. The coherence coefficient maps show the
combined effects of the shear layer and turbulence interaction on the wall pressure.
The wall pressure spectral level is higher in the low frequency region (Stδ < 0.8).
This is the same frequency range that the flow low-rank behaviours are identified,
and the coherence maps present a similar shape to the first SPOD modes, indicating
the influences of the KH-roll up and vortex pairing mechanism on wall pressure
production.
Chapter 6

Forward-backward Facing Acoustic


Characteristics

6.1 Overview

In this chapter, the results of the acoustic measurements are presented. FBS with
AR = 1, 2, 4, and 8 were considered. The data collected by single microphone with
the observation angle θ = 83.4 are shown is Section 6.2. The sound spectra of
background noise (undisturbed flow) shows a spectral hump at 4 k-10 kHz which
is considered to be the facility noise. The sound spectra are broadband in nature
without any apparent tonal noise. The fixed observation angle spectra for different
flow speeds collapse reasonably using an AR−0.7 scaling. In Section 6.2.2, the FBS
data are compared with a FFS sound prediction model. The good agreement indi-
cates that the noise generation mechanism of FBS is similar to that of FFS, where
the geometry of the step front face is the most critical factor. Also, similar to FFS,
the dipole like directivity of FBS are shown in Section 6.2.3.
In Section 6.3, the acoustic results obtained using the microphone phased array
are presented. The acoustic beamforming maps show that the noise source of FBS
mainly located at the step leading edge across the frequency of interest (1 k-10 kHz).

123
Chapter 6. Forward-backward Facing Acoustic Characteristics 124

Also, in Section 6.3.2, the sound maps are spatially integrated to obtain the far-field
sound spectra. The integrated spectra indicates that the noise production is lower
xr1
when the reattachment is not seen on the top surface of FBS (AR < h
). This
is likely due to the lack of separated shear layer flapping motion on the step top
surface.
Chapter 6. Forward-backward Facing Acoustic Characteristics 125

6.2 Single Microphone Data

6.2.1 Background Noise and Step Induced Noise

The background noise was measured with the absence of the step at free stream
velocities of 20, 30, and 40 m/s. Figure 6.1 shows the background noise recorded
by the center microphone of the phased-array (observation angle θ = 83.4◦ , see Fig.
3.4) at different flow speeds. A small spectral hump is observed at around 4 kHz to
8 kHz and a spike is seen at 10 kHz. Apart from level, these spectral features do
not change with the incoming flow speed. The small spectral hump is suspected to
be the facility response which is not able to be removed and is also seen in the step
spectra. The 10 kHz spike is due to the fan motor controller noise.

60

50

40

30

20

10

-10
103 104

Figure 6.1: Sound spectra for background noise (undisturbed flow) at three different flow
speeds.

The far-field acoustic measurements were conducted for steps with AR = 1, 2, 4,


and 8. Figure 6.2 compares the background noise and the step induced noise mea-
sured at θ = 83.4◦ and U∞ = 40 m/s. It can be seen that the sound spectral level
are elevated above the background in all step cases. The signal-to-noise ratio (SNR)
(see Section 3.2.3) is used to determine the noise produced by the step. As shown in
Chapter 6. Forward-backward Facing Acoustic Characteristics 126

Figure 6.3, SNR for step noise and background noise is mostly above 2 dB between
1 kHz to 10 kHz (white region) for all step cases, indicating that the influence of the
step in far-field sound radiation mainly occurs in this frequency range. The same
frequency range of noise induced by FBS are also noticed by Springer et al.[74] who
considered an AR = 1 step. Thus, the following discussion and the sound spectra
results will only focus on this frequency range. Note that in Fig. 6.3 the SNR is
presented in 1/12th octave bands.

70 70

60 60

50 50

40 40

30 30

20 20

10 10
10 2 10 3 10 4 10 2 10 3 10 4

70 70
Undisturbed
60 60 With step

50 50

40 40

30 30

20 20

10 10
10 2 10 3 10 4 10 2 10 3 10 4

Figure 6.2: Narrow band sound spectra for different AR steps compared with undisturbed
flow at U∞ = 40 m/s. (a) Step1. (b) Step2. (c) Step4. (d) Step8.
Chapter 6. Forward-backward Facing Acoustic Characteristics 127

Figure 6.3: SNR for different AR steps (1/12th octave band) at U∞ = 40m/s. (a) Step1.(b)
Step2.(c) Step4.(d) Step8.

The background subtracted spectra obtained from a single microphone at an


83.4◦ observation angle for different steps are shown in Fig. 6.4. The results are
presented in 1/12-octave bands. The noise produced by the step is generally broad-
band. The sound spectra for different AR steps at different velocities have nearly
identical shape throughout the frequencies of interest. Awasthi et al. [8] suggested
that a spectral dip may occur when the half wavelength is half of the step height,
i.e. f = c0 /2h, which in this case corresponds to a frequency of 17 kHz. Such a
frequency is slightly outside of the SNR > 2 dB frequency region, yet a descending
trend in the spectra can still be observed at around 12-13 kHz.
Chapter 6. Forward-backward Facing Acoustic Characteristics 128

60 60

50 50

40 40

30 30

20 20

10 10

0 0

10 3 10 4 10 3 10 4

60 60

50 50

40 40

30 30

20 20

10 10

0 0

10 3 10 4 10 3 10 4

Figure 6.4: 1/12th octave band sound spectra at θ = 83.4◦ . (a) Step1.(b) Step2.(c)
Step4.(d) Step8.

The spectra show an almost flat shape at low frequencies followed by two small
humps peak at around 5 kHz and 10 kHz. These small humps/peaks are also seen
in the background sound spectra at the similar frequencies. It is believed that these
humps/peaks is due to the facility effects. It is supposed that the background noise
from the inlet is related to the facility and can not be removed by the subtraction.
The result shows the strong influence of flow velocity. At the highest mean flow
velocity, the sound level exceeds that at U∞ = 30m/s and 20m/s by 8 dB and 18
dB between 1 kHz to 10 kHz, respectively. A scaling law for the acoustic spectrum
at a fixed observation angle was proposed by Doolan and Moreau [19]. The non-
dimensional model was given as

Φr2
 
StL ∼ 10 log10 (6.1)
p20 M 3 LL
Chapter 6. Forward-backward Facing Acoustic Characteristics 129

where Φ is the acoustic pressure spectrum, r is the distance from the step leading
edge to the microphone, L is the width of step span, M is Mach number, and
p0 = ρU 2 , StL is the Strouhal number based on the characteristic length L of
the flow, which was selected to be the step height. The scaling law maintains
the 7th power velocity scaling proposed by Farabee and Zoccola [25]. The scaled
acoustics data are shown in Fig. 6.5. A good collapse of the data can be seen for
Sth < 1 for all step cases. For 1 < Sth ≤ 2, the scaling results are affected by the
humps/peaks caused by the non flow-related background noise and therefore show
some discrepancy. For Sth above 2, the scaling law does not collapse the data well
because the step is becoming acoustically non-compact. Catlett et al. [66] suggests
that the mixed scaling (frequency scaled on δ and sound spectrum scaled on step
height) may collapse the data better as the step height will affect the step acoustic
compactness in the high frequency region.
At a similar observation angle, Doolan and Moreau [19] observed an good data
collapse at Stδ1 between 0.3 to 2 for the FBS. This frequency range is similar to the
current study. It is worth noting that in Ref. [19], the boundary layer thickness
downstream of the step was chosen as the flow characteristic length whereas the step
height was selected here.
Chapter 6. Forward-backward Facing Acoustic Characteristics 130

-30 -30

-40 -40

-50 -50

-60 -60

-70 -70

-80 -80

-90 -90
10 -1 10 0 10 1 10 -1 10 0 10 1

-30 -30

-40 -40

-50 -50

-60 -60

-70 -70

-80 -80

-90 -90
10 -1 10 0 10 1 10 -1 10 0 10 1

Figure 6.5: Scaled 1/12th octave band sound spectra. (a) Step1. (b) Step2. (c) Step4.
(d) Step8.

Figure 6.6 (a) and (b) show the sound spectra measured for 4 different AR steps
at U∞ = 40 m/s, the scaled results using Eq.6.1, respectively. Since the step height
is the same for all testing cases, the step length l is chosen as the characteristic
length L here. It can be seen that the scaling law does not collapse the data well
for different AR at the same flow speed. The scaling law seems to capture the
velocity dependence of the sound spectra (Fig. 6.5) but does not sufficiently model
the dependence on step length (Fig. 6.6 (b)).
In the previous discussions (Chapter 4 and 5), the r.m.s. fluctuating wall pressure
xr1
level is found to be lower on the top surface of AR < h
steps due to the lack of
shear layer flapping motion at the wall. To address this, the reattachment length
xr1
to the step AR ratio h
/AR is considered in Equation 6.1. Meanwhile, since the
occurrence of the spectral dip is associated with the step geometry and independent
of the flow speed, the Helmholtz number (He) is used here instead of the Strouhal
Chapter 6. Forward-backward Facing Acoustic Characteristics 131

number. The new scaling law can be written as



xr1


Φr2 K
 K = 1,
 if /AR <1
h
HeL v.s. 10 log10 , (6.2)
p20 M 3 LL
K = ( xr1 /AR)n , if xr1


h h
/AR ≥1

where L is selected as the step height and n = 0.7 is found to provide the best
collapse. Note that for those small AR steps where the reattachment point is not
observed on the top surface, xr1 should be determined by the longer FBS that is
under similar flow conditions (Reh and h/δ). Here xr1 is chosen as 2h. Figure 6.6
(c) shows the results scaled on step height (h) using the new scaling law at U∞ = 40
m/s. It can be seen that the new scaling law successfully collapses the spectra
for different AR steps at 0.2 < Heh < 2. It also captures the the dependence on
the flow speed as shown in Fig. 6.6 (d). This suggests that once the step is long
xr1
enough to keep the reattachment point on its top surface, i.e. AR > h
, the far-
field sound pressure level is only dependent on the step height. However, for small
AR steps where the separated shear layer does not reattach on the top wall, the
noise generation is lower and is influenced by the step aspect ratio. As mentioned in
Chapter 5, the unsteadiness of the separated shear layer formed at the step leading
edge is determined by the KH-roll up & vortex shedding and shear layer flapping
xr1
motion. Both of these mechanisms are accountable for AR > h
steps wall pressure
fluctuations and therefore the far-field sound pressure levels are similar regardless
xr1
the step length. While for AR < h
steps, only the KH-roll up & vortex shedding
is apparent. The lack of shear layer flapping motion on the wall leads to the lower
fluctuating wall pressure and presumably results in lower levels of noise generation.
It should be noted that Eq.6.2 assumes the noise generation level is lower when
xr1 xr1
the AR is smaller for h
/AR < 1 step. Only one h
/AR < 1 case was investigated
in this study. Future work is required to fully valid the new scaling law.
Chapter 6. Forward-backward Facing Acoustic Characteristics 132

70
-50
60
-60
50
-70
40
-80
30
-90
20
10 3 10 4 10 0 10 1

-50 -50

-60 -60

-70 -70
U = 20m/s U=30 m/s U=40 m/s
Step1
-80 -80 Step1 Step1 Step1
Step2
Step2 Step2 Step2
Step4
Step4 Step4 Step4
-90 Step8 -90 Step8 Step8 Step8

10 0 10 0

Figure 6.6: Comparison of different scaling law. (a) Unscaled 1/12th octave band sound
spectra at U∞ =30m/s. (b) Scaled sound spectra at U∞ =30m/s using Eq.6.1. (c) Scaled
sound spectra at U∞ =30m/s using Eq.6.2. (d) Scaled sound spectra for all cases using
Eq.6.2.

6.2.2 Comparison with Analytical Model

Awasthi et al. [9] proposed an analytical model to predict the sound generation for
an unrounded forward facing step. The prediction formula for the far-field spectra
is given as

2 2
Kh ρ2 U∞
7
hA2

x1  khx2  x2  khx2 
Φpp (ω) = cos + sin (6.3)
(Rc0 f h2 ) R R R R

where Φpp is the spectral density of acoustic pressure (Pa2 /Hz), Kh is a constant
of 2.08 × 10−7 , U∞ is the flow speed, ρ is the air density, h is the step height, A is
the step front-face area, c0 is the speed of sound, f is the frequency, k = ω/c0 is
the acoustic wave number, and x1 , x2 , R are the corresponding distances from the
observer to the step (defined in Fig. 6.7).
Chapter 6. Forward-backward Facing Acoustic Characteristics 133

Figure 6.7: Configuration of the forward-facing step. Taken from Awasthi et al.[9].

The SPL value predicted by the FFS analytical model (Eq.6.3) and the acoustic
data for the FBS are compared in Fig. 6.8. The observation angle θ is 83.4◦ . The
comparison shows good agreements for Step2, Step4, and Step8. For Step1, the
measured values are around 5 dB less than the predicted value in the mid-frequency
range (3-6 kHz). Nonetheless, the prediction formula is able to capture the overall
trend and the SPL for each FBS. This indicates that the noise production mechanism
of the FBS resembles to that of the FFS. In the prediction formula, it can be seen
that the step front-face geometry (step height and area) plays a critical role for the
noise generation.
Figure 6.9 shows that the FBS data and the values predicted by the FFS model
collapse well using the new scaling law (Eq.6.2). In the FFS scenario, K in Eq.6.2
is always 1 as the reattachment point is guaranteed to be on the step top surface
and therefore all FFS data collapse to one curve in Fig. 6.9. The good data collapse
xr1
suggests that the far-field acoustic spectra for the FFS and the FBS with AR > h

are only dependent on the flow speed and the step height, and the step AR becomes
irrelevant once the flow is able to reattach on the top surface.
Chapter 6. Forward-backward Facing Acoustic Characteristics 134

60 60

50 50

40 40

30 30

20 20

10 10

0 0

10 3 10 4 10 3 10 4

60 60

50 50

40 40

30 30

20 20

10 10

0 0

10 3 10 4 10 3 10 4

Figure 6.8: 1/12th octave band sound spectra of the FBS compared with the FFS analyt-
ical model [9]. (a) Step1. (b) Step2. (c) Step4. (d) Step8.

-45

-50

-55

-60

-65

-70

-75
U=20 m/s U=30 m/s U=40 m/s
-80 Step1 Step1 Step1
Step2 Step2 Step2
-85 Step4 Step4 Step4
Step8 Step8 Step8
-90 FFS model FFS model FFS model

-95
100

Figure 6.9: 1/12th octave band sound spectra of the FBS compared with the FFS analyt-
ical model [9]
Chapter 6. Forward-backward Facing Acoustic Characteristics 135

6.2.3 Directivity

Previous studies [11, 39, 72] reported the streamwise-aligned dipole-like directiv-
ity with no noise radiated directly above the step edge for the FFS. Doolan and
Moreau[19] suggested that for the acoustically compact FBS, the sound source can
be regarded as the superpositions of compact dipoles. Awasthi et al. [9] proposed
an analytical model (Eq.6.3) showing that the directivity is no longer dipole-like at
higher frequencies where the FFS is not acoustically compact.
In this study, the directivity of the noise radiated by the FBS is investigated
using a microphone array. To compare with the analytical model prediction, the
directivity results are calculated using only microphones located around the central
area of the array (grey area in Fig. 6.10 (a), within 5 mm of the center microphone
in the z direction). The corresponding observation angles for each microphone are
given in Fig. 6.10 (b). The pressure spectral density is normalised on 4πr2 , where
r is the distance from the microphone to the leading edge location.

ο ο ο ο ο ο
100.1
ο ο
95.9 91.4 83.4 76.05 75.7 73.6 68.8

0.01 m

Flow direction

(b) Observation angles of the microphones.

(a) Microphone phased array. Red dots represent


the microphones used in Fig.6.10b

Figure 6.10: Directivity measurement setup

Figure 6.11a shows the directivity map at 1782 Hz (1/12th octave band) for
FBS obtained from the current measurement and for FFS predicted by Eq.6.3. The
overall directivity for the FBS and FFS is similar where the dipole like pattern with
Chapter 6. Forward-backward Facing Acoustic Characteristics 136

a dip above the leading edge can be observed. Downstream of the step leading edge
(θ < 90), a lobe is seen in every FBS case. Ji & Wang [39] claimed that the BFS noise
generation is diffraction-dominated and its sound level is lower than that of FFS,
which is dominated by the turbulence distortion and generation at the leading edge.
In Chapters 4 and 5, the numerical and wall pressure results show that remnants
of the turbulence created at the step leading edge remain in the downstream flow
field. The diffraction of this turbulence at the step trailing edge is considered the
secondary source in FBS noise generation. Despite being weaker than the source at
the leading edge, this secondary source may alter the acoustic field and therefore
affect directivity in the downstream region.
In Fig.6.12 at 7280 Hz, the FFS analytical model suggests a non-dipole like
behaviour at this frequency as the step is no longer acoustically compact. However,
the measurements present a similar pattern are seen in lower frequency (Fig.6.11)
where a dip is observed roughly above the step trailing edge. This is perhaps due
to the influences of the trailing edge noise source which alters the overall acoustic
field.
90
105 75

120 60

45

3 10-6

2 10-6

Step1
Step2 1 10-6
Step4
Step8
FFS model

Figure 6.11: Directivity map at 1728 Hz.


Chapter 6. Forward-backward Facing Acoustic Characteristics 137

90
105 75

120 60

45

2 10-6

Step1 1 10-6
Step2
Step4
Step8
FFS model
0

Figure 6.12: Directivity map at 7280 Hz.

6.3 Acoustic Beamforming

6.3.1 Acoustic Beamforming Maps

To visualize the location of the noise sources, the acoustic data were beamformed
using the conventional beamforming algorithm. The background noise was removed
in all cases. Figure 6.13 shows pre-subtracted and subtracted results for step8 at
5040 Hz (1/12th octave band). It can be seen that the background noise is mainly
from the wind tunnel inlet and the level is ∼ 8 dB lower than the step induced noise.
Background With Step (Unsubtracted) Background Subtracted
(undisturbed boundary layer)

Background noise
from tunnel inlet

Figure 6.13: Acoustic beamforming maps for background noise, pre-background-


subtraction, and background subtraction at 5040 Hz. The results are given for Step8
at Uinf ty = 40 m/s and the level is in dB.. Green rectangle indicates the location of the
step.
Chapter 6. Forward-backward Facing Acoustic Characteristics 138

The acoustic beamforming was performed in 1/12th octave bands. The beam-
forming maps for different AR steps at flow speed of 40 m/s are given in Figures 6.14
to 6.16. In the figures, the flow is coming from the left and the green rectangle repre-
sents the location of the steps. The results show that the noise produced by the FBS
is mainly located at the step leading edge. Note that due to the large beamwidth
of the sound map at lower frequency (Fig. 6.14), the source region is ambiguous
but still centred at the step leading edge. It can be observed that the sound maps
are nearly uniform in the spanwise direction indicating that the source is two di-
mensional. The sound maps for step1 are seen to be slightly concentrated at each
end of the step span. This is potentially due to the octagonal shape of the wind
tunnel, where the beamforming results sometimes appear curved due to additional
refraction at the corners. At high frequency, the sources for each step are clearly
seen concentrated at the leading edge, indicating that the leading edge dominates
the noise production in the FBS scenario. This is consistent with the CFD results
and wall pressure measurement results where the step leading edge is found to be
the most important factor of flow modification.
It should be noted that in the beamforming calculation, refraction effects, which
appear as a downstream shift in source position, were taken into account by using
correction term[61]. However, in [61] the correction method assumes the model
was placed in the middle of a potential core whereas in the current measurement
the steps were immersed in the boundary layer of the wall. The influences of the
thick boundary layer and separated shear layer on the refraction of the acoustic
waves is still unclear and further work is required to fully understand their affect on
beamforming.
Chapter 6. Forward-backward Facing Acoustic Characteristics 139

(a) (b)

(c) (d)

Figure 6.14: Acoustic beamforming maps at 2670 Hz at U∞ =40 m/s. Flow is form the
left and the level is in dB. Green rectangular indicate the step location. (a) Step1. (b)
Step2. (c) Step4. (d) Step8.
Chapter 6. Forward-backward Facing Acoustic Characteristics 140

(a) (b)

(c) (d)

Figure 6.15: Acoustic beamforming maps at 6350 Hz at U∞ =40 m/s. Flow is form the
left and the level is in dB. Green rectangular indicate the step location. (a) Step1. (b)
Step2. (c) Step4. (d) Step8.
Chapter 6. Forward-backward Facing Acoustic Characteristics 141

(b)
(a)

(c) (d)

Figure 6.16: Acoustic beamforming maps at 10079 Hz at U∞ =40 m/s. Flow is form the
left and the level is in dB. Green rectangular indicate the step location. (a) Step1. (b)
Step2. (c) Step4. (d) Step8.

6.3.2 Integrated Spectra

The acoustic beamforming map is spatially integrated over the source area to obtain
the far-field sound spectra. The integration region is shown in Fig.6.17 (blue dashed
rectangle). Due to the low SNR at high frequency region, some false noise sources
(sidelobes) are visible in the sound maps. To acquire the sound spectra only con-
sisting of step noise, these side lobes need to be excluded. Therefore, the integration
region is adopted to suit each frequency where a larger area is considered at lower
frequencies and a smaller region is applied at higher frequencies to include only the
relevant sources. For the present application, the integration regions for frequencies
below and above 4 kHz are shown in Fig.6.17 a and b, respectively.
Chapter 6. Forward-backward Facing Acoustic Characteristics 142

(a) (b)
0.5 m 0.3 m

0.24 m 0.24 m

Figure 6.17: Sound maps integration region at different frequencies. Frequencies below 4
kHz using (a), and frequencies above 4k Hz using (b).

Figure 6.18 (a) compares the integrated spectra using the fixed integration area
(using the large integration region in Fig.6.17a across the whole frequency range)
and dynamic integration area (switch to the smaller region for frequencies above 4
kHz). The integrated spectra is shown in 1/12th octave band for step8 at mean
flow speed of 40 m/s. It can be seen that the differences between two are marginal
throughout the frequency of interest (1k-10kHz). At higher frequencies (>10 kHz)
a minor offset are noticed as the large integration area may include the spurious
sources (Fig.6.17b).
Figure 6.18 (b) shows the integrated spectra for background noise, step8 back-
ground noise included, and step8 background noise subtracted. Note that in this
figure the fixed (large) integration region were used and the subtraction process was
performed before beamforming calculation. It can be noticed that the SNR in high
frequency region > 10 kHz) is higher in the integrated spectra as the beamforming
extend the frequency range of interest. Therefore, in the following discussion the
results will be presented between 1 kHz to 15kHz. Also, spectral humps can be
spotted above 3.5 kHz in all integrated spectra. As mentioned in Section 6.2, this
is speculated to be the facility response.
Chapter 6. Forward-backward Facing Acoustic Characteristics 143

60 60

50 50

40 40

30 30

Background (undisturbed flow)


20 Fixed Integration Region 20 With step (baskground included)
Dynamic Integration Region With step (background subtracted)
10 10
103 104 103 104

(a) (b)

Figure 6.18: 1/12th octave band integrated spectra (Pa2 /Hz) for Step8 at U∞ =40 m/s.
(a) Influences of integration region. (b) Comparison integrated spectra of subtracted,
unsubtracted, and background noise.

Figure 6.19 shows the 1/12th octave band integrated spectra (in Pa2 /Hz) for
different AR steps at flow speed of 20, 30, and 40 m/s. The integrated spectra
show a nearly identical shape at different mean stream velocities where two small
spectral humps are observed at 4-8 kHz and 10 kHz followed by a downward trend at
even higher frequency. This downward trend is associated with the step height and
seemed to be independent to step aspect ratio as no apparent difference is noticed
with the varied AR.
It can be seen that between 1.5 kHz-8 kHz, the spectral level of smallest AR
step (Step1) is ∼ 5dB lower than the other cases at different velocities. The lower
r.m.s value of wall pressure fluctuations are also recorded for step1 (Section 4.4). As
mentioned in Section 5.6.1, the unsteadiness of the separated shear layer created at
the leading edge is controlled by KH vorticity roll up and shear layer flapping. For
the small AR step where no reattachment point is seen on the step top surface, the
vortex will be shed directly downstream and the shear layer flapping motion will
blend into the wake instead of oscillating over the wall. This potentially leads to
lower fluctuating wall pressure levels. And the lower far-field acoustic spectral level
xr1
in this small AR step (AR < h
) can be attributed to the lack of this shear layer
flapping motion on the top surface.
For Step2 (AR = 2), since the reattachment length (xr1 ) is close to the step
Chapter 6. Forward-backward Facing Acoustic Characteristics 144

length, the shear layer flapping motion does not always stay on the top surface.
This possibly results in 1∼2 dB lower spectral level between 1.5 kHz - 3.5 kHz
compare to step4 and step8 where the step is long enough to keep the reattachment
point on the top surface.
40 50

30 40

20 30

10 20

0 10

-10 0
10 3 10 4 10 3 10 4

60

step1
50
step2
step4
40
step8

30

20

10
10 3 10 4

Figure 6.19: Integrated spectra (Pa2 /Hz) at different flow speeds. (a) U∞ =20 m/s. (b)
U∞ =30 m/s. (c) U∞ =40 m/s.

The new scaling law (Eq.6.2) were applied to the integrated spectra. The distance
from the observer to the testing model r in Eq.6.2 is determined using the averaged
distance (1.3 m) for the microphone array. Reasonable data collapses for U∞ = 40
m/s and for all testing cases can be observed in Fig.6.20.
To summarize, the influences of the FBS aspect ratio on integrated sound spectra
are visible only when the step length is shorter than the reattachment length. For
xr1
smaller AR step when AR < h
, the separated shear layer flapping motion are
not reattaching to the top surface, which leads to a lower level of r.m.s fluctuating
Chapter 6. Forward-backward Facing Acoustic Characteristics 145

wall pressure and lower levels noise production. However, despite the lack of the
shear layer flapping motion over the wall, the difference in far-field spectra is still
not significant. This is perhaps because the the main mechanism of KH vortex roll
up and shedding, which is controlled by the leading edge geometry. In the current
study, the leading edge geometry for all testing cases are similar; thus, the far-field
power spectra in Fig.6.19 are also similar.

-40 -40

-50 -50

-60 -60

-70 -70

-80 -80 U=20 m/s U=30 m/s U=40m/s


Step1
Step1 Step1 Step1
Step2
Step2 Step2 Step2
-90 Step4 -90 Step4 Step4 Step4
Step8
Step8 Step8 Step8
-100 -100
10 0 10 0

Figure 6.20: Scaled integrated spectra. (a) U∞ =40 m/s. (b) All testing cases.
Chapter 6. Forward-backward Facing Acoustic Characteristics 146

6.4 Summary

This chapter has provided the results of acoustic measurements. The step influences
in noise production was found mainly between 1 and 10k Hz. The sound spectra
obtained using a single microphone shows that the noise produced by the step is
generally broadband without any distinct tonal noise.
The step aspect ratio is included in the scaling law proposed by Doolan &
Moreau[19] to address the effects of the step length. The new scaling law sug-
gests that the far-field acoustic spectra is not dependant on the step aspect ratio
once the reattachment point is able to stay on the FBS top surface ( xhr1 /AR ≤ 1).
xr1
While for h
/AR > 1 steps, the data collapse reasonably well when scaled with
( xhr1 /AR)−0.7 . Also, the single microphone spectra for FBS were compared with the
value predicted by an analytical model[9] for FFS. The quantitatively good agree-
ments suggests that the noise generation mechanism for FBS is similar to FFS where
the geometry of FBS leading edge is the most critical factor. Meanwhile, the direc-
tivity maps show a dipole-like pattern for FBS with a dip roughly above the step
leading edge. Downstream the step, the FBS displays a different directivity pattern
from FFS. It is speculated that the noise generated by the rear edge of the FBS,
despite the low level, alters the acoustic field downstream of the step.
The acoustic beamforming maps at different frequencies identified the same noise
source locations of FBS at the leading edge regardless the step AR, which indicates
that the leading edge dominates the noise production of FBS. The sound source
location (around leading edge) is consistent with the locations of highest level of TKE
(Section 5.4) and the highest level of r.m.s. wall pressure fluctuations (Section 4.4.1),
implying that the FBS induce noise is highly associated with the large turbulence
modification cause by the leading edge.
Finally, the sound maps are spatially integrated to yield the integrated spectra
of the FBS. The integrated spectra shows that the influences of the FBS aspect
ratio are visible only when the step length is shorter than the reattachment length
Chapter 6. Forward-backward Facing Acoustic Characteristics 147

xr1
(AR < h
). As discussed in Chapter 5, the separated shear layer flapping motion
does not occur on the step top surface when the reattachment point is absent. The
lack of this separated shear layer flapping motion potentially leads to the lower
fluctuating wall pressure and the lower spectral level for small AR steps.
Chapter 7

Conclusions

7.1 Summary of Work

This thesis examines the turbulent flow field and the acoustic features of the FBS
with different aspect ratios using a combined experimental and numerical approach.
The experimental setup, numerical method, and the data analysis techniques have
been described in Chapter 3.
Insight to the flow field features were explored using the wall pressure measure-
ments and the CFD simulation. The flow reattachment length and the unsteady wall
pressure measurements are presented in Chapter 4 for AR = 1, 2, 4, and 8 steps.
Results of the LES calculations are given in Chapter 5 to elucidate the turbulent
flow fields. In the LES calculation, AR = 1, 2.4, 4, and 8 steps are considered and
the numerical model was validated by the experimental data.
Chapter 6 discusses the far-field acoustic characteristics of the FBS. The acoustic
beamforming maps identify the locations of the noise source. The sound spectra
obtained from a fixed observation angle and integrated from the beamforming maps
are presented, and the influences of the AR are justified combined with the evidence
observed in the flow field.

148
Chapter 7. Conclusions 149

7.2 Main Findings and Achievement

• Flow field separation-reattachment features of the FBS and the lead-


ing edge influences on the step downstream area

The flow separation-reattachment region around the FBS was investigated.


The upstream separation lengths were found independent of the step aspect
ratio with the similar values recorded for different AR steps. The separation
lengths were seen to be around 0.7 ∼ 0.8 step heights in CFD simulation
and 1.5 ∼ 1.6 step heights in oil flow visualization measurements. The differ-
ences are believed to be due to the different upstream flow conditions. The
first downstream reattachment length xr1 was found between 1.8 ∼ 2.5 step
heights in both numerical and experimental results and the values are weakly
dependent on the step AR. The second downstream reattachment length (xr2 )
measured from the step trailing edge was longer for smaller AR ratio steps and
xr1
remained at the nearly constant value of ∼ 5 step heights for AR > h
steps.
Similar observations were also made by Bergeles & Athanassiadis who claimed
that the downstream reattachment length (xr2 ) varies with step aspect ratio in
xr1
a negative linear relationship for AR < h
steps and stays almost consistent
xr1
for AR > h
steps.

• Step AR influence on the flow field

The scale of the FBS AR compared to the step reattachment length on its
top surface ( xhr1 ) was found to be a critical factor determining the overall
flow field characteristics. The flow unsteadiness on the step top surface is
controlled by the KH-roll up and vortex pairing mechanism and separated
shear layer flapping motion. The instantaneous vorticity flow fields indicate
xr1
that for AR < h
steps, the separated shear layer is not able to flap on the
top surface and will merge into the downstream wake area. The lack of this
shear layer flapping motion results in lower r.m.s fluctuating wall pressure on
its top surface.
Chapter 7. Conclusions 150

xr1
In the step downstream area, the spanwise coherence lengths for AR < h
xr1
steps are generally larger than that for AR > h
steps as the large scale
structures in the separated shear layer are able to mix into the downstream flow
field. This also enhances the turbulence kinetic energy and the wall pressure
fluctuations downstream of the step as the high momentum structures are
preserved in the flow without being broken down by reattachment on the top
surface.

• SPOD identifies low rank behaviour

The SPOD analysis identified the low-rank behaviour of the flow field around
the step leading edge at frequency between 0.06 < Stδ < 1, indicating the first
SPOD mode dominance in this area. The first SPOD mode is considered to be
associated with the KH-roll up and vortex pairing mechanism of the separated
shear layer, which is found to be coherent with the wall pressure fluctuations
near the leading edge (x = 0.5h). This indicates that the unsteadiness of the
separated shear layer is the main contributor to the wall pressure fluctuations.
Also, due to the first mode dominance, this KH-roll up and vortex pairing
mechanism is regarded as the primary effect of the turbulence modification in
the near leading edge region.

• Step leading edge dominates the noise generation

The far-field sound spectra of the FBS display a good agreement with the FFS
model [9] predicted results, implying that the noise production mechanism of
the FBS resembles that of the FFS, where the geometry of the leading edge
is the most critical factor. The acoustic beamforming maps also show that
the noise sources of the FBS are mainly located at its leading edge, where the
highest TKE and highest r.m.s fluctuating wall pressure were recorded. This
indicates that the leading edge of the FBS dominates both flow field alteration
and far-field sound generation.

• FBS far-field acoustic features


Chapter 7. Conclusions 151

The frequency range of the noise generated by the FBS is 1 kHz-10 kHz. The
far-field sound spectra of the FBS is in general broadband regardless of the
aspect ratio. The spectra present a downward trend in high frequency region
as the step becomes acoustic non-compact. The FBS show a dipole like far-
field sound directivity with a dip observed roughly above the step leading edge.
However downstream the step, the FBS displays a different directivity pattern
from FFS. It is speculated that although the level is lower, the noise generated
by the rear edge of the FBS changes the acoustic field downstream of the step.

• The influence of FBS aspect ratio on far-field acoustic features

The sound spectra shows that the influences of the FBS aspect ratio is only
noticeable when the step length is shorter than the reattachment length. For
xr1
AR < h
steps, the separated shear layer flapping motion does not occur on
the top surface and therefore results in lower wall pressure fluctuations. The
lack of this separated shear layer flapping motion is believed to be the reason
xr1
for the lower level of noise production for small AR steps (AR < h
).

A new scaling law for FBS is proposed, which separates the FBS into two
xr1
categories. For longer steps (AR > h
), the step AR seems to be irrelevant.
xr1
While for small AR steps (AR < h
), the data collapse reasonably well when
scaled with (xr1 /AR)−0.7 .

7.3 Recommendations for Future Work

• The noise production for FBS with AR < xr1


h
has been found to be lower than
xr1
large AR step (AR > h
). The new scaling law proposed in the study sug-
gested that the data scale well with (xr1 /AR)−0.7 for small AR steps. However,
xr1
since only one case with AR < h
was considered in the measurements, more
investigations regarding these small AR steps are required. A measurement
xr1
specifically targeting the AR < h
steps can help understand how the even
smaller AR alter the overall flow field.
Chapter 7. Conclusions 152

• In the present study, an incompressible numerical model is adopted to cal-


culate the flow field. The far-field sound radiation can be further calculated
using integral methods. The results can be validated by comparing with the
experimental data collected in this study.

• The Mach number effect of the opened test section wind tunnel was taken into
account using the correction term[61] in the present beamforming calculation.
The correction term assumes the sound source is located in the mean flow;
however, in this study the model is immersed in the boundary layer. The in-
fluences of the thick boundary layer and separated shear layer on the refraction
of the acoustic waves need to be examined to fully understand their affect on
beamforming.

• In the current study, the FBS is two-dimensional where it is considered in-


finitely long in the spanwise direction. When the step spanwise length be-
comes comparable to the height, the effects of edges on the sides need to be
taken into account. The three-dimensional flow features such as the horseshoe
vortex will further alter the turbulent flow field. Further studies can focus on
the aeroacoustic features of this three-dimensional effects of the FBS.
Bibliography
[1] Y. Addad, D. Laurence, C. Talotte, and M. C. Jacob. Large eddy simulation of
a forward–backward facing step for acoustic source identification. International
Journal of Heat and Fluid Flow, 24(4):562–571, 2003. ISSN 0142-727X.

[2] K. Akselvoll and P. Moin. Large-eddy simulation of turbulent confined coan-


nular jets. Journal of Fluid Mechanics, 315:387–411, 1996. ISSN 0022-1120.

[3] M. Awasthi. High Reynolds Number Turbulent Boundary Layer Flow over Small
Forward Facing Steps. Thesis, Virginia Polytechnic Institute and State Univer-
sity, 2012.

[4] M. Awasthi. Sound Radiated from Turbulent Flow over Two and Three-
Dimensional Surface Discontinuities. Phd thesis, Virginia Polytechnic Institute
and State University, 2015.

[5] M. Awasthi, W. Devenport, S. Glegg, and J. Forest. Pressure fluctuations


produced by forward steps immersed in a turbulent boundary layer. Journal of
Fluid Mechanics, 756:384–421, 2014.

[6] M. Awasthi, W. J. Devenport, W. N. Alexander, and S. A. L. Glegg. Aeroa-


coustics of rounded forward-facing steps: Near-field behavior. AIAA Journal,
57(3):1237–1249, 2018. ISSN 0001-1452.

[7] M. Awasthi, J. Rowlands, D. J. Moreau, and C. J. Doolan. Two-step hybrid

153
Bibliography 154

calibration of remote microphones. The Journal of the Acoustical Society of


America, 144(5):EL477–EL483, 2018.

[8] M. Awasthi, W. J. Devenport, W. N. Alexander, and S. A. L. Glegg. Aeroa-


coustics of rounded forward-facing steps: Far-field behavior. AIAA Journal, 57
(5):1899–1913, 2019.

[9] M. Awasthi, W. J. Devenport, and S. A. L. Glegg. Sound radiated by turbulent


flow over unrounded and rounded forward-facing steps. Journal of Sound and
Vibration, 488:115635, 2020. ISSN 0022-460X.

[10] J. Bardina, J. Ferziger, and W. Reynolds. Improved subgrid-scale models for


large-eddy simulation. Fluid Dynamics and Co-located Conferences. American
Institute of Aeronautics and Astronautics, 1980.

[11] S. Becker, M. Escobar, C. Hahn, I. Ali, M. Kaltenbacher, B. Basel, and


M. Grunewald. Experimental and Numerical Investigation of the Flow Induced
Noise from a Forward Facing Step. Aeroacoustics Conferences. American In-
stitute of Aeronautics and Astronautics, 2005.

[12] G. Bergeles and N. Athanassiadis. The flow past a surface-mounted obstacle.


Journal of Fluids Engineering, 105(4):461–463, 1983. ISSN 0098-2202.

[13] R. Camussi, M. Felli, F. Pereira, G. Aloisio, and A. Di Marco. Statistical


properties of wall pressure fluctuations over a forward-facing step. Physics of
Fluids, 20(7):075113, 2008. ISSN 1070-6631.

[14] C. Chandrsuda, R. D. Mehta, A. D. Weir, and P. Bradshaw. Effect of free-


stream turbulence on large structure in turbulent mixing layers. Journal of
Fluid Mechanics, 85(4):693–704, 1978. ISSN 0022-1120.

[15] F. H. Clauser. Turbulent boundary layers in adverse pressure gradients. Journal


of the Aeronautical Sciences, 21(2):91–108, 1954.

[16] Y. Co. Educational wind tunnel (ewt), Jun 2019.


Bibliography 155

[17] G. M. Corcos. The resolution of turbulent pressures at the wall of a boundary


layer. Journal of Sound and Vibration, 6(1):59–70, 1967. ISSN 0022-460X.

[18] N. Curle and M. J. Lighthill. The influence of solid boundaries upon aerody-
namic sound. Proceedings of the Royal Society of London. Series A. Mathemat-
ical and Physical Sciences, 231(1187):505–514, 1955.

[19] C. Doolan and D. Moreau. Flow-induced noise generated by sub-boundary layer


steps. Experimental Thermal and Fluid Science, 72, 2015.

[20] C. Doolan, D. Moreau, M. Awasthi, and C. Jiang. The unsw anechoic wind
tunnel. In WESPAC.

[21] D. M. Driver, H. L. Seegmiller, and J. G. Marvin. Time-dependent behavior of


a reattaching shear layer. AIAA Journal, 25(7):914–919, 1987. ISSN 0001-1452.

[22] J. K. Eaton and J. P. Johnston. Low-frequency unsteadyness of a reattaching


turbulent shear layer. page 162.

[23] X. Fang and M. F. Tachie. On the unsteady characteristics of turbulent sepa-


rations over a forward–backward-facing step. Journal of Fluid Mechanics, 863:
994–1030, 2019. ISSN 0022-1120.

[24] X. Fang and M. F. Tachie. Spatio-temporal dynamics of flow separation induced


by a forward-facing step submerged in a thick turbulent boundary layer. Journal
of Fluid Mechanics, 892:A40, 2020. ISSN 0022-1120.

[25] T. Farabee and P. Zoccola. Experimental evaluation of noise due to flow over
surface steps. In 1998 ASME International Mechanical Engineering Congress
and Exposition.

[26] T. M. Farabee and M. J. Casarella. Effects of surface irregularity on turbu-


lent boundary layer wall pressure fluctuations. ASME Journal of Vibration
Acoustics, 106:343, 1984. ISSN 0739-3717.
Bibliography 156

[27] J. E. Ffowcs Williams, D. L. Hawkings, and M. J. Lighthill. Sound generation


by turbulence and surfaces in arbitrary motion. Philosophical Transactions of
the Royal Society of London. Series A, Mathematical and Physical Sciences,
264(1151):321–342, 1969.

[28] W. K. George, P. D. Beuther, and R. E. A. Arndt. Pressure spectra in turbulent


free shear flows. Journal of Fluid Mechanics, 148:155–191, 1984. ISSN 0022-
1120.

[29] S. A. Glegg, B. Bryan, W. J. Devenport, and M. Awasthi. The Noise From


Separated Flow. AIAA AVIATION Forum. American Institute of Aeronautics
and Astronautics, 2014.

[30] M. Goody. Empirical spectral model of surface pressure fluctuations. AIAA


Journal, 42(9):1788–1794, 2004.

[31] A. Graziani, F. Kerhervé, R. J. Martinuzzi, and L. Keirsbulck. Dynamics of


the recirculating areas of a forward-facing step. Experiments in Fluids, 59(10):
154, 2018. ISSN 1432-1114.

[32] G. Green. An essay on the application of mathematical analysis to the theories


of electricity and magnetism. 1828.

[33] Y. Guan, C. R. Berntsen, M. J. Bilka, and S. C. Morris. The measurement


of unsteady surface pressure using a remote microphone probe. Journal of
visualized experiments : JoVE, (118):53627, 2016. ISSN 1940-087X.

[34] H. Hattori and Y. Nagano. Investigation of turbulent boundary layer over


forward-facing step via direct numerical simulation. International Journal of
Heat and Fluid Flow, 31(3):284–294, 2010. ISSN 0142-727X.

[35] A. F. Heenan and J. F. Morrison. Passive control of pressure fluctuations


generated by separated flow. AIAA Journal, 36(6):1014–1022, 1998. ISSN
0001-1452.
Bibliography 157

[36] M. S. Howe. Sound produced by turbulent boundary layer flow over a finite
region of wall roughness, and over a forward facing step. Journal of Fluids and
Structures, 3(1):83–96, 1989. ISSN 0889-9746.

[37] L. M. Hudy, A. Naguib, and W. M. Humphreys. Stochastic estimation of a


separated-flow field using wall-pressure-array measurements. Physics of Fluids,
19(2):024103, 2007. ISSN 1070-6631.

[38] M. C. Jacob, A. Louisot, D. Juvé, and S. Guerrand. Experimental study of


sound generated by backward-facing steps under wall jet. AIAA Journal, 39
(7):1254–1260, 2001. ISSN 0001-1452.

[39] M. Ji and M. Wang. Sound generation by turbulent boundary-layer flow over


small steps. Journal of Fluid Mechanics, 654:161–193, 2010. ISSN 0022-1120.

[40] M. Ji and M. Wang. Surface pressure fluctuations on steps immersed in tur-


bulent boundary layers. Journal of Fluid Mechanics, 712:471–504, 2012. ISSN
0022-1120.

[41] M. Kiya and K. Sasaki. Structure of a turbulent separation bubble. Journal of


Fluid Mechanics, 137:83–113, 1983. ISSN 0022-1120.

[42] M. Kiya and K. Sasaki. Structure of large-scale vortices and unsteady reverse
flow in the reattaching zone of a turbulent separation bubble. Journal of Fluid
Mechanics, 154:463–491, 1985. ISSN 0022-1120.

[43] P. A. Krogstad, J. H. Kaspersen, and S. Rimestad. Convection velocities in a


turbulent boundary layer. Physics of Fluids, 10(4):949–957, 1998. ISSN 1070-
6631.

[44] Y. C. Küçükosman, C. Schram, and J. van Beeck. A remote microphone tech-


nique for aeroacoustic measurements in large wind tunnels. Applied Acoustics,
129:346–353, 2018. ISSN 0003-682X.
Bibliography 158

[45] Y. C. Küçükosman, C. Schram, and J. van Beeck. A remote microphone tech-


nique for aeroacoustic measurements in large wind tunnels. Applied Acoustics,
129:346–353, 2018. ISSN 0003-682X.

[46] D. C. Lander, C. W. Letchford, M. Amitay, and G. A. Kopp. Influence of


the bluff body shear layers on the wake of a square prism in a turbulent flow.
Physical Review Fluids, 1(4):044406, 2016.

[47] J. F. Largeau and V. Moriniere. Wall pressure fluctuations and topology in


separated flows over a forward-facing step. Experiments in Fluids, 42(1):21,
2006. ISSN 1432-1114.

[48] H. Le, P. Moin, and J. Kim. Direct numerical simulation of turbulent flow over
a backward-facing step. Journal of Fluid Mechanics, 330:349–374, 1997. ISSN
0022-1120.

[49] D. Leclerq, M. Jacob, A. Louisot, and C. Talotte. Forward-backward facing


step pair - aerodynamic flow, wall pressure and acoustic characterisation. 2001.

[50] I. Lee and H. J. Sung. Characteristics of wall pressure fluctuations in separated


and reattaching flows over a backward-facing step:part i. time-mean statistics
and cross-spectral analyses. Experiments in Fluids, 30(3):262–272, 2001. ISSN
1432-1114.

[51] M. J. Lighthill and M. H. A. Newman. On sound generated aerodynamically i.


general theory. Proceedings of the Royal Society of London. Series A. Mathe-
matical and Physical Sciences, 211(1107):564–587, 1952.

[52] J. L. Lumley. The structure of inhomogeneous turbulent flows. Atmospheric


Turbulence and Radio Wave Propagation, 1967.

[53] D. G. Mabey. Analysis and correlation of data on pressure fluctuations in


separated flow. Journal of Aircraft, 9(9):642–645, 1972.
Bibliography 159

[54] D. M. Moore, C. W. Letchford, and M. Amitay. Energetic scales in a bluff body


shear layer. Journal of Fluid Mechanics, 875:543–575, 2019. ISSN 0022-1120.

[55] W. D. Moss and S. Baker. Re-circulating flows associated with two-dimensional


steps. Aeronautical Quarterly, 31(3):151–172, 1980.

[56] T. J. Mueller. Aeroacoustic measurements. Springer, Berlin ; New York, 2002.

[57] A. Nematollahi and M. F. Tachie. Time-resolved piv measurement of influ-


ence of upstream roughness on separated and reattached turbulent flows over
a forward-facing step. AIP Advances, 8(10):105110, 2018.

[58] F. Nicoud and F. Ducros. Subgrid-scale stress modelling based on the square of
the velocity gradient tensor. Flow, Turbulence and Combustion, 62(3):183–200,
1999. ISSN 1573-1987.

[59] T. Ota, Y. Asano, and J.-i. Okawa. Reattachment length and transition of the
separated flow over blunt flat plates. Bulletin of JSME, 24(192):941–947, 1981.

[60] D. Palumbo. Determining correlation and coherence lengths in turbulent


boundary layer flight data. Journal of Sound and Vibration, 331(16):3721–
3737, 2012. ISSN 0022-460X.

[61] T. PaOPTDOIs, C. Prax, and V. Valeau. Numerical validation of shear flow


corrections for beamforming acoustic source localisation in open wind-tunnels.
Applied Acoustics, 74(4):591–601, 2013. ISSN 0003-682X.

[62] D. S. Pearson, P. J. Goulart, and B. Ganapathisubramani. Turbulent separation


upstream of a forward-facing step. Journal of Fluid Mechanics, 724:284–304,
2013. ISSN 0022-1120.

[63] Z. Prime, C. Doolan, and B. Zajamsek. Beamforming array optimisation and


phase averaged sound source mapping on a model wind turbine. 11 2014.

[64] L. F. Richardson and J. A. Gaunt. The deferred approach to the limit. part
i. single lattice. part ii. interpenetrating lattices. Philosophical Transactions of
Bibliography 160

the Royal Society of London Series A, 226:299, 1927. ISSN 1364-503X0080-


46140962-8436.

[65] P. J. Roache. Quantification of uncertainty in computational fluid dynamics.


Annual Review of Fluid Mechanics, 29(1):123–160, 1997.

[66] M. Ryan Catlett, W. Devenport, and S. A. L. Glegg. Sound from boundary


layer flow over steps and gaps. Journal of Sound and Vibration, 333(18):4170–
4186, 2014. ISSN 0022-460X.

[67] O. T. Schmidt, A. Towne, G. Rigas, T. Colonius, and G. A. Brès. Spectral


analysis of jet turbulence. Journal of Fluid Mechanics, 855:953–982, 2018.
ISSN 0022-1120.

[68] F. SchÄFer, M. Breuer, and F. Durst. The dynamics of the transitional flow
over a backward-facing step. Journal of Fluid Mechanics, 623:85–119, 2009.
ISSN 0022-1120.

[69] M. Sherry, D. Jacono, J. Sheridan, R. Mathis, and I. Marusic. Flow separation


characterisation of a forward facing step immersed in a turbulent boundary
layer. Sixth International Symposium on Turbulence and Shear Flow Phenom-
ena, 2009.

[70] M. Sieber, C. O. Paschereit, and K. Oberleithner. Spectral proper orthogonal


decomposition. Journal of Fluid Mechanics, 792:798–828, Mar 2016. ISSN
1469-7645.

[71] L. W. Sigurdson. The structure and control of a turbulent reattaching flow.


Journal of Fluid Mechanics, 298:139–165, 1995. ISSN 0022-1120.

[72] J. Slomski. Numerical Investigation of Sound from Turbulent Flow over a For-
ward Facing Step. Aeroacoustics Conferences. American Institute of Aeronau-
tics and Astronautics, 2011.
Bibliography 161

[73] P. G. Spazzini, G. Iuso, M. Onorato, N. ZOPTURLo, and G. M. Di Cicca.


Unsteady behavior of back-facing step flow. Experiments in Fluids, 30(5):551–
561, 2001. ISSN 1432-1114.

[74] M. Springer, C. Scheit, and S. Becker. Flow-induced noise of a forward-


backward facing step. In A. Dillmann, G. Heller, E. Krämer, C. Wagner, and
C. Breitsamter, editors, New Results in Numerical and Experimental Fluid Me-
chanics X, pages 767–776. Springer International Publishing. ISBN 978-3-319-
27279-5.
Appendices

162
Appendix A

Microphone Details for Wall Pres-


sure Measurements
The 1/4” GRAS 40PH microphone were used in the unsteady wall pressure measure-
ment (Chapter 4). Details of each microphone such as the serial number, sensitivity
and the testing location are given in this Appendix.

163
Appendix A. Microphone Details for Wall Pressure Measurements 164

A.1 Undisturbed Flow

A.1.1 Streamwise Measurement Locations

x/h z/h Microphone Serial No. Sensitivity (V /P a)


-15 0 223099 0.05664284
-8 0 223073 0.04709581
-4 0 222978 0.04794154
-2 0 222981 0.05298306
-1 0 213645 0.05339791
-0.5 0 223101 0.05324862
0.6 0 223060 0.05458556
1.2 0 223062 0.05265562
1.8 0 213600 0.05109956
2.4 0 223039 0.0543242
3 0 222989 0.05418316
3.6 0 213638 0.05127358
4.2 0 213642 0.04767384
4.8 0 223040 0.05416149
5.4 0 223080 0.05122234
5.9 0 223070 0.05216302
6.4 0 223083 0.04918456
6.9 0 223085 0.04729401
7.4 0 223053 0.04756909
7.9 0 222251 0.04658997
9.4 0 213595 0.04911576
12.4 0 223096 0.0505406
18.4 0 213635 0.05228835
24.4 0 213583 0.0545201
30.4 0 213593 0.0480087
36.4 0 213602 0.04333638
43.4 0 223052 0.05875416
50.4 0 213640 0.05131461

Table A.1: Undisturbed flow streamwise microphone locations


Appendix A. Microphone Details for Wall Pressure Measurements 165

A.1.2 Spanwise Measurement Locations

x/h z/h Microphone Serial No. Sensitivity (V /P a)


-4 0 223099 0.05664284
-4 0.2 223073 0.04709581
-4 0.5 222978 0.04794154
-4 1 222981 0.05298306
-4 2 213645 0.05339791
-0.5 0 223060 0.05458556
-0.5 0.2 223062 0.05265562
-0.5 0.5 213600 0.05109956
-0.5 1 223039 0.0543242
-0.5 2 222989 0.05418316
2.4 0 223116 0.04449558
2.4 0.2 223084 0.05209005
2.4 0.5 222251 0.04658997
2.4 1 223102 0.05866024
2.4 2 223112 0.05407491
6.4 0 223080 0.05122234
6.4 0.2 223070 0.05216302
6.4 0.5 223083 0.04918456
6.4 1 223085 0.04729401
6.4 2 223053 0.04756909

Table A.2: Undisturbed flow spanwise microphone Locations


Figure A.1: Wall pressure measurement locations for undisturbed flow. Red dashed line indicates step leading edge location (x = 0).
Appendix A. Microphone Details for Wall Pressure Measurements
166
Appendix A. Microphone Details for Wall Pressure Measurements 167

A.2 Step1

A.2.1 Streamwise Measurement locations

x/h z/h Microphone Serial No. Sensitivity (V /P a)


-15 0 223099 0.05664284
-8 0 223073 0.04709581
-4 0 222978 0.04794154
-2 0 222981 0.05298306
-1 0 213645 0.05339791
-0.5 0 223101 0.05324862
0.5 0 223049 0.0574988
1.2 0 223060 0.05458556
1.8 0 223062 0.05265562
2.4 0 213600 0.05109956
3 0 223039 0.0543242
3.6 0 222989 0.05418316
4.2 0 213638 0.05127358
4.8 0 213642 0.04767384
5.4 0 223040 0.05416149
6 0 223080 0.05122234
6.5 0 223070 0.05216302
7 0 223083 0.04918456
7.5 0 223085 0.04729401
8 0 223053 0.04756909
8.5 0 222251 0.04658997
10 0 213595 0.04911576
13 0 223096 0.0505406
19 0 213635 0.05228835
25 0 213583 0.0545201
31 0 213593 0.0480087
37 0 213602 0.04333638
44 0 223052 0.05875416
51 0 213640 0.05131461

Table A.3: Step1 streamwise microphone locations


Appendix A. Microphone Details for Wall Pressure Measurements 168

A.2.2 Spanwise Measurement Locations

x/h z/h Microphone Serial No. Sensitivity (V /P a)


-4 0 223099 0.05664284
-4 0.2 223073 0.04709581
-4 0.5 222978 0.04794154
-4 1 222981 0.05298306
-4 2 213645 0.05339791
-0.5 0 223060 0.05458556
-0.5 0.2 223062 0.05265562
-0.5 0.5 213600 0.05109956
-0.5 1 223039 0.0543242
-0.5 2 222989 0.05418316
0.5 0 222793 0.05628152
0.5 0.2 223109 0.05801858
0.5 0.5 223131 0.05158212
0.5 1 223113 0.05479336
0.5 2 222254 0.04841363
3 0 223116 0.04449558
3 0.2 223084 0.05209005
3 0.5 222251 0.04658997
3 1 223102 0.05866024
3 2 223112 0.05407491
7 0 223080 0.05122234
7 0.2 223070 0.05216302
7 0.5 223083 0.04918456
7 1 223085 0.04729401
7 2 223053 0.04756909

Table A.4: Step1 spanwise microphone locations


Figure A.2: Wall pressure measurement locations for Step1. Red dashed line indicates step leading edge, black dashed line indicates step trailing
edge.
Appendix A. Microphone Details for Wall Pressure Measurements
169
Appendix A. Microphone Details for Wall Pressure Measurements 170

A.3 Step2

A.3.1 Streamwise Measurement Locations

x/h z/h Microphone Serial No. Sensitivity (V /P a)


-15 0 223099 0.05664284
-8 0 223073 0.04709581
-4 0 222978 0.04794154
-2 0 222981 0.05298306
-1 0 213645 0.05339791
-0.5 0 223101 0.05324862
0.5 0 222793 0.05628152
1 0 223109 0.05801858
1.5 0 223131 0.05158212
2.2 0 223060 0.05458556
2.8 0 223062 0.05265562
3.4 0 213600 0.05109956
4 0 223039 0.0543242
4.6 0 222989 0.05418316
5.2 0 213638 0.05127358
5.8 0 213642 0.04767384
6.4 0 223040 0.05416149
7 0 223080 0.05122234
7.5 0 223070 0.05216302
8 0 223083 0.04918456
8.5 0 223085 0.04729401
9 0 223053 0.04756909
9.5 0 222251 0.04658997
11 0 213595 0.04911576
14 0 223096 0.0505406
20 0 213635 0.05228835
26 0 213583 0.0545201
32 0 213593 0.0480087
38 0 213602 0.04333638
45 0 223052 0.05875416
52 0 213640 0.05131461

Table A.5: Step2 streamwise microphone locations


Appendix A. Microphone Details for Wall Pressure Measurements 171

A.3.2 Spanwise Measurement Locations

x/h z/h Microphone Serial No. Sensitivity (V /P a)


-4 0 223099 0.05664284
-4 0.2 223073 0.04709581
-4 0.5 222978 0.04794154
-4 1 222981 0.05298306
-4 2 213645 0.05339791
-0.5 0 223060 0.05458556
-0.5 0.2 223062 0.05265562
-0.5 0.5 213600 0.05109956
-0.5 1 223039 0.0543242
-0.5 2 222989 0.05418316
1 0 222793 0.05628152
1 0.2 223109 0.05801858
1 0.5 223131 0.05158212
1 1 223113 0.05479336
1 2 222254 0.04841363
4 0 223116 0.04449558
4 0.2 223084 0.05209005
4 0.5 222251 0.04658997
4 1 223102 0.05866024
4 2 223112 0.05407491
8 0 223080 0.05122234
8 0.2 223070 0.05216302
8 0.5 223083 0.04918456
8 1 223085 0.04729401
8 2 223053 0.04756909

Table A.6: Step2 spanwise microphone locations


Figure A.3: Wall pressure measurement locations for Step2. Red dashed line indicates step leading edge, black dashed line indicates step trailing
edge.
Appendix A. Microphone Details for Wall Pressure Measurements
172
Appendix A. Microphone Details for Wall Pressure Measurements 173

A.4 Step4

A.4.1 Streamwise Measurement Locations

x/h z/h Microphone Serial No. Sensitivity (V /P a)


-15 0 223099 0.05664284
-8 0 223073 0.04709581
-4 0 222978 0.04794154
-2 0 222981 0.05298306
-1 0 213645 0.05339791
-0.5 0 223101 0.05324862
0.5 0 222793 0.05628152
1 0 223109 0.05801858
1.5 0 223131 0.05158212
2 0 223113 0.05479336
2.5 0 222254 0.04841363
3 0 223135 0.05004777
3.5 0 223130 0.05522237
4.2 0 223060 0.05458556
4.8 0 223062 0.05265562
5.4 0 213600 0.05109956
6 0 223039 0.0543242
6.6 0 222989 0.05418316
7.2 0 213638 0.05127358
7.8 0 213642 0.04767384
8.4 0 223040 0.05416149
9 0 223080 0.05122234
9.5 0 223070 0.05216302
10 0 223083 0.04918456
10.5 0 223085 0.04729401
11 0 223053 0.04756909
11.5 0 222251 0.04658997
13 0 213595 0.04911576
16 0 223096 0.0505406
22 0 213635 0.05228835
Appendix A. Microphone Details for Wall Pressure Measurements 174

28 0 213583 0.0545201
34 0 213593 0.0480087
40 0 213602 0.04333638
47 0 223052 0.05875416
54 0 213640 0.05131461

Table A.7: Step4 streamwise microphone locations


Appendix A. Microphone Details for Wall Pressure Measurements 175

A.4.2 Spanwise Measurement Locations

x/h z/h Microphone Serial No. Sensitivity (V /P a)


-4 0 223099 0.05664284
-4 0.2 223073 0.04709581
-4 0.5 222978 0.04794154
-4 1 222981 0.05298306
-4 2 213645 0.05339791
-0.5 0 223060 0.05458556
-0.5 0.2 223062 0.05265562
-0.5 0.5 213600 0.05109956
-0.5 1 223039 0.0543242
-0.5 2 222989 0.05418316
2 0 222793 0.05628152
2 0.2 223109 0.05801858
2 0.5 223131 0.05158212
2 1 223113 0.05479336
2 2 222254 0.04841363
6 0 223116 0.04449558
6 0.2 223084 0.05209005
6 0.5 222251 0.04658997
6 1 223102 0.05866024
6 2 223112 0.05407491
10 0 223080 0.05122234
10 0.2 223070 0.05216302
10 0.5 223083 0.04918456
10 1 223085 0.04729401
10 2 223053 0.04756909

Table A.8: Step4 spanwise microphone locations


Figure A.4: Wall pressure measurement locations for Step4. Red dashed line indicates step leading edge, black dashed line indicates step trailing
edge.
Appendix A. Microphone Details for Wall Pressure Measurements
176
Appendix A. Microphone Details for Wall Pressure Measurements 177

A.5 Step8

A.5.1 Streamwise Measurement Locations

x/h z/h Microphone Serial No. Sensitivity (V /P a)


-15 0 223099 0.05664284
-8 0 223073 0.04709581
-4 0 222978 0.04794154
-2 0 222981 0.05298306
-1 0 213645 0.05339791
-0.5 0 223101 0.05324862
0.5 0 223049 0.0574988
1 0 222793 0.05628152
1.5 0 223109 0.05801858
2 0 223131 0.05158212
2.5 0 223113 0.05479336
3 0 222254 0.04841363
3.5 0 223135 0.05004777
4 0 223130 0.05522237
4.5 0 223116 0.04449558
5 0 223084 0.05209005
5.5 0 222251 0.04658997
6 0 223102 0.05866024
6.5 0 223112 0.05407491
7 0 223103 0.05057093
7.5 0 213585 0.05764848
8.2 0 223060 0.05458556
8.8 0 223062 0.05265562
9.4 0 213600 0.05109956
10 0 223039 0.0543242
10.6 0 222989 0.05418316
11.2 0 213638 0.05127358
11.8 0 213642 0.04767384
12.4 0 223040 0.05416149
13 0 223080 0.05122234
Appendix A. Microphone Details for Wall Pressure Measurements 178

13.5 0 223070 0.05216302


14 0 223083 0.04918456
14.5 0 223085 0.04729401
15 0 223053 0.04756909
15.5 0 222251 0.04658997
17 0 213595 0.04911576
20 0 223096 0.0505406
26 0 213635 0.05228835
32 0 213583 0.0545201
38 0 213593 0.0480087
44 0 213602 0.04333638
51 0 223052 0.05875416
58 0 213640 0.05131461

Table A.9: Step8 streamwise microphone locations


Bibliography 179

A.5.2 Spanwise Measurement Locations

x/h z/h Microphone Serial No. Sensitivity (V /P a)


-4 0 223099 0.05664284
-4 0.2 223073 0.04709581
-4 0.5 222978 0.04794154
-4 1 222981 0.05298306
-4 2 213645 0.05339791
-0.5 0 223060 0.05458556
-0.5 0.2 223062 0.05265562
-0.5 0.5 213600 0.05109956
-0.5 1 223039 0.0543242
-0.5 2 222989 0.05418316
4 0 222793 0.05628152
4 0.2 223109 0.05801858
4 0.5 223131 0.05158212
4 1 223113 0.05479336
4 2 222254 0.04841363
10 0 223116 0.04449558
10 0.2 223084 0.05209005
10 0.5 222251 0.04658997
10 1 223102 0.05866024
10 2 223112 0.05407491
14 0 223080 0.05122234
14 0.2 223070 0.05216302
14 0.5 223083 0.04918456
14 1 223085 0.04729401
14 2 223053 0.04756909

Table A.10: Step8 spanwise microphone locations


Bibliography

Figure A.5: Wall pressure measurement locations for Step8. Red dashed line indicates step leading edge, black dashed line indicates step trailing
edge.
180

You might also like