Download as pdf or txt
Download as pdf or txt
You are on page 1of 202

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/34174370

Large-scale oscillatory stabilisation of the boundary layer : a thesis submitted


in partial fulfilment of the requirements for the degree of Master of
Mechanical Engineering, Depar...

Article
Source: OAI

CITATIONS READS

0 155

1 author:

Reuben D Rusk
Mindquip Limited
6 PUBLICATIONS   109 CITATIONS   

SEE PROFILE

All content following this page was uploaded by Reuben D Rusk on 27 May 2014.

The user has requested enhancement of the downloaded file.


Large-Scale Oscillatory Stabilisation of
the Boundary Layer

Reuben D. Rusk

A thesis submitted in partial fulfilment of the requirements for the Degree of


Master of Mechanical Engineering

Department of Mechanical Engineering


University of Canterbury
Canterbury
New Zealand

31st of May, 2004


Typeset in LaTeX using MiKTeX 2.4.

ii
Abstract

In recent decades, considerable research has greatly improved understanding of bound-


ary layer structure, stability mechanisms, and turbulence regeneration. The under-
standing gained from this research has led to new methods for control these dynamics.
Spanwise oscillations induced in the turbulent boundary layer are known to result in
considerable modification of the coherent structures present. Forcing is thought to
interfere with the instability mechanisms which drive turbulence regeneration. The
large drag reductions reported in many experiments suggest such control may be
useful on water vessels and aircraft.
The present work reviews turbulent boundary layer structure, the mechanisms
of transition, and turbulence phenomena to provide understanding into the mech-
anisms governed by oscillations. The linear stability of laminar boundary layers is
analysed, both for non-oscillatory and oscillatory flows. Using a forcing profile de-
rived analytically for a spanwise travelling surface wave in Blasius flow, it is found
low frequency streamwise oscillations with peak oscillatory velocity above 20% of the
free-stream can significantly increase the critical Reynolds number. Spanwise oscil-
lations are found to have a similar effect, dramatically stabilising three-dimensional
instability modes. It appears the stability of sufficiently large-amplitude oscillatory
flow dominates over the instability of the boundary layer, thereby stabilising the flow.
The frequency at which stabilisation occurs corresponds to frequencies at which drag
reduction has been observed using spanwise oscillations in turbulent boundary layers.
Higher-frequency oscillations, however, are found to destabilise the boundary layer.
Results suggest that the drag reduction observed due to spanwise forcing in turbu-
lent boundary layers is a consequence of a generic stabilisation mechanism by oscilla-
tory forcing. This hypothesis is supported by recent studies indicating an instability-
based turbulence regeneration mechanism. Further research is required to validate
these results and investigate the stability of forced turbulent flows and the effect of
forcing on transient growth.

iii
Table 1: Nomenclature
Latin Letters Description
a Non-dimensional forcing wavelength / 2π
A Non-dimensional forcing amplitude
D Chebyshev differentiation matrix
E transformed wall-normal disturbance vorticity
k Index of vector, matrix or mode
M Number of Floquet modes
N Number of spectral collocation points
p disturbance pressure
P Pressure
Re Reynolds number based on boundary layer thickness
T Chebyshev transformation matrix
u streamwise disturbance velocity
U streamwise velocity
V transformed wall-normal disturbance velocity
V wall-normal velocity
v wall-normal disturbance velocity
V wall-normal velocity
w spanwise disturbance velocity
W spanwise velocity
Wo Womersly number
x streamwise distance
y wall-normal distance
y∗ Nondimensional wall-normal distance
z spanwise distance
Greek Letters Description
α streamwise disturbance wavenumber
β spanwise disturbance wavenumber
δ boundary layer thickness / Mathieu equation constant
 Mathieu equation forcing amplitude
η vorticity
κ vertical decay rate of forcing pressure
λ Floquet exponent / or forcing wavelength
µ absolute fluid viscosity
∇ Divergence
ν kinematic fluid viscosity
ω disturbance frequency
Ω Nondimensional forcing frequency
ρ fluid density
ξi transformed wall-normal domain, −1 ≤ ξi ≤ 1

iv
Table 2: Nomenclature (continued)

Indices Description
0
(prime) Derivative in the wall-normal direction
∞ Free-stream value
e(tilde) Denotes a vector of values over computational domain
ˆ(hat) Denotes a vector of spectral coefficients
¯(bar) Oscillatory component
i Imaginary part
k Collocation point value / or mode number
r Real part
rms Root-mean-squared value
U Pertaining to streamwise velocity
W Pertaining to spanwise velocity
x Streamwise component
y Wall-normal component
z Spanwise component
Abbreviations Description
BLC Boundary Layer Control
DNS Direct Numerical Simulation
LEBU Large-Eddy Break Up device
LTSE Linearised Time-Dependent Stability Equations
MEMS Micro-Electro-Mechanical-System
OSE Orr-Sommerfeld Equation
PIV Particle Image Velocimetry
PSE Parabolised Stability Equations
STG Streak Transient Growth
TS Tollmien-Schlichting

v
Acknowledgement

The author wishes to thank Dr. J. Geoffrey Chase1 and Professor Tim David2 for their
supervision, helpful guidance, useful discussions and valuable feedback. Many thanks
also to my Father, family, and friends for their support, guidance and encouragement.

1
Sr. Lecturer, Dept of Mechanical Engineering, University of Canterbury
2
Professor, Dept of Mechanical Engineering, University of Canterbury

vi
Preface

This thesis has three main objectives. First, it provides a complete review of flow
control technology and turbulence phenomena, presented in Parts I and II respec-
tively. Part I provides an introduction to the boundary layer and addresses current
flow control technologies and applications. Chapter 1 briefly introduces boundary
layer fluid dynamics and terminology, and Chapter 2 introduces and discusses the
importance of flow control, and current methods and progress of flow control.
Part II reviews current understanding of turbulence phenomena, with particular
attention to the instability mechanisms, regeneration, and structure of turbulence.
Chapter 3 outlines the transition processes from laminar to turbulent flow, which
covers boundary layer receptivity, primary and secondary instabilities, bypass mech-
anisms and transient growth. The coherent fluid structures of turbulence are discussed
in Chapter 4, and Chapter 5 outlines the regeneration mechanisms of these structures.
The second objective of this thesis is to investigate the influence of boundary layer
oscillation, given in Part III. It is known that spanwise oscillation reduces turbulence
production and therefore drag in turbulent boundary layers, and a brief review of
experimental and direct numerical simulation results is presented in Chapter 6. To
investigate the effect of spanwise oscillations on the stability of various modes in the
boundary layer, a Matlab code is developed for investigating the linear stability of
the laminar boundary layer based on the analysis given in Chapter 7.
Little theoretical work has been done on the stabilising effect of parametric os-
cillations in boundary layer flows, likely due to large computational cost. However,
increases in computational power are making simple analysis possible. The linearised
time-dependent equations are derived for simple oscillatory parallel flows. The re-
sults of this analysis are presented in Chapter 8 and validated against the work of
several others, including Orszag (1971) and Schmid and Henningson (2001). Turbu-
lent boundary layers are far more complex in structure however, so it is likely that
a three-dimensional nonlinear analysis will be required to fully understand turbulent
boundary layer stabilisation due to oscillations. Such a non-linear analysis, based on
the present work, is left for future work.
Finally, although it is known that spanwise oscillations are effective, practical

vii
actuation technology is currently in its infancy. Wall oscillation in the form of a trav-
elling wave is one of the most promising methods for large-scale drag reduction. The
author is unaware of relevant previous research on the flow-field generated from such
actuation, however, to successfully design and optimise an actuator. It is desirable to
know the flow field induced, hence Part IV presents an analytically derived equation
to determine the flow field for small amplitudes of oscillation in Chapter 9. Using
a finite-element numerical method described in Chapter 10 for validation, results of
this analysis are presented in Chapter 11.

viii
Contents

I Introduction 1

1 Introduction 2
1.1 The boundary layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Problem definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Wavelike disturbances . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2 Flow control 11
2.1 Purpose . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2.1 Suction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.2.2 Compliant coatings . . . . . . . . . . . . . . . . . . . . . . . . 16
2.2.3 Fixed small-scale structures . . . . . . . . . . . . . . . . . . . 17
2.2.4 Micro-electro-mechanical technologies . . . . . . . . . . . . . . 18
2.2.5 Micro-bubbles . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.2.6 Polymer addition . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2.7 Riblets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

II Turbulence phenomena 23

3 Transition to turbulence 24
3.1 Transition pathways . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.2 Receptivity processes . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.3 Primary mode instabilities . . . . . . . . . . . . . . . . . . . . . . . . 28
3.4 Secondary instability mechanisms . . . . . . . . . . . . . . . . . . . . 28
3.4.1 Two-dimensional TS wave secondary instability . . . . . . . . 29
3.4.2 Streak secondary instability . . . . . . . . . . . . . . . . . . . 33
3.4.3 Oblique secondary instability . . . . . . . . . . . . . . . . . . 33
3.5 Bypass mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.6 Transient growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

ix
4 Coherent structures 39
4.1 Streamwise vortices . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.2 Boundary layer streaks . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.3 Hairpin vortices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

5 Turbulence regeneration 46
5.1 Parent-offspring scenarios . . . . . . . . . . . . . . . . . . . . . . . . 46
5.1.1 Hairpin vortex formation . . . . . . . . . . . . . . . . . . . . . 47
5.1.2 Vorticity sheet generation and roll-up . . . . . . . . . . . . . . 47
5.2 Instability-based regeneration . . . . . . . . . . . . . . . . . . . . . . 48
5.2.1 Wave-shear instabilities . . . . . . . . . . . . . . . . . . . . . . 48
5.2.2 Streak instability . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.2.3 Streak transient growth . . . . . . . . . . . . . . . . . . . . . 50
5.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

III Oscillatory Stabilisation 53

6 Review of oscillatory forcing 54


6.1 Mechanism of drag reduction . . . . . . . . . . . . . . . . . . . . . . . 56
6.2 Prediction of drag reduction . . . . . . . . . . . . . . . . . . . . . . . 57
6.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

7 Linear stability analysis 59


7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
7.2 Governing equations . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
7.3 Chebyshev spectral method . . . . . . . . . . . . . . . . . . . . . . . 64
7.3.1 Chebyshev collocation and differentiation . . . . . . . . . . . . 64
7.3.2 Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . 65
7.4 Temporal formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
7.5 Spatial formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
7.5.1 4th order formulation . . . . . . . . . . . . . . . . . . . . . . . 69
7.5.2 Second order formulation . . . . . . . . . . . . . . . . . . . . . 70
7.6 Time dependent formulation . . . . . . . . . . . . . . . . . . . . . . . 72
7.6.1 Introduction to parametric forcing . . . . . . . . . . . . . . . . 72
7.6.2 Linearised stability equations . . . . . . . . . . . . . . . . . . 75
7.7 Solution method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

x
8 Stability analysis results 79
8.1 Temporal formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
8.1.1 Plane Couette flow . . . . . . . . . . . . . . . . . . . . . . . . 79
8.1.2 Plane Poiseuille flow . . . . . . . . . . . . . . . . . . . . . . . 80
8.1.3 Blasius boundary layer flow . . . . . . . . . . . . . . . . . . . 81
8.2 Spatial formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
8.2.1 Plane Poiseuille flow . . . . . . . . . . . . . . . . . . . . . . . 88
8.2.2 Blasius boundary layer flow . . . . . . . . . . . . . . . . . . . 88
8.3 Time-dependent temporal formulation . . . . . . . . . . . . . . . . . 90
8.3.1 Streamwise forcing . . . . . . . . . . . . . . . . . . . . . . . . 91
8.3.2 Spanwise forcing . . . . . . . . . . . . . . . . . . . . . . . . . 95
8.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
8.4.1 Unforced flow results . . . . . . . . . . . . . . . . . . . . . . . 102
8.4.2 Forced flow results . . . . . . . . . . . . . . . . . . . . . . . . 102
8.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
8.6 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

IV Actuation method 111

9 Analytical solution 112


9.1 Problem specification . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
9.2 Nonlinear formulation . . . . . . . . . . . . . . . . . . . . . . . . . . 114
9.3 Derivation of approximated governing equation . . . . . . . . . . . . 117
9.4 Stokes oscillating plate solution . . . . . . . . . . . . . . . . . . . . . 122

10 2-Dimensional numerical simulation 123


10.1 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
10.1.1 Notation of tensors and derivatives . . . . . . . . . . . . . . . 123
10.1.2 Conservation equations . . . . . . . . . . . . . . . . . . . . . . 125
10.1.3 Mass conservation . . . . . . . . . . . . . . . . . . . . . . . . . 125
10.1.4 Momentum conservation . . . . . . . . . . . . . . . . . . . . . 126
10.1.5 Thermal energy conservation . . . . . . . . . . . . . . . . . . . 126
10.1.6 Formulation of the discrete problem . . . . . . . . . . . . . . . 126
10.1.7 Derivation of matrix coefficients . . . . . . . . . . . . . . . . . 127
10.2 Solution method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
10.2.1 Non-dimensionalisation . . . . . . . . . . . . . . . . . . . . . . 129
10.2.2 Domain and boundary conditions . . . . . . . . . . . . . . . . 129
10.2.3 Numerical scheme . . . . . . . . . . . . . . . . . . . . . . . . . 130

xi
11 Actuation results 132
11.1 Stokes oscillating plate solution . . . . . . . . . . . . . . . . . . . . . 132
11.2 Travelling-wave forcing . . . . . . . . . . . . . . . . . . . . . . . . . . 133
11.3 Errors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
11.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
11.5 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

Bibliography 142

Appendices 157

A Matlab codes 157


A.1 stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
A.1.1 stability.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
A.1.2 makematrix.m . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
A.1.3 postprocess.m . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
A.1.4 cheb.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
A.1.5 couette.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
A.1.6 poiseuille.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
A.1.7 blasius.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
A.1.8 chebtransform.m . . . . . . . . . . . . . . . . . . . . . . . . . 166
A.1.9 solveblasiusODE.m . . . . . . . . . . . . . . . . . . . . . . . . 167
A.2 FIDAP read-file generator . . . . . . . . . . . . . . . . . . . . . . . . 168

B Additional wave-forcing results 173

xii
List of Figures

1.1 Three methods of boundary layer actuation. . . . . . . . . . . . . . . 4


1.2 Typical visualisation of a turbulent boundary layer structure in plan. 5
1.3 Boundary layer definition showing streamwise, spanwise and wall-normal
components. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Blasius boundary layer profile. . . . . . . . . . . . . . . . . . . . . . . 6
1.5 Instability wave propagating at an angle to the mean flow. . . . . . . 8
1.6 Common modes of instability in the boundary layer. . . . . . . . . . . 10

2.1 Flow control objectives and methods. . . . . . . . . . . . . . . . . . . 12


2.2 A Micro-Electro-Mechanical-System (MEMS) fluid control actuator. . 18
2.3 The riblet structure of shark skin. . . . . . . . . . . . . . . . . . . . . 21

3.1 Transition mechanisms in wall bounded flows. . . . . . . . . . . . . . 25


3.2 Particle Image Velocimetry (PIV) visualisations of oblique (left), H-
type (center), and K-type (right) transition. . . . . . . . . . . . . . . 29
3.3 Fundamental secondary instability of Tollmien-Schlichting waves. . . 30
3.4 Subharmonic secondary instability of Tollmien-Schlichting waves. . . 31
3.5 Relationship between streaks and secondary instability modes. . . . . 31
3.6 Contours of streamwise disturbance velocity for Tollmien-Schlichting
wave secondary instability. . . . . . . . . . . . . . . . . . . . . . . . . 32
3.7 Contours of streamwise disturbance velocity for secondary instability
of Görtler streaks. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.8 Sinuous and varicose secondary instability of a synthetic low-speed streak. 35
3.9 Contours of streamwise disturbance velocity for oblique secondary in-
stability. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.10 Flow visualisation showing a typical turbulent spot in a zero-pressure
gradient boundary layer. . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.11 Illustration of transient growth due to the nonorthogonal superposition
of two vectors that decay at different rates. . . . . . . . . . . . . . . . 37

4.1 Formation of a low-speed streak by streamwise vortices. . . . . . . . . 40

xiii
4.2 Typical low-speed streak structure. . . . . . . . . . . . . . . . . . . . 42
4.3 Hairpin vortex structure. . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.4 Typical hairpin vortex structure with Reynolds number. . . . . . . . 44
4.5 Varicose streak instability and breakdown. . . . . . . . . . . . . . . . 45

6.1 Drag reduction correlation by Choi et al. (2002). . . . . . . . . . . . . 57

7.1 Stability diagram for Mathieu’s equation. . . . . . . . . . . . . . . . . 72


7.2 Eigenvalues of Mathieu’s equation. . . . . . . . . . . . . . . . . . . . 73

8.1 Orr-Sommerfeld spectrum for plane Couette flow. . . . . . . . . . . . 80


8.2 Orr-Sommerfeld spectrum for plane Poiseuille flow. . . . . . . . . . . 81
8.3 Contours of constant growth rate and constant phase velocity of Poiseuille
flow for β = 0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
8.4 Maximum growth and phase of the least stable mode of Poiseuille flow
for Re = 10000 with non-zero β. . . . . . . . . . . . . . . . . . . . . . 83
8.5 Typical temporal eigenvalue distribution for the Blasius boundary layer. 84
8.6 Maximum growth and phase speed of the least stable mode for the
Blasius boundary layer for β = 0. . . . . . . . . . . . . . . . . . . . . 85
8.7 Maximum growth and phase of the least stable mode for the Blasius
boundary layer for Re = 1000 with non-zero β. . . . . . . . . . . . . . 86
8.8 Eigenfunction corresponding to the least stable growth rate of the Bla-
sius boundary layer at Re = 500, α = 0.2 and β = 0. . . . . . . . . . 87
8.9 Spatial eigenvalue spectrum of streamwise wavenumbers of Poiseuille
flow for Re = 2000, ω = 0.3, β = 0 and N = 200. . . . . . . . . . . . . 88
8.10 Plots of the spatial eigenvalue spectrum for Blasius flow for Re = 1000,
ω = 0.26, β = 0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
8.11 Typical appearance of the eigenvalue spectrum for the linearised time-
dependent stability equation. . . . . . . . . . . . . . . . . . . . . . . . 90
8.12 Stability curves of Blasius flow for streamwise forcing of amplitude 0.05
and Ω = 0.18. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
8.13 Stability curves of Blasius flow for streamwise forcing at Re = 700 and
α = 0.25. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
8.14 Forcing profile A. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
8.15 Comparison of different number of collocation points and modes used
for Re = 700. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
8.16 Stability curves of Blasius flow for streamwise forcing of amplitude 0.05
and 0.1 with Ω = 0.06. . . . . . . . . . . . . . . . . . . . . . . . . . . 96

xiv
8.17 Stability curves of Blasius flow for streamwise forcing of amplitude 0.2,
frequency of 0.4; and amplitude 0.03 with Ω = 0.1. . . . . . . . . . . 97
8.18 Forcing profile B. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
8.19 Stability curves of Blasius flow for streamwise forcing of amplitude 0.3
and Ω = 0.04. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
8.20 Accuracy comparison of stability curves of Blasius flow for streamwise
forcing of amplitude 0.2 and Ω = 0.03, using different numbers of modes.100
8.21 Growth rates of Blasius flow for spanwise forcing at Re = 700 and
α = 0.25. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
8.22 Oblique transition delay found by Berlin (1998) using oscillating-wall
forcing. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

9.1 Domain mapping. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114


9.2 Numerical results showing exponential pressure decay. . . . . . . . . . 120

10.1 Computational mesh for numerical simulation. . . . . . . . . . . . . . 130


10.2 Smooth ramp-up of amplitude used in simulations. . . . . . . . . . . 131

11.1 Stokes oscillating-plate solution obtained. . . . . . . . . . . . . . . . . 132


11.2 Typical stream function calculated for a Stokes oscillating plate. . . . 133
11.3 Flow field for Re = 102 at wave amplitudes of 0.005 and 0.020. . . . . 135
11.4 Flow field for Re = 102 at wave amplitudes of 0.035 and 0.050. . . . . 136
11.5 Flow field for Re = 103 and Re = 104 at wave amplitude of 0.050. . . 137
11.6 Analytical streamfunction calculated for Reynolds number of 102 and
wave amplitude of 0.050. . . . . . . . . . . . . . . . . . . . . . . . . . 138
11.7 Profiles of mean (quasi-steady) spanwise velocity induced in the nu-
merical solution for Re = 102 and Re = 105 . . . . . . . . . . . . . . . 139
11.8 Profiles of mean (quasi-steady) spanwise velocity induced in the nu-
merical solution at t = 60. . . . . . . . . . . . . . . . . . . . . . . . . 139

B.1 Additional plots of the actuated flow field at Re = 102 , wave amplitude
is 0.005. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
B.2 Additional plots of the actuated flow field at Re = 102 , wave amplitude
is 0.020. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
B.3 Additional plots of the actuated flow field at Re = 102 , wave amplitude
is 0.035. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
B.4 Actuated flow field at Re = 101 , wave amplitude is 0.050. . . . . . . . 177
B.5 Addition plots of the actuated flow field at Re = 101 , wave amplitude
is 0.050. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178

xv
B.6 Addition plots of the actuated flow field at Re = 102 , wave amplitude
is 0.050. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
B.7 Addition plots of the actuated flow field at Re = 103 , wave amplitude
is 0.050. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
B.8 Addition plots of the actuated flow field at Re = 104 , wave amplitude
is 0.050. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
B.9 Actuated flow field at Re = 105 , wave amplitude is 0.050. . . . . . . . 182
B.10 Addition plots of the actuated flow field at Re = 105 , wave amplitude
is 0.050. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183

xvi
List of Tables

1 Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv
2 Nomenclature (continued) . . . . . . . . . . . . . . . . . . . . . . . . v

xvii
xviii
Part I

Introduction

1
Chapter 1

Introduction

The boundary layer is the layer of fluid near the wall of external flows, such as
those over the skin of aircraft or water vessels. The instability mechanisms in the
boundary layer are responsible for transition to turbulent flow and likely contribute
to the regeneration of turbulence (Schoppa and Hussain, 2002). If these instability
mechanisms can be hindered or stabilised it could yield drag reduction in both form
drag and skin friction. One of the most promising methods of stabilisation involves
large-scale oscillatory forcing of the fluid near the wall of the boundary layer.
While little research has been undertaken on streamwise forcing, there is a large
body of research relating to spanwise oscillations. Such spanwise forcing has been
accomplished by use of wall oscillation (e.g. Choi and Clayton, 2001; Choi et al.,
2002; Dhanak and Si, 1999), electromagnetic Lorentz forcing (e.g. Berger et al., 2000;
Weier et al., 2000) and wavelike motion of the surface (e.g. Du et al., 2002; Karniadakis
and Choi, 2003). Each of these methods is shown schematically in Figure 1.1. These
methods do not require complex control methods or often delicate actuators, while
still offering significant drag reductions.
However, achieving a net power reduction through such spanwise oscillations is
difficult (Quadrio and Ricco, 2003), hence optimisation of the actuation method is
necessary. Travelling wave oscillation is currently one the most efficient methods
of introducing oscillations into the boundary layer. Work by Choi et al. (2002) and
Quadrio and Ricco (2003) indicates that there are optimal combinations of oscillation
frequency, depth of influence, and amplitude of maximum oscillatory velocity. Both of
these studies indicate that oscillations broadly influence the instability mechanisms in
the boundary layer. These studies, combined with recent research on the stabilisation
of Taylor-Couette flows by axial oscillations (e.g. Marques and Lopez, 1997, 2000;
Meseguer and Marques, 2000) suggest a stabilisation mechanism inherent to near-
wall oscillations. The author is not aware of previous published research on the linear
stability of Blasius boundary layers subject to oscillatory forcing.

2
1.1. THE BOUNDARY LAYER 3

1.1 The boundary layer


The boundary layer is the area of fluid close to the wall in external flows. External
flows have extents far greater than the surface in question, and contrast to the finite
domain on internal flows, such as pipe or channel flow. Flow in the boundary layer
is categorised as either laminar or turbulent. In laminar flow, the fluid moves in
parallel shear. However, turbulent flow involves seemingly chaotic disruption in the
form of eddies (areas of local vorticity), which break down into smaller and smaller
eddies in the so-called turbulent cascade before being completely attenuated due to
viscosity. The origin of these turbulent eddies is thought to be at the wall, in the
turbulent boundary layer. A similar process occurs in pipe and channel flows, but the
turbulence spreads throughout the flow because it is bounded by turbulent boundary
layers.
Laminar boundary layers are typically unstable at high speeds, which correspond
to a high Reynolds number. This instability causes small perturbations in the flow
to grow and eventually interact in non-linear fashion, which leads to transition to
turbulence through a variety of mechanisms. Once the boundary layer is turbulent,
several typical types of fluid motion are found in the turbulent boundary layer.
An area of fluid undergoing similar motion is termed a coherent structure. In
the turbulent boundary layer, there are areas of fluid close to the wall, aligned in
the streamwise direction, which move at higher or lower speed relative to the mean
flow. Such areas are called high-speed and low-speed streaks, respectively. Low speed
streaks can be seen as black areas of Figure 1.2, which shows a wall-normal view
of a typical turbulent boundary layer. These streaks are thought to be caused by
so-called quasi-streamwise vortices, or simply streamwise vortices, and appear as grey
features in Figure 1.2. Streamwise vortices occur near the wall and therefore undergo
convection and shearing with the mean flow. There can also exist small hairpin or
arch vortices which rise up away from the wall and into the higher regions of the
boundary layer.
In cases where the streamlines within the boundary layer break away from the
wall, the boundary layer is said to have separated. The most obvious illustration of
the effects of boundary layer separation is the stall of an aircraft wing, where the flow
no longer follows the top contour of the wing, dramatically reducing lift.

1.2 Problem definition


To better understand the analysis of following chapters, it is necessary to define the
nomenclature and terminology used. A basic understanding of fluid mechanics will
4 CHAPTER 1. INTRODUCTION

wall
oscillation

time

Lorentz
forcing

N S N S N S
Electromagnets

travelling
wave

Figure 1.1: Three methods of boundary layer actuation. Wall oscillation (top),
Lorentz forcing (center), and travelling wave oscillation (bottom) all induce an oscil-
latory flow near the wall.
1.2. PROBLEM DEFINITION 5

Figure 1.2: Typical visualisation of a turbulent boundary layer structure in plan.


Flow is from left to right. Black indicates a low-speed streak, and grey indicates a
streamwise vortex (from Schoppa and Hussain, 2002)

Wall-normal, y
v
w
Spanwise, z u x
a m w ise, ss
Stre ickne
er th
ar y lay
nd
Bou
Le
ad

ream
in

t
g

e - s
ed

Fre velocity
ge

Figure 1.3: Three-dimensional view of boundary layer showing streamwise, spanwise


and wall-normal components.
6 CHAPTER 1. INTRODUCTION

be assumed. Figure 1.3 shows the basic terminology and the coordinate system of
the Blasius boundary layer flow under consideration. The main flow U∞ is treated as
uniform far away from the plate, although in reality some free-stream turbulence may
exist. The plate causes the formation of the boundary layer through shear because of
the no-slip condition at the wall. The global coordinate system is defined by the X,
Y , and Z axes, which lie along the streamwise, wall-normal, and spanwise directions
respectively. A natural point to define as the streamwise origin, X = 0, is the point
where the flow meets the plate, termed the leading edge. However, in this work the
domain of interest is typically much smaller than the extent of the plate, and so a
local coordinate system of x, y, and z will be used. Often the streamwise boundary
layer thickness δ with X 1/2 will be neglected for calculations in the local coordinate
system. The direction normal to the wall will be denoted y, with y = 0 at the plate
surface. The direction parallel to the leading edge, the spanwise direction, is denoted
by z. The plate is considered to be infinite in spanwise extent and z = 0 is defined
as the center of the domain in consideration.

10
U
9 dU/dy
2 2
d U/dy
8
nondimensional height

0
−0.4 −0.2 0 0.2 0.4 0.6 0.8 1
U/U
inf

Figure 1.4: Blasius boundary layer profile used in stability calculations.

The flow considered is close to the wall, within the boundary layer, where the
local flow may be in any direction. This flow will be represented by U , V , and W
1.2. PROBLEM DEFINITION 7

components in the x, y, and z directions respectively. These components will be


linearly decomposed into steady components and small disturbances:

U (t) = U + u (t) V (t) = V + v (t)


(1.1)
W (t) = W + w (t) P (t) = P + p (t)

where U , V , and W are the base flows that would be present without disturbance
(Blasius flow), and u, v, and w are the disturbance velocities. Likewise, P and p are
steady and disturbance components of pressure, respectively. In the present work,
the spanwise steady flow component W will always be zero.
The constant part of a disturbance may be separated by calculating averages of
the disturbance over several periods of the lowest oscillatory Fourier component in
the flow. These mean disturbance velocities are denoted by u, v, and w, leaving the
fluctuating components u e, ve, and w.
e The fluctuating part can also be studied by
computing the root mean square of the disturbance, denoted urms , vrms , and wrms .
Vorticity in the three coordinate directions are defined as,

∂w ∂v
ηx = −
∂y ∂z
∂w ∂v
ηy = − (1.2)
∂z ∂x
∂w ∂v
ηz = −
∂x ∂y

These can be decomposed in the same way as the velocity components. A vortex
in the x direction is a swirling motion around an axis parallel to the x axis and is
associated with vorticity $x . It is important to note that vorticity itself does not
imply the presence of a vortex. Table 1 summarises the use of symbols and subscripts
in this document.
The Blasius boundary layer profile under consideration is defined by the similarity
solution given by Equation 1.3 (Blasius, 1908).

f 000 + f f 00 = 0 (1.3)

Where a prime (0 ) indicates differentiation of the streamfunction f with respect to


the similarity variable, y ∗ , given by:
8 CHAPTER 1. INTRODUCTION

v U
ëz mean
flow

ëz x

Figure 1.5: Diagram of an instability wave propagating at an angle to the mean flow.
The diagonal lines represent constant phase of the disturbance wave.

r
U∞
y∗ = y (1.4)
2νX
The mean velocity profile is then given as U (y) = U∞ f 0 (y ∗ ). Equation 1.3 can be
rearranged into a linear system of equations and readily solved numerically,

2U̇3 + U1 U3 = 0
U̇1 = U2 (1.5)
U̇2 = U3

1.3 Wavelike disturbances - a prelude to stability


theory
To understand the review of turbulent boundary layer phenomena in Part II, a brief
introduction to stability theory is required to define the terminology. A given steady
flow field may be disturbed such that the total velocity at a certain point is given by
a sum of the mean and disturbance values. For example, U (t) = U + u (t). In shear
flows these disturbances are often wave-like, which suggests a decomposition of the
total disturbance into a sum of waves.
Fourier decomposition allows a given arbitrary disturbance to be separated into
several component modes. Figure 1.5 displays a wall-normal view of a flow with a
wave propagating at an angle to the mean flow direction U , with the lines representing
1.3. WAVELIKE DISTURBANCES 9

positions of constant phase. Wavelengths λx and λz in the streamwise and spanwise


directions respectively can then be defined. A stationary observer will register a
temporal frequency ω. A velocity component, such as v, of a single wave disturbance
depending on all coordinates and time t, may now be represented as,

v (x, y, z, t) = v̂ (y) ei(kαx+lβz−mωt) (1.6)

where

2π 2π
α= , β= (1.7)
λx λz

are the streamwise and spanwise wavenumbers, respectively. Note that the frequency
ω is closely related to the streamwise wavenumber, α, by the speed with which the
disturbance travels downstream. α, β, and ω are chosen to represent the fundamental
wave disturbance. The integers (or integer fractions) k, l, and m, can then be used
to represent all other disturbances in the wave decomposition and relate them to the
primary one. Considering the flow at a specific time t, a disturbance will be denoted
as a mode (k; l), meaning a disturbance with streamwise wavenumber k · α and
spanwise wavenumber l · β. However, considering the flow at a specific downstream
position, x, the modes will be of the form (m; l), representing frequency m · ω and
spanwise wavenumber l · β. These two cases represent the same type of disturbance
in the flow. Some examples of the three major disturbance types are displayed in
Figure 1.6. Modes of the form (α,0) produce two-dimensional TS waves, (0, β) modes
form low-speed streaks, and modes of form (α, ±β) produce oblique waves.
10 CHAPTER 1. INTRODUCTION

(1,0) (0,1) (1,1)

2D Streaks Oblique

(2,0) (0,2) (1,1) + (1,-1)

Figure 1.6: Common modes of instability in the Blasius boundary layer.


Chapter 2

Flow control

In his ground-breaking presentation, Prandtl (1904) introduced boundary layer theory


and was the first to show the potential of flow control. He used suction to delay the
separation of the boundary layer over the surface of a cylinder, thus reducing the
drag on the cylinder. Following Prandtl’s discovery, flow control has been under
considerable research for the last century.
Flow control is large-scale modification of the flow to induce more favourable flow
dynamics. This interaction is usually most effective when implemented in the fluid
near the wall - hence the term boundary layer flow control. Boundary Layer Control
(BLC) finds application in many areas, due to the potentially large gains in efficiency
and performance for relatively low energy input. By exploiting natural flow responses
to small-scale disturbances, the entire flow-field may be dramatically modified. Flow
control technologies are therefore of immense technological importance. The potential
for flow control technologies in a variety of systems, such as aircraft drag reduction,
has been demonstrated to be both significant and dramatic.
With rapid advances in computational power, flow control technology is advancing
faster than ever before. The present work focuses on using boundary layer flow control
to stabilise the boundary layer, resulting in delayed transition, turbulence suppres-
sion, and drag reduction. In the last 20 years, considerable research has increased
understanding of near-wall dynamics has indicated the potential of flow control for
drag reduction. Despite the large accumulated body of knowledge, boundary layer
control remains the most challenging and promising areas of fluid mechanics.

2.1 Purpose
With rising fuel prices, there is a greater demand for fuel efficiency, particularly in the
shipping and commercial aircraft industries. Drag reductions of as little as 5% could
save millions in fuel costs per year. There is therefore considerable incentive to find

11
12 CHAPTER 2. FLOW CONTROL

Boundary layers Vortices Jets, wakes, and mixing


Separation control Forebody vortex control Mixing enhancement
Drag reduction Blade-vortex interaction Jet vectoring
Transition delay Wing-tip vortex dynamics Noise suppression
Virtual surface shaping Vortex generation and control Wake modification

Modification of flow phenomenon

Flow
control

Actuators Sensing and control


Structural element Passive
Fluidic (None)
Thermal Open-loop
Acousic Basic control system
Piezoelectric and sensing
Synthetic jets
Closed-loop
Electromagnetic
Shape memory alloys Control
MEMS Neural Networks
Adaptive
Physical-model based
Dynamical systems based
Optimal control
Sensing
Conventional
Optical
MEMS

Figure 2.1: Flow control objectives and methods (after Kral, 2001).

practical methods to reduce drag of aircraft and ships. Savings can be achieved by
either reducing form drag or drag due to skin friction. This work indicates that drag
reduction can be achieved through stabilisation of the boundary layer. Flow control
can also reduce aircraft noise by minimising near wall turbulence or modifying vortex
dynamics, and spin-off applications in manufacturing and biomedical industries are
not unlikely.
Although the technology to implement active boundary layer control systems is
still in its infancy, considerable progress has been made. Braslow (1999) reviews
some practical implementations and experiments of boundary layer control to several
aircraft. While cost-effective active boundary layer control technology is still several
years away, the theoretical and experimental understanding of control parameters can
be improved greatly with current research. The variety of control objectives is shown
in Figure 2.1, with several actuator types and control schemes also given.
Considerable knowledge has been accumulated, but recent advances in computa-
2.2. METHODS 13

tional and manufacturing technologies has resulted in renewed research into turbu-
lence dynamics and drag reduction techniques. Several methods of flow control exist,
which are briefly reviewed and discussed herein.

2.2 Methods
Drag reduction methods can be separated into three broad categories; separation de-
lay, transition delay, and turbulent drag reduction. Separation delay reduces form
drag by maintaining attachment of the boundary layer over a larger percentage of
the surface, thus recovering more pressure toward the rear. Delaying or preventing
transition can be even more effective, because the laminar boundary layer has con-
siderably less skin friction than the turbulent boundary layer, although a laminar
boundary layer is more liable to separate. Turbulent drag reduction can be used
where transition cannot be prevented to lower the skin friction toward laminar levels.
Typically, reduction of the turbulence level within the boundary layer is necessary to
reduce the turbulent skin friction.
Within these categories exist both passive and active control strategies. Passive
strategies require no energy input, while active methods require an energy source and
actuation. In closed-loop active control, information about the downstream flow is
used to modify the control at an upstream location, while such feedback does not
occur in open-loop systems. Itemising the control strategies:

• Separation delay

– Passive
∗ Streamwise vortex generators
∗ Boundary layer trips
– Open-loop
∗ Polymeric
∗ Suction
∗ Wall jets
∗ Synthetic jets

• Transition delay

– Passive
∗ Compliant coatings
– Closed-loop control
14 CHAPTER 2. FLOW CONTROL

∗ Wall suction or jets


∗ Surface deformation

• Turbulent drag reduction

– Passive
∗ Riblets
∗ LEBU’s
∗ Compliant coatings
∗ Streamwise vortex generators
∗ Randomised surface bumps
– Open-loop control
∗ Tangential jets
∗ Polymeric
∗ Microbubbles
∗ Wall heating or cooling
∗ Spanwise wall oscillations or forcing
– Closed-loop control
∗ MEMS devices
∗ Wall deformation
∗ Microbubbles
∗ Wall heating or cooling
∗ Spanwise wall oscillations or forcing

This thesis highlights the potential of open-loop forcing of the boundary layer as
a means of stabilising the boundary layer, thus reducing drag. Open-loop control
has potential to allow large drag reductions to be achieved without the need for
expensive and delicate sensors, which are required for closed-loop control systems.
To understand this work in the broader context of flow control for drag reduction, a
brief review of drag reduction strategies is presented here.

2.2.1 Suction
Wall suction (and blowing) is undoubtedly the most researched and currently, the
most effective method of flow control. Suction was used by Prandtl (1904) to demon-
strate the role of the boundary layer in his 1904 experiment. The effectiveness of
such techniques lies in modification of the momentum of the near-wall fluid. Much
2.2. METHODS 15

progress has been made on using spanwise suction slots, tangential wall jets, or dis-
crete suction (blowing) micro-holes to delay flow transition or separation (Gad-el Hak,
2000). The small amounts of fluid withdrawn from the near-wall region of the bound-
ary layer draw higher-momentum fluid towards the wall, which modifies the mean
velocity profile and thus has a stabilising effect on the boundary layer. Stabilisation
methods have been successfully used by MacCormack et al. (2002) among others to
delay boundary layer transition.
The friction drag in a laminar flow is considerably lower than the friction drag
in a turbulent flow. In laminar boundary layers, stabilisation can lead to delays
in turbulence transition, and thus result in a substantial drag reduction. On the
other hand, stabilisation of a turbulent boundary layer can inhibit flow separation
and therefore lead to dramatic net drag reductions in airfoil applications. However, it
should be noted that the increase in momentum transfer results in an increase in skin-
friction drag along the surface. Boundary layers thicken with streamwise distance,
finally resulting in greater likelihood of boundary layer separation or transition to
turbulence. Suction inhibits this growth of the boundary layer, inhibiting separation
and delaying transition.
High frequency, open-loop oscillatory forcing in the form of blowing and suction
was used successfully by Choi (2002a) to achieve significant drag reductions on bluff
bodies. The forcing, which increased momentum in the boundary layer, was found
to relaminarise the flow in some cases and reduce separation in the majority of cases
studied. Work by Tardu (2001) using oscillatory suction and blowing through a
spanwise slot also achieved a large 40% drag reduction. It was concluded that this
actuation displaced the spanwise vorticity and quasi-streamwise vortical structures
away from the wall, reducing turbulent drag.
A promising open-loop control strategy similar to that of Tardu (2001) involves
large-scale, colliding, z-directed wall jets, which have may reduce drag by up to 50%
according to DNS studies by Schoppa and Hussain (1998). For a realistic application,
where flow length scales are typically on the order of 0.1mm for aircraft, such large-
scale forcing has many advantages. The need for sensing is eliminated and viable
actuation methods are available. However, this work did not address net power
saving or loss.
Active control methods are also used to increase the effectiveness and efficiency
of boundary layer control. Baker et al. (2002) concluded from numerical simulations
that closed-loop control methods achieved significant drag reductions through a small
number of sensors and actuators. The study also highlighted the importance of ac-
tuator and sensor locations for effective control of the flow. Drag reductions on the
order of 20% have been computed numerically by blowing and suction using feedback
16 CHAPTER 2. FLOW CONTROL

control (Chang et al., 2002). However, at higher Reynolds numbers (Reτ > 720) the
energy required for forcing makes control less feasible.
Wright and Nelson (2001) studied the energy cost for suction experimentally, by
delaying the transition over an airfoil. They found effective optimisation techniques
significantly reduce the actuation cost. Even using such optimisations and closed-loop
control methods, a cost-effective actuation suction method for airfoil or ship applica-
tions has not yet be found, due to technological issues relating to cost, maintenance
and reliability issues (Gad-el Hak, 1996b).
Active methods have also been used to stabilise boundary layer instabilities.
Högberg and Henningson (2002) used feedback-controlled suction and blowing to
successfully stabilise instabilities in a Blasius and Falkner-Skan-Cooke boundary lay-
ers. While effective, such closed-loop control methods are limited by the observability
of the flow components and sensor location. Lee et al. (1997) used a neural network
to control surface suction and blowing based on the measured spanwise wall-stress.
Their numerical simulation showed drag reductions of up to 20% were possible.
The challenge and current limitation of suction technology lies in a feasible method
of actuation. Many simulations have assumed an idealised distributed suction, when
in real applications it is impractical to achieve controlled distributed suction. Porous
media are also liable to blockage, and typically lack desirable surface characteris-
tics. Thus, discrete suction through small holes must be investigated. However, such
discrete suction results in the introduction of three-dimensional structures and lon-
gitudinal vortices to the boundary layer, which tend to destabilise the flow and can
result in transition at super-critical suction rates (MacManus and Eaton, 2000). Nev-
ertheless, considerable ongoing research and technological advances could result in a
commercial implementation of suction-based laminar flow control (Braslow, 1999).

2.2.2 Compliant coatings


Compliant coatings show great promise for drag reduction, by delaying transition
and potentially reducing near-wall turbulence levels. Compliant coatings are usually
a soft, thin layers which lie between the fluid and rigid wall of the channel or pipe.
The main advantage of compliant coatings is that they are a passive means of flow
control, and do not require constant energy input. Furthermore, they are typically
simple in structure and would therefore be relatively easy to mass-produce. A recent
review of compliant technology is given by Gad-el Hak (2000).
Many studies have focused on the potential of compliant coatings in laminar flow
for delaying transition to turbulence. In a wind tunnel experiment by Lee et al.
(1995), it was found that a silicone-based compliant coating had a significant stabil-
2.2. METHODS 17

ising effect on the boundary layer when applied to a surface. Reductions in near-wall
velocity fluctuations of up to 40% were found, suggesting the practical use of compli-
ant coatings for delaying transition. Experimentally, transitional Reynolds numbers
that exceed those on rigid-surface boundary layers by an order of magnitude can be
achieved (Gad-el Hak, 1996a). For example, Carpenter (1993) calculated an increase
in transition Reynolds number from 2.25 × 106 for a rigid wall to 12.62 × 106 for a
compliant surface. In marine applications, such an increase could create significant
drag reductions (Gad-el Hak, 2000). Further work by Carpenter (1998) suggested the
possibility of complete suppression of turbulence by compliant coatings for suitable
flows, where Tollmien-Schlichting instabilities are the primary mode of transition.
In turbulent flows, Satake and Kasagi (1996) found that vortex regeneration,
as well as the primary turbulence mechanisms should be considerably suppressed
by the selective damping of the near-wall spanwise velocity fluctuations caused by
compliance. In more recent work by Xu et al. (2003) the compliant wall was simulated
as a homogeneous spring-supported plate. In contrast to Satake and Kasagi’s results,
it was found that there was little change in the very long term behaviour of the friction
drag and little modification to the near-wall turbulent coherent structures. In fact,
the values of pertinent statistical quantities of the turbulence near the compliant walls
converged to those near a rigid wall. Their results concluded that the statistical effect
of the wall compliance on the turbulent channel flow was small. It seems, therefore,
that while compliance is likely to be a viable option for transition delay, its application
in turbulent boundary layers remains uncertain.

2.2.3 Fixed small-scale structures

In the mid 1980’s, several experiments were made using so-called Large-Eddy Break-
Up (LEBU) devices. These devices typically consist of small, thin airfoils placed
near the wall. Balakumar and Widnall (1986) and Sahlin et al (1986) investigated
these devices as a means of reducing vertical velocity fluctuations in the boundary
layer. Despite reporting local skin-friction reductions of up to 40%, Sahlin et al.
(1988) concluded it is unlikely LEBU devices can be used for substantial net drag
reduction at practical Reynolds numbers. Drag associated with the LEBU structures
outweighed the local reduction in skin friction. However, Rashidnia and Ebiana (1999)
experimentally tested a tandem arrangement of airfoils and reported 2-10% net drag
reductions. Nevertheless, the difficulty in manufacture and maintenance of the small
structural elements that make up a LEBU device suggest these devices will not be a
practical means of drag reduction.
18 CHAPTER 2. FLOW CONTROL

Figure 2.2: A Micro-Electro-Mechanical-System (MEMS) fluid control actuator (from


Koosh et al., 1999).

2.2.4 Micro-electro-mechanical technologies


In the last decade, microfabrication technology has become well developed. This ad-
vance in technology has enabled the experimental testing of Micro-Electro-Mechanical
Systems (MEMS) for boundary layer control. The potential of MEMS for drag reduc-
tion has been experimentally tested. Gad-el Hak (1996b) gives a brief review of early
MEMS implementations for active control of near-wall turbulent structures. Other
reviews include those of McMichael (1996), and Ho and Tai (1996, 1998). A wide
variety of actuator types exist, one of which is shown in Figure 2.2. Although the
possibilities of MEMS-based active control are interesting, it is unlikely that such
systems will soon be practical for many applications. For instance, on aircraft wings,
insect-strikes, ice-formation, and corrosion would make current MEMS-based drag
reduction system unlikely to be practical or cost-effective.

2.2.5 Micro-bubbles
The injection of micro-bubbles into the turbulent boundary layer has great promise
for hydrodynamic applications. Early work on drag reduction using microbubbles was
done by Madavan et al. (1984), who reported very significant skin friction reductions of
greater than 80%. Micro-bubbles were created near the wall using a sintered stainless
steel plate with a nominal pore size of 5 µm. It was suggested the mechanism of skin
friction reduction was similar to that of polymeric interaction, where the sublayer
thickness increases causing a reduction in drag. The drag reductions reported by
Madavan et al. (1984) were confirmed by Deutsch and Castano (1986) and Clark III
2.2. METHODS 19

and Deutsch (1991) who each reported similar drag reductions. Unexpectedly, it
was found by Deutsch and Castano (1986) that the type of gas used to generate the
bubbles had a significant effect on drag reduction.
Using a three-metre test section, Kodama et al. (2000) found a 40% drag reduction
was possible. The mechanism of drag reduction caused by micro-bubbles remains
unconfirmed, because the complex two-phase nature of the flow makes simulation
difficult. It is known, however, that the average volume fraction, or void ratio, close
to the wall is important. Recent simulations by Sugiyama et al. (2002) failed to
determine the drag-reducing mechanism. Numerical simulations by Xu et al. (2002)
found that greater sustained drag reduction is possible with smaller bubbles, but did
not suggest any underlying mechanism. Gad-el Hak (2000) suggests that the bubbles
suppress the Reynolds stress production in the buffer zone that links the linear and
the logarithmic portions of the mean velocity profile, thus inhibiting turbulent mixing
and reducing drag.
Although large drag reductions are possible, considerable energy is required to
eject the gas into the boundary layer, significantly limiting the net energy savings
possible. Kodama et al. (2001) calculated that the reductions produced by microbub-
ble injection would have to become at least twice as efficient as current technology
before it would be practical for large ships. Currently, maximum net energy gains of
approximately 2% could be achieved, which is uneconomical in light of installation
and maintenance.

2.2.6 Polymer addition


The addition of polymers to fluid flows is now a mature and well-researched field, with
over 5000 papers published (Gad-el Hak, 2000). Skin friction reductions of up to 80%
are easily achieved with 100 parts per million or less of some polymers (Gad-el Hak,
2000), and drag reductions up to 85% have been reported (Hetsroni et al., 1997). This
technology has been implemented widely, including crude oil pipelines, fire-fighting
equipment, and high-speed water jet cutting (Gad-el Hak, 2000). Toms (1948) first
observed that the addition of a few parts per million of polymethyl methacrylate to
a turbulent pipe flow of monochlorobenzene reduced the pressure drop substantially
below that of the solvent alone at the same flow rate. This result sparked considerable
experimental research, which is reviewed by Lumley (1969).
The polymer additives are typically coiled polymer chains where each monomer is
joined to the preceding at a random angle. Notable proposed mechanisms of polymeric
drag reduction include those of Lumley (1977) and Landahl (1977). According to
Lumley’s theory, drag reduction is caused when the polymers lying in the wall region
20 CHAPTER 2. FLOW CONTROL

are elongated by the shear in the fluid, increasing turbulent length-scales and thereby
thickening the buffer layer. Landahl’s theory holds that the drag reduction is due to
the suppression of secondary instabilities in the near wall region.
Early numerical modelling of polymeric flows was done by Patterson et al. (1977),
but it was not until the 1990’s when increasing in computational power allowed direct
numerical simulations of polymeric solutions (e.g. Sureshkumar et al., 1997). Numer-
ical simulations by Hagiwara et al. (2000) confirm earlier theories with findings that
polymeric addition attenuates small-scale turbulent structure.
Although promising, the practicality of polymeric drag reduction limits its appli-
cation to contained liquid flows. Thus far, the high production cost of the polymers
combined with storage and application considerations have made this technology un-
viable for ships. However, decreasing manufacturing costs and increasing fuel prices
may see this technology increasingly used in the future (Gad-el Hak, 2000).

2.2.7 Riblets
Riblets are small, typically v-shaped grooves that run along the wall in the stream-
wise direction. Because riblets are a relatively inexpensive, passive means of drag
reduction, they have been used for a number of years. The first practical use of ri-
blet technology was on the U.S. men’s rowing boat at the 1984 Los Angeles Olympic
Games. They were also used successfully by Stars and Stripes to regain the America’s
Cup in 1987 (Karniadakis and Choi, 2003), and subsequently disallowed. Riblets pro-
vide up to 10% frictional drag decrease for specifically matched turbulent Reynolds
numbers (Benhalilou et al., 1994), but can cause large increases in drag (greater than
35%) for Reynolds numbers significantly away from design conditions (Han et al.,
2002). These large increase in drag are due to a larger wetted surface area.
It is thought that riblets restrain the spanwise movement of fluid (e.g. Chu and
Karniadakis, 1993), thereby stabilising some of the coherent structures in the bound-
ary layer and impeding the cascade of energy to small scales. For an optimised riblet
surface, the skin friction is 10% higher in the peaks and 40% lower in the valleys,
compared to the flat-walled case (Gad-el Hak, 2000). It is thought that the riblet-
like grooves on shark skin are a biological type of riblet (as shown in Figure 2.3),
as these have been shown to be optimally configured for drag reduction by Bechert
et al. (1985). A riblet film has been manufactured by 3M corporation, with potential
for use in commercial aircraft applications, which obtained 5-7% drag reduction in
experiments by Liu et al. (1990).
In fully developed laminar channel flow, Djenidi et al. (1994) found using exper-
imental laser doppler anemometry and numerical simulations that despite the large
2.2. METHODS 21

Figure 2.3: The riblet structure of shark skin (from Bechert and Bartenwerfer, 1989).

wetted area increase due to the riblets, the frictional drag was not measurably in-
creased. This finding is in contrast with work by Choi et al. (1991), who calculated
that riblets increased the net drag in such laminar flows. Such results support the
conclusion that riblets are only effective in turbulent boundary layers. However, man-
ufacturing, cost, and maintenance issues are still to be resolved before widespread
commercial use.
22 CHAPTER 2. FLOW CONTROL
Part II

Turbulence phenomena

23
Chapter 3

Transition to turbulence

Establishing the origins of turbulent flow and the mechanisms of transition from
laminar to turbulent flow remains an important challenge in fluid mechanics. These
phenomena are important in understanding the nature of turbulence. Although some
transition phenomena appear unique to the early onset of turbulence, similarities
exist between transition mechanisms and turbulence regeneration mechanisms. For
completeness, all basic instability and transition mechanisms for boundary layer flows
are briefly reviewed herein.
No mathematical model presently exists that can accurately predict the transition
Reynolds number on a flat plate. One obvious reason is the variety of influences such
as freestream turbulence, surface roughness, sound, etc. are not completely under-
stood, and are difficult to include mathematically. However, considerable theoretical
research has been made possible by the linear hydrodynamic stability theory initi-
ated by Rayleigh, Kelvin, and Helmholtz (von Helmholtz, 1868; Rayleigh, 1880, 1887;
Kelvin, 1910). Their work was furthered by Orr (1907); Sommerfeld (1908), who the
well-known Orr-Sommerfeld stability equation is named after.

3.1 Transition pathways


According to classical stability theory, early transition to turbulence begins with a
sequence of instabilities occurring on a succession of more and more complicated basic
flows, as originally proposed by Landau in 1944 (Bayly and Orszag, 1988). Typically,
the initial state is a plane, parallel shear flow such as plane Blasius flow. At sufficiently
high Reynolds number, long two-dimensional disturbance waves, known as Tollmien-
Schlichting (TS) waves (Tollmien, 1929; Schlichting, 1933), become unstable and grow
in magnitude exponentially. The Orr-Sommerfeld equation of classical stability theory
predicts the instability of these waves in parallel Blasius boundary layers, of which
so-called two-dimensional waves are usually initially observed.

24
3.1. TRANSITION PATHWAYS 25

Forcing Environmental Disturbances


Amplitude

Receptivity

Transient Growth
A E

B C D

Primary Modes Bypass

Secondary Mechanisms

Breakdown

Turbulence

Figure 3.1: Transition mechanisms in wall bounded flows (after Saric et al., 2002).

TS waves are essentially streamwise-periodic concentrations of spanwise vortic-


ity, which eventually equilibrate at some finite amplitude if undisturbed by further
perturbations. However, this TS flow state can be considered as a new periodic
base-state, and it is typically unstable to short-wave three-dimensional disturbances
- known as secondary instabilities. These secondary instabilities do not equilibrate at
finite amplitude, and the flow evolves directly into fully developed turbulence.
In spite of considerable progress, an overall theory remains rather incomplete with
regard to predicting the complete transition process (Saric et al., 2002). However,
basic mechanisms of transition for boundary layers in external flows are known. These
processes can be qualitatively assembled into ‘pahtways’ that lead to turbulence, as
proposed by Saric et al. (2002) in Figure 3.1. These transition mechanisms will be first
qualitatively overviewed before details of each stage are given in subsequent sections.
Velocity disturbances in the freestream, such as sound or vorticity, enter the
26 CHAPTER 3. TRANSITION TO TURBULENCE

boundary layer as fluctuations of the basic state, and induce instability the flow. The
susceptibility of the flow to such disturbances is called receptivity. Current research
is aimed at understanding the source of initial disturbances and controlling undesir-
able conditions. Larger amplitudes of freestream turbulence and larger roughness on
the surface lead to different receptivity modes being activated. Environmental dis-
turbances determine the initial conditions of disturbance amplitude, frequency, and
phase for the evolution of boundary layer instabilities. These initial disturbances are
therefore largely responsible for which mechanism of transition operates (e.g. Singer
et al., 1989). Research is currently seeking to quantify how freestream disturbances
are entrained into the boundary layer and create the initial amplitudes of unstable
waves. In Figure 3.1, the environmental disturbance amplitude increases schemati-
cally from left to right over all of the figure.
If weak velocity disturbances are generated and path A is followed, the initial
growth of these disturbances is described by linear stability theory of primary modes
(the linearised unsteady Navier-Stokes equations). This primary growth is weak, oc-
curs over a long streamwise length scale, and can be modulated by pressure gradients,
surface mass transfer, temperature gradients, and other effects. However, if stronger
disturbances are generated, the velocity disturbances may undergo transient growth.
This latter phenomenon has received much attention in recent years, and much
remains to be learned, but streamwise vorticity and wall-normal vorticity appear to
be important. Qualitatively, transient growth occurs when two nonorthogonal modes
of velocity perturbation interact and grow algebraically before decaying exponentially,
resulting in a comparatively rapid change in the basic flow state. This new state may
lead to primary (path B), secondary (path C), or bypass-type (path D) instabilities.
Strong disturbances may lead directly to bypass-type instabilities (path E), which
result in strong, nonlinear localised velocity perturbations which rapidly breakdown
and lead to turbulence. The bypass phenomenon is not well understood because linear
theories fail to describe the dynamics. Bypass transition has been well documented in
cases of roughness and high freestream turbulence, where transient growth is believed
to be significant (e.g. Reshotko, 2001).
However, if no bypass-type instabilities grow, secondary growth mechanisms dom-
inate the transition process. As the amplitude of primary instabilities and transient
growth mechanisms grow, effectively resulting in a new base flow state, interactions
occur in the form of secondary instabilities. If the stability of this flow is analysed, lin-
earising about this new base flow, these secondary instabilities can be approximated.
These secondary instabilities typically grow quickly over only 5 or 6 wavelengths of
the base-flow period, such as a TS wavelength, and rapidly lead to breakdown and
turbulence. Because primary instabilities grow at a considerably slower rate and thus
3.2. RECEPTIVITY PROCESSES 27

have a large streamwise extent compared to the secondary instabilities, an approxi-


mate location of transition is often estimated using only primary stability analyses.

3.2 Receptivity processes

For incompressible flows, there are many different paths through which disturbances
can be introduced into the boundary layer. The dynamics of the very early growth
of disturbance within the boundary layer is known as the boundary layer receptiv-
ity, and standard linear theory typically does not hold. Initially these disturbances
may be too small to measure, and they are observed only after they are magnified
through instability mechanisms. These initial disturbances include the interaction of
freestream sound or turbulence with leading-edge curvature, discontinuities in surface
curvature, or surface inhomogeneities. Moreover, the dynamics of three-dimensional
flows are different than those of two-dimensional flows.
A number of different instabilities can occur independently or concurrently. The
effect of a given free-stream disturbance depends on Reynolds number, wall curvature,
sweep, roughness, and initial conditions. Introduction of a boundary layer disturbance
is usually initiated by incoming freestream disturbance, which can be represented by a
Fourier series. According to receptivity theory, this freestream disturbance interacts
with an inhomogeneity of the body, causing its frequency spectrum to broaden to
include the response disturbance (Saric et al., 2002).
Small initial disturbances tend to excite the linear normal modes of the boundary
layer that are of the TS type (Saric et al., 2002), which corresponds to path A of
Figure 3.1. However, three-dimensional disturbances can lead to transient growth
and low-speed streak formation. Larger, localised, three-dimensional disturbances,
such as those caused by surface inhomogeneities, can also lead to bypass transition
(path E in Figure 3.1).
Recent work using a localised oscillating surface by Bake et al. (2002) showed
that the boundary layer is more receptive to three-dimensional disturbances than
two-dimensional ones. The receptivity amplitudes were found to increase with the
spanwise wavenumber, the wave propagation angle, and the disturbance frequency.
Early work by Hultgren and Gustavsson (1981) also pointed out the increased recep-
tivity of plane flows to spanwise disturbances. More information on receptivity can
be found in work by Goldstein and Hultgren (1989), King (2000), Luchini (2000), and
Bertolotti (2000), among others.
28 CHAPTER 3. TRANSITION TO TURBULENCE

3.3 Primary mode instabilities


Traditional research on laminar-turbulent transition has focused on the exponen-
tial growth of two-dimensional Tollmien-Schlichting (TS) waves (Berlin et al., 1994),
which is given by solutions of the Orr-Sommerfeld equation (OSE). In low-noise en-
vironments, transition to turbulence often occurs via the primary growth of these
waves, which are two-dimensional wavelike vertical perturbations of the parallel flow
field. The majority of research over recent decades has focused on instabilities due
to TS waves because the growth of these waves can be predicted accurately by lin-
ear stability analysis using the OSE (Berlin et al., 1994) or the Parabolic Stability
Equations (PSE) (Schmid and Henningson, 2001).
So-called two-dimensional TS waves result in a streamwise-periodic concentration
of disturbance velocity. These are given by a wavenumber (see Section 1.3) of the
form (α, 0). In Blasius boundary layers, these two-dimensional TS waves become
unstable at approximately Re = 520. TS-like waves can also evolve in the form
of spanwise periodic concentrations of disturbance velocity, which form low-speed
streaks. Streak-like TS modes are given by wavenumbers of the form (0, β). Oblique
TS modes (α, ±β) are also common primary instabilities, particularly in environments
of higher free-stream turbulence. These oblique modes usually are excited in pairs of
(α, β) and (α, −β) modes.
Finite-amplitude TS waves have been studied using perturbation-based stability
theory and by numerically solving the Navier-Stokes equations (Herbert, 1988). It
has been found that at low amplitudes, which are typically of concern in the growth
of these primary instabilities, linear perturbation theory provides accurate solutions
and nonlinear effects can be neglected (Schmid and Henningson, 2001).

3.4 Secondary instability mechanisms


Considerable numerical research in the last decade has clarified the so-called secondary
instability mechanisms. These secondary instabilities can be loosely divided into three
categories, TS-wave, streak, and oblique secondary instabilities. Considerable work
on secondary instabilities has been presented by Schmid and Henningson (2001), and
this section follows their work to explain these concepts.
Secondary instabilities can be divided into three main categories; TS wave, streak,
and oblique. Each of these mechanisms leads to turbulence, but the appearance of
the transition zone differs for each instability type. Using particle image velocimetry
(PIV) on artificially induced turbulence in a series of experiments, Berlin et al. (1999)
showed the different form of these instability types. Figure 3.2 shows examples of
3.4. SECONDARY INSTABILITY MECHANISMS 29

flow

Figure 3.2: Particle Image Velocimetry (PIV) visualisations of oblique (left), H-type
(center), and K-type (right) transition (from Berlin et al., 1999).

H-type, K-type and oblique transitions, which are different categories of secondary
instability.

3.4.1 Two-dimensional TS wave secondary instability


In low-noise environments, turbulent transition over a flat-plate usually involves a
predominantly two-dimensional linear stage, followed by a three-dimensional stage
in which nonlinear effects become significant. It is this three-dimensional stage in
which secondary instability mechanisms dominate. The three-dimensionality of the
disturbance growth then quickly leads to turbulence, and therefore it is important to
understand these mechanisms.
Secondary instability can be understood as the instability of a base flow, where
the base flow is not parallel but disturbed in the form of a primary instability mode.
Thus, the governing equations become more complex because the periodic modu-
lation of the base flow must be considered. Floquet theory is usually used when
dealing with secondary instability (Schmid and Henningson, 2001). The initial de-
velopment of secondary instabilities is often due to a resonant interaction between
a two-dimensional TS wave and a pair of oblique instability waves, with equal but
opposite spanwise wavenumbers. As this resonance develops the oblique waves begin
to interact and nonlinear interaction between the oblique waves dominates, quickly
leading to turbulence.
This resonant interaction on the two-dimensional base state by oblique waves has
been found to develop into one, of two, three-dimensional states. One is known as
30 CHAPTER 3. TRANSITION TO TURBULENCE

Figure 3.3: Fundamental (Klebanoff) secondary instability of TS waves (photograph


by W. S. Saric, from Herbert, 1988).

fundamental secondary instability and the other is called the subharmonic secondary
instability. Herbert (1983) found that these two three-dimensional states were caused
by secondary instabilities of the two-dimensional state, which begin to dominate as
the TS wave amplitude exceeds 1% of the freestream velocity (Herbert, 1988). Funda-
mental secondary instability, also known as K-type, was first recorded experimentally
by Klebanoff et al. (1962) and is shown in Figure 3.3. Subharmonic secondary insta-
bility is similar, but results in a different pattern of so-called peak-valley splitting, as
seen in Figure 3.4. The major difference between these modes of instability is in the
layout of the secondary vortices.

It is believed that if high enough levels of free-stream turbulence exist, turbulent


flow structures such as streamwise vortices and streaks can be caused by secondary
instability mechanisms (Wu and Luo, 2003) which are also characteristic of turbu-
lent flow structures. Berlin demonstrated that streamwise vortices co-exist and grow
with the secondary structures (Berlin et al., 1999). The relationship between these
structures is shown in Figure 3.5 (Berlin et al., 1999)

Many direct numerical simulations have been used to study TS wave secondary
instability. Results of a simulation by Schmid and Henningson (2001) are shown in
Figure 3.6. The basic TS wave can be seen in the top view of Figure 3.6 at t = 0,
and a subharmonic disturbance can be seen growing in time. Prior to full transition
to turbulence, distinctive Λ-vortices can be seen, which correspond to the Λ-vortices
in Figure 3.4.
3.4. SECONDARY INSTABILITY MECHANISMS 31

Figure 3.4: Subharmonic secondary instability of TS waves (photograph by J. T.


Kegelman, from Herbert, 1988).

(a) (b)

W W
FLO FLO

Figure 3.5: Relationship between streaks and secondary instability modes (from
Berlin et al., 1999).
32 CHAPTER 3. TRANSITION TO TURBULENCE

Figure 3.6: Contours of streamwise disturbance velocity at y = 1.5 for TS wave


secondary instability. (a) t = 0 (b) t = 500 (c) t = 1400. Dashed contours represent
negative disturbance velocity. (from Schmid and Henningson, 2001)
3.4. SECONDARY INSTABILITY MECHANISMS 33

3.4.2 Streak secondary instability


Streaks can form in the laminar boundary layer through transient growth mechanisms
or wave interactions. As similar streak-like structures appear in turbulent boundary
layers, the dynamics of streak instability is important to understanding both transi-
tion and regeneration phenomena. Once the disturbance velocity of a streak exceeds a
critical value, fundamental and subharmonic secondary instabilities begin to develop.
These instabilities are manifested by a streamwise modulation of the streak. Each of
these fundamental and subharmonic modes can occur as a sinuous or a varicose mode,
which correspond to velocity resonances in the spanwise and wall-normal directions
respectively. The differences between the sinuous and varicose mode is seen clearly in
the DNS results of Li and Malik (1995), shown in Figure 3.7, and the experimental
visualisations of Asai et al. (2002) in Figure 3.8.
Waleffe (1995, 1997) and Waleffe and Kim (1997) numerically simulated the linear
stability of streaks in a plane Couette flow at a low Reynolds number. Assuming a
simple two-dimensional base flow they demonstrated that streaks can develop sinu-
ous instabilities. These sinuous modes have often been observed experimentally and
numerically, and initially grow via a linear secondary instability mechanism. Waleffe
and Kim stated that the instability originates from spanwise inflection points in the
streamwise velocity of streaky flows. Reddy et al. (1998) investigated the same insta-
bility systematically in plane Poiseuille flow as well as in plane Couette flow to study
subcritical transition. Varicose instability of streaks has also been demonstrated in
several computational and experimental studies.

3.4.3 Oblique secondary instability


Oblique transition is a scenario initiated by two oblique waves with opposite wave
angle and in which non-modal growth plays an important role. The potential for this
transition method to dominate was shown by Schmid and Henningson (1992), who
calculated oblique transition in channel flow using a temporal DNS code. Reddy et al.
(1998) computed the energy needed to initiate oblique transition in channel flow, and
found that oblique transition required less energy than secondary instabilities of TS
waves. The same result holds true for Blasius flows, where the energy required for
oblique mode disturbance is almost an order of magnitude less than the TS mode for
a similar transition time (Schmid and Henningson, 2001).
Oblique secondary instability arises from the vortices generated by the oblique
waves, which occur without a TS wave present. A pair of oblique waves interact
in a nonlinear fashion to produce strong streamwise vortices, which in turn produce
streamwise streaks (Berlin et al., 1994). The streaks then become unstable as de-
34 CHAPTER 3. TRANSITION TO TURBULENCE

Figure 3.7: Contours of streamwise disturbance velocity for sinuous (above) and
varicose (below) modes of instability of streaks induced by Görtler vortices (from Li
and Malik, 1995).

scribed in 3.4.2. Figure 3.9 shows the initial streamwise disturbance induced by the
oblique waves (a), which grow streamwise streak-structures in (b) that subsequently
develop sinuous secondary instability (c).

3.5 Bypass mechanisms

It is generally accepted that bypass refers to a transition process whose initial growth
is not described by the primary modes of the Orr-Sommerfeld equation (OSE). These
instabilities are not described by the eigenvalues of the OSE, and can often lead to
significant transient growth and generation of streamwise streaks. Modal transition
can also be bypassed in environments of high free-stream turbulence or surface in-
homogeneities, which can lead to the formation of turbulent spots. Turbulent spots
are the typical transitional objects found in boundary layer flows at high Reynolds
numbers (Bayly and Orszag, 1988). A visualisation of a turbulent spot in a wind
tunnel experiment by Cantwell et al. (1978) is shown in Figure 3.10. Turbulent spots
can decay and the flow relaminarise, as Figure 3.10 shows, or can lead directly to
regenerating turbulence throughout the boundary layer.
3.5. BYPASS MECHANISMS 35

(a)

(b)

(c)

50 100 150 200


x-x0 (mm)

(a)

(b)

(c)

50 100 150 200


x-x0 (mm)

Figure 3.8: Sinuous (top) and varicose (bottom) secondary instability of a synthetic
low-speed streak (from Asai et al., 2002). Photographs (a), (b) and (c) of the varicose
mode are in streamwise succession.
36 CHAPTER 3. TRANSITION TO TURBULENCE

Figure 3.9: Contours of streamwise disturbance velocity for oblique secondary in-
stability at y = 2. (a) t = 250 (b) t = 700. Dashed contours represent negative
disturbance velocity. (from Schmid and Henningson, 2001)
3.5. BYPASS MECHANISMS 37

Figure 3.10: Flow visualisation showing a typical turbulent spot in a zero-pressure


gradient boundary layer (from Cantwell et al., 1978). Flow is from left to right.

Figure 3.11: Illustration of transient growth due to the nonorthogonal superposition


of two vectors that decay at different rates (from Schmid and Henningson, 2001).
38 CHAPTER 3. TRANSITION TO TURBULENCE

3.6 Transient growth


Transient growth is a nonlinear mechanism due to the nonorthogonality of the Orr-
Sommerfeld operator. This mechanism was first elucidated by Landahl (1980) and
then by Hultgren and Gustavsson (1981), who demonstrated transient growth could
lead to streak formation. The idea was used by Henningson et al. (1993) and others.
Recent reviews appear in Andersson et al. (1999), Reshotko (2001), and Schmid and
Henningson (2001).
Even though individual modes may be undergoing exponential decay, the nonorthog-
onality of the modes gives rise to an algebraic growth in disturbance, which is illus-
trated by the arrow in Figure 3.11. Thus, transient growth initially causes an alge-
braic (and sometimes rapid) instability growth, followed by an exponential decay as
the nonorthogonal effects subside. Studies have shown that large growth amplitudes
can be achieved through transient growth when the boundary layer is provided with
appropriate initial conditions. Thus, the spectrum of initial conditions depends on
receptivity processes. Finally, as shown in Figure 3.1, transient growth can lead to
secondary instabilities, direct distortion of the basic state that leads to secondary or
subcritical instabilities (path C), or direct bypass (path D) to turbulence.
Chapter 4

Coherent structures in the


turbulent boundary layer

The basic understanding of turbulent boundary layer dynamics was developed in


the 1940s, when rudimentary hot-wire sampling at the edge of turbulent flows first
showed evidence of organised motions in turbulent boundary layers (Corrsin, 1943;
Corrsin and Kistler, 1954). The discovery of coherent boundary motions prompted
considerable research in order to quantify the boundary layer dynamics, and flow
visualisation was the dominant experimental technique. Extensive research was con-
ducted that supported the basic model proposed by Theodorsen (1952). Runstadler
et al. (1963) made early visualisation experiments of violent near-wall ejections, and
was the first called the process ‘bursting’. Using hydrogen bubble visualisation tech-
niques, Kline et al. (1967) detailed the basic temporal and special evolution of the low
speed streaks and consequent bursts. Their findings were supported with the studies
done by Corrino and Brodkey (1969). Later work by Kim et al. (1971) identified
three stages of streak evolution, streak growth, oscillation, and breakup. Offen and
Kline (1974, 1975) suggested these were all due to the same underlying mechanism.

More recently, many numerical simulations and Particle Image Velocimetry (PIV)
experiments of turbulent boundary flows have been performed. These studies have
given considerable insight into the near wall structure of the turbulent boundary layer.
However, insufficient computational power limited the extent of Direct Numerical
Simulations (DNS) until the last decade, when higher computational power became
more readily available. Although the typical structure of the turbulent boundary
layer is now generally agreed upon, the underlying dynamics remain controversial
(Karniadakis and Choi, 2003).

39
40 CHAPTER 4. COHERENT STRUCTURES

Figure 4.1: Formation of a low-speed streak by streamwise vortices (from Blackwelder


and Eckelmann, 1979).
4.1. STREAMWISE VORTICES 41

4.1 Streamwise vortices

The dominant role of streamwise vortices near the wall in turbulence production and
drag generation is now widely accepted (e.g. Kim et al., 1987; Robinson, 1991; Pan-
ton, 1997). Blackwelder and Eckelmann (1979) initially proposed the accepted ‘lift-up’
mechanism of low-speed streak formation via counter-rotating quasi-streamwise vor-
tices, as shown in Figure 4.1. In this mechanism, the vortices lift low-speed fluid from
very near the wall up into the near-wall region. The lifted fluid is of lower streamwise
velocity and thus forms a so-called low-speed streak.
Computer simulation of turbulent flows was pioneered in the work of Murlis et al.
(1982), Blackwelder and Haritonidis (1983), Alfredsson and Johansson (1984), and
Kim and Spalart (1987). The numerical simulation by Kim et al. (1987) confirmed
that streamwise vortices are the most common vortical structures in the near wall
region. Further simulations by Jiménez and Moin (1991) and Jeong et al. (1997)
developed these results further. Although the outer region of the boundary layer
contains large, energetic structures (Adrian et al., 2000), numerical experiments by
Jiménez and Pinelli (1999) confirmed the suggestion by Kline et al. (1967) and Kim
and Spalart (1987) that the essential inner-layer dynamics (namely y + < 60) can
operate autonomously. It is these inner-layer dynamics that are key to understanding
and controlling the turbulent boundary layer.
Direct evidence linking the organized vortical structure of turbulent boundary lay-
ers and measurable stress quantities in the flow has been provided in many numerical
simulation studies (e.g. Bernard et al., 1993; Kravchenko et al., 1993; Jiménez and
Pinelli, 1999). Analysis of numerical data for turbulent flow in a channel demonstrated
that Reynolds stress production (and therefore surface friction) is linked directly to
the dynamics of quasi-streamwise vortical structures in the wall region.
In an analysis of DNS data by Kravchenko et al. (1993), streamwise vortices were
generally observed to lie directly above and displaced laterally from high-skin-friction
regions. The net drag associated with regions of strong streamwise vorticity was
larger than the mean skin friction for the entire wall. These results can be under-
stood by considering the high-momentum fluid away from the wall being advected
toward the wall by the vortices. In Jiménez and Pinelli (1999), it was shown that
artificially interrupting the streak generation cycle led to large drag reduction and
even relaminarisation of the flow.
Most research, both simulated and experimental, has been performed using chan-
nel or pipe flow at relatively low Reynolds numbers. Due to experimental and com-
putational limitations, the Reynolds numbers investigated have been orders of magni-
tude lower that typical real-world flows. Naturally, the results found at low Reynolds
42 CHAPTER 4. COHERENT STRUCTURES

(a)
y
x
90
+ =2
Lx
z
L+ =
z 100
100

80
(d )
(b) (c) 60
y+
40
h20
20
y

z 0 15
U +0

Figure 4.2: Typical streak structure (after Schoppa and Hussain, 2002). (a) isosurface
of constant streamwise disturbance velocity; (b) streak profile, where the thick contour
is common to both (a) and (b); (c) typical idealised streak distribution.

number cannot necessarily be applied to high Reynolds number flows. However, recent
work by Schoppa and Hussain (2002) suggests that the near wall dynamic mechanisms
analysed at relatively low Reynolds numbers remain similar into the high Reynolds
number regime.

4.2 Boundary layer streaks

Turbulent boundary layer streaks are significantly less well-ordered than laminar
streaks, and are inherently three-dimensional. The near-wall streaky structure of
developed turbulent boundary layers was first visualised by Kline et al. (1967). Their
results showed that turbulent low-speed streaks have a characteristic spacing which
has been confirmed by many later studies.
Streamwise vortices are thought to lift low-momentum fluid away from the wall,
and thereby initiate streaks. These vortices have a short lifespan and are convected
by the mean flow, giving the resulting turbulent streaks a wavy appearance, as shown
in Figures 1.2 and 4.2.
4.3. HAIRPIN VORTICES 43

Figure 4.3: Hairpin vortex structure originally proposed by Theodorsen (1952)


.

4.3 Hairpin vortices


Theodorsen (1952) was the first to propose the now-accepted concept of the hairpin
vortex. After analysing the vorticity-transport form of the Navier-Stokes equations,
Theodorsen suggested these structures to explaining turbulence production and dis-
sipation. Figure 4.3 shows the general structure of the hairpin vortices proposed by
Theodorsen.
Some years later, large eddy motions in the outer regions of the boundary layer
were discovered (Townsend, 1956; Favre et al., 1957; Grant, 1958), providing evidence
of coherent vortical structures in the boundary layer. Further research provided
supporting evidence of coherent structures near the wall in the form of streaks and
violent ejections (Kline and Runstadler, 1959; Kline et al., 1967). These streaks were
regions of low-speed fluid, relative to the mean flow, and were suggested to give rise
to the vortices observed in previous studies.
Further research gave insight into the interaction of the hairpin vortices with the
near wall region (Head and Bandyopadhyay, 1981). At low Reynolds numbers, ‘arch’
vortices are typically formed, while at higher Reynolds numbers the vortices were
shaped like hairpins, as shown in Figure 4.4. It was suggested that the legs of the
hairpin vortices form the quasi-streamwise vortices that lift up low speed fluid behind
44 CHAPTER 4. COHERENT STRUCTURES

“Hairpin”

“Arch”
or “Horseshoe”
HEAD

NECK
mber
n u
lds
e yno
R
ng
FLOW
LEG
r e asi
Inc

Figure 4.4: Typical hairpin vortex structure with Reynolds number (from Robinson,
1991).

the hairpins. However, symmetrical arch or hairpin vortices are rare in most turbulent
boundary layers, and asymmetric hairpins predominate.
During early experiments, considerable importance was placed on the energetic
‘burst’ and ‘sweep’ events. It was thought that these events were of fundamental
importance in turbulence. However, Schoppa and Hussain (2002) comment that the
so-called bursting - used to describe the intermittent, energetic process perceived from
scalar markers in flow visualization or from stationary sensors - does not reflect any
particular event, but is primarily the consequence of passage of near-wall vortices.
In turbulent flows at relatively low Reynolds numbers, hairpin vortices occur in
streamwise-aligned packets that propagate upwards in the streamwise direction at a
mean angle of between 12 and 20 degrees to the wall (Head and Bandyopadhyay,
1981; Adrian et al., 2000). The hairpins in these packets are typically spaced several
hundred viscous length-scales apart in the streamwise direction.
Acarlar and Smith (1987) performed experiments to determine possible genera-
tion mechanisms for the hairpin vortices. Their experiments showed that a sufficiently
large streak could cause roll-up of the crest shear layer and thereby create hairpin
vortices, as shown in Figure 4.5. Once streaks reached a critical amplitude, there
were subject to a linear instability mechanism, which caused growing, oscillatory dis-
turbance at the crest of the streak. Nonlinear instability mechanisms finally resulted
in the shear layers rolling up into hairpin vortices. This generation mechanism of
4.3. HAIRPIN VORTICES 45

Linear instability, Hairpin


oscillation and growth Nonlinear vortices
Synthetic low-speed streak vortex roll-up

Streamwise
vortices

Figure 4.5: Varicose streak instability and breakdown, resulting in the formation of
hairpin vortices (from Acarlar and Smith, 1987).

hairpin vortices has now been confirmed by several studies (e.g. Asai et al., 2002)
Chapter 5

Turbulence regeneration

Once the boundary layer undergoes transition to turbulence, it tends to remain tur-
bulent because the turbulent structures undergo continual generation. The turbulent
structures present in turbulence are thought to promote the formation of new struc-
tures, thereby regenerating the turbulence in the boundary layer. This self-sustaining
mechanism of near-wall turbulence has received much attention recently, in both nu-
merical simulations and experimental testing (Jiménez and Moin, 1991; Hamilton
et al., 1995; Waleffe, 1997; Jiménez and Pinelli, 1999).
The evolutionary dynamics of near-wall coherent structures have until recently
been poorly understood, presenting a barrier to the development of effective drag
reduction strategies. While a considerable body of knowledge has been accumulated,
widely disparate mechanisms for vortex formation and coherent structure regeneration
in near-wall turbulence have been suggested. Two primary categories are:

(i) Parent-offspring scenarios

New vortices are generated by the direct action (induction) of existing


vortices.

(ii) Instability-based mechanisms

Local instability of a quasi-steady base flow generates new vortices, without


requiring the presence of parent vortices. Note that the base flow is made
unsteady by previous events - resulting in an indirect feedback
mechanism.

5.1 Parent-offspring scenarios


Two main parent-offspring scenarios have been suggested. The first of these relies on
dominance of hairpin vortices, and the second scenario suggests a mechanism based

46
5.1. PARENT-OFFSPRING SCENARIOS 47

on Kelvin-Helmholtz instability.

5.1.1 Hairpin vortex formation


A parent-offspring mechanism was proposed by Smith and Walker (1998). They
presented a mechanism in which parent hairpin vortices spawn offspring hairpins.
In this scenario, vortex formation is driven by unsteady separation near the wall
(Doligalski and Walker, 1984), in which the parent hairpin pumps low speed fluid
away from the wall near its head and legs to generate a low-speed streak. The
resulting inflectional shear flow then rolling up by Kelvin-Helmholtz instability and
giving birth to new hairpins, as shown in Figure 4.5. Smits and Delo (2002) present a
mechanism similar to that of Acarlar and Smith (1987) and Smith and Walker (1998),
whereby symmetric and asymmetric hairpin vortices generate streamwise streaks.
These streaks in turn become unstable and generate more hairpins. This process is
shown in Figure (ADD FIG).
In contrast, using direct numerical simulation (DNS) studies initialized with con-
ditionally averaged streamwise vortices, Zhou et al. (1999) found that a sufficiently
strong single hairpin can generate a packet of hairpin vortices, both upstream and
downstream of the parent hairpin. In this mechanism, the induction of the parent
vortex generates intense local shear layers of predominantly spanwise vorticity. Sub-
sequently, these shear layers roll up into arch vortices that link up with the existing
streamwise-oriented legs, and are stretched by the mean shear into offspring hairpin
vortices, detached from the primary hairpin vortex.

5.1.2 Vorticity sheet generation and roll-up


Other studies propose that the vorticity inherent around streaks generates streamwise
vortices directly. The vortex surface geometry in these idealized flows exhibiting
hairpin vortex formation is characterized by an x-localised, lifted low-speed streak.
These studies suggest that the vortex lines in the lifted region tilt forward, due to the
faster streamwise velocity of their crests. Subsequently, the vortex sheet consisting of
these tilted vortex lines will collapse due to sustained stretching by the mean shear,
generating a hairpin vortex with two well-defined legs. The legs of these hairpins are
then suggested to generate new low-speed streaks.
However, detailed studies by Schoppa and Hussain (2002) point out that for the
highly x-elongated streaks observed in typical flows, vortex lines coincide with local
velocity isocontours, and therefore experience no tilting and therefore no amplifi-
cation, as the velocity is constant along each vortex line. Furthermore, two-legged
hairpins are rare in near-wall turbulence (Robinson, 1991; Brooke and Hanratty, 1993;
48 CHAPTER 5. TURBULENCE REGENERATION

Jeong et al., 1997). Typically, streamwise vortices are more numerous and more in-
tense than symmetric hairpins. This evidence supports the proposal of Schoppa and
Hussain (2002) that streamwise vortices are generated via a different mechanism to
that of two-legged hairpins.

5.2 Instability-based regeneration


Unlike in parent-offspring regeneration scenarios, existing vortices may play a more
indirect role. Several instability mechanisms by which low-speed streaks and hairpins
are generated have been suggested.

5.2.1 Wave-shear instabilities


Some studies have suggested that streamwise vortices originate from locally curved
streamlines generated near the wall. For example, the condition for Taylor-Görtler
instability - sufficient concave curvature of near-wall streamlines - is locally satisfied
above y + = 50 (Brown and Thomas, 1977). Alternatively, Phillips et al. (1996)
consider a instability mechanism involving x-dependent perturbation growth in shear
flows with small-amplitude streamwise undulation. However, such models have been
criticised in light of experimental data and other studies, because they fail to account
for all observed turbulence phenomena (Schoppa and Hussain, 2002).

5.2.2 Streak instability


Several studies have targeted instability of lifted low-speed streaks as the dominant
agent of turbulence production, although the details of the instability mechanism
vary widely in the literature. This instability concept germinated from the early
flow visualization studies of low-speed fluid (Kline et al., 1967), which suggested
spatial, unstable oscillations of local shear layers at the crest of low-speed streaks.
These unstable oscillations are analogous to the Kelvin-Helmholtz instability of free-
shear layers. Once the oscillatory amplitude grows sufficiently, nonlinear mechanisms
breakdown the streak into smaller structures.
Two secondary modes of streak instability can be demonstrated using linear sta-
bility analysis. These modes, called varicose and sinuous modes, are now widely
accepted. The varicose modes correspond to the oscillatory behaviour described by
Acarlar and Smith (1987) in Figure 4.5 (page 43), and also correspond to the vertical
oscillations observed by Kline et al. (1967). Sinuous modes occur due to the same
instability mechanism, but different parameters of stability result in oscillation of
5.2. INSTABILITY-BASED REGENERATION 49

the streak in the spanwise direction. This oscillatory instability grows exponentially,
consequently results in breakdown.
In the case of varicose instability, an arch vortex is generated by the roll-up of
an internal shear layer atop a streak, which links up with the downstream end of a
nearby streamwise vortex due to the shear-induced collapse of the connecting vortex
lines. Because the vorticity layer is of finite extent, circulation pile-up causes vortex
roll-up at the tip of the internal shear layer. Note that because no perturbations are
required for internal shear layer roll-up, the arch formation is not strictly an instability
process. However, the roll-up is due to vorticity concentration by two-dimensional self-
advection, and is therefore analogous to Kelvin-Helmholtz instability. Schoppa and
Hussain (2002) suggest that two-dimensional instability mechanisms fail to account
for the inherent three-dimensionality of this roll-up mechanism.
Based on the evolution of instantaneous structures visualized via DNS, Robin-
son (1991) proposes streaks are responsible for turbulence regeneration. Robinson
suggests that lifted low-speed streaks, left behind by (faster convecting) streamwise
vortices, contain locally unstable shear on the streak crest which then gives rise to
new spanwise arch vortices (varicose instability). One leg of the arch is stretched into
a streamwise vortex, which in turn generates a new unstable streak in its wake via
the lift-up mechanism to close the cycle.
Görtler instability arises in laminar flow where concave wall curvature exists, and
gives rise to pairs of Görtler vortices which create low speed streaks by lifting low-
momentum fluid away from the wall. The streaks generated by Görtler vortices are
similar in form to the streaks found in turbulence. In studies of x-independent streak
distributions generated by Görtler vortices (Hall and Horseman, 1991; Yu and Liu,
1991), the growth rates of varicose modes are found to be relatively small (approxi-
mately one-half) compared to the dominant sinuous modes - even with exceedingly
strong streak-crest shear. This finding casts doubt on the dominance of the regen-
eration mechanism suggested by Robinson (1991). Furthermore, for x-independent
streak distributions with more realistic shear magnitudes, varicose modes are typ-
ically stable (Schoppa and Hussain, 2002). Finally, the varicose instability would
necessarily be accompanied by formation of two-legged hairpin vortices, which are
indeed rarely found near the wall.
An alternative streak instability mechanism is proposed by Swearingen and Black-
welder (1987). Based on experimental analysis of streak breakdown induced by
Görtler vortices, a dominant sinuous mode of instability was found. From smoke-
visualization and hot-wire probe measurements, they inferred that turbulence pro-
duction is caused by local, wake-like instability of the shear layers flanking low-speed
streaks. This concept is distinct from Kline et al. (1967) and Robinson (1991). Sub-
50 CHAPTER 5. TURBULENCE REGENERATION

sequent stability analysis by Yu and Liu (1991) revealed that Görtler streak distribu-
tions, representative of the experiments by Swearingen and Blackwelder (1987), are
indeed unstable to the predominant sinuous modes.
As further support of the regeneration model based on sinuous streak instability,
Hamilton et al. (1995) studied vortex regeneration using the ‘minimal flow unit’ con-
cept of Jiménez and Moin (1991) in plane Couette flow. Hamilton et al revealed a
surprisingly cyclic flow evolution and identified a three-step closed cycle: (i) streak
formation by streamwise vortices, (ii) streak ‘breakdown’ via (normal-mode) sinuous
instability, and (iii) ‘regeneration’ of streamwise vortices due to nonlinear interactions
in the post-breakdown flow. This cycle is in agreement with the early experiments
by Kline et al. (1967) and Offen and Kline (1975).

5.2.3 Streak transient growth


An alternative explanation of observations made by Robinson (1991) on normal-
mode streak instability was suggested by Kawahara et al. (1998), who preformed a
linear secondary instability analysis of streak base flow. They found some modes
of secondary streak instability could directly induce streamwise vortices, suggesting
a mechanism unrelated to normal-mode streak instability. Furthermore, Kawahara
et al. (2003) conclude from a recent study that advective lifting is not by itself enough
to destabilise a streak, which suggests a more dominant instability mechanism drives
regeneration.
Schoppa and Hussain (2002) proposed that transient growth of three-dimensional
streak instability is the driving factor of turbulent regeneration. In contrast to pre-
vious studies involving transient growth of a streakless base flow, they performed a
linear stability analysis of a streaky base flow and found a linear mechanism they
called Streak Transient Growth (STG). STG is capable of causing ten-fold linear
amplification of x-dependent streaks.
Linear stability analysis performed by Schoppa and Hussain (2002) on typical
streak structures in the boundary layer showed that approximately 20% of streaks in
the boundary layer exceed the threshold for linear sinuous instability. STG occurs on
the far more numerous smaller streaks, stable to normal mode growth. The mecha-
nism they propose leads to the generation of streamwise vortices through STG from
these smaller streaks. This mechanism does not directly lead to hairpin or arch vor-
tices, instead these structures are generated though varicose instability of the streaks
produced by STG.
An energetic perturbation of streamwise velocity results in formation of a small
x-independent streak. Initiated by disturbances typical in developed near-wall tur-
5.3. CONCLUSIONS 51

bulence, STG causes this perturbation to excite a nonlinear amplification mechanism


of streamwise vorticity, resulting in the rapid generation of streamwise vortices. The
underlying mechanism of the exponential transient amplification is due to a combi-
nation of x-dependent spanwise velocity perturbations with vertical shear due to the
mean flow. As STG reaches nonlinear amplitude, the sheet of streamwise vorticity
collapses due to the stretching induced by mean flow shear, creating streamwise vor-
tices. These streamwise vortices lift up low speed fluid from near the wall to generate
new small-amplitude streaks, completing the regeneration cycle.
Schoppa and Hussain (2002) find good correlation between the coherent structures
predicted by the STG mechanism and those observed in numerical and experimental
studies. They conclude that STG is the prominent, possibly dominant, turbulence
regeneration mechanism.

5.3 Conclusions
Transient growth of turbulent streaks accounts for many observed turbulence phe-
nomena (Schoppa and Hussain, 2002) and appears to be the most likely regeneration
mechanism. This instability-based mechanism is inherently three-dimensional, and
leads to the rapid growth of low speed streaks without the need for parent stream-
wise vortices. Instead, the transient linear instability of low amplitude streaks leads
directly to generation of streamwise vortices, which then create streak through the
lift-up mechanism. If these streaks reach critical amplitude, Kelvin-Helmholtz insta-
bility and subsequent vorticity roll-up may create hairpin or arch vortices.
Due to the linear (transient) nature of this instability, similarities with laminar
(transient) instability mechanisms and secondary instabilities can be seen. If stabil-
isation of the primary and secondary laminar instabilities can be demonstrated, it
is likely that similar stabilisation may also be possible for the STG mechanism in
turbulent boundary layers. It is shown herein that such stabilisation of laminar flow
may be possible using large-scale, near-wall oscillatory forcing.
52 CHAPTER 5. TURBULENCE REGENERATION
Part III

Oscillatory Stabilisation

53
Chapter 6

Review of oscillatory forcing

Little research has been done on oscillations in laminar boundary layers. Berlin (1998)
performed experimental testing on oscillatory forcing in an oblique transition scenario.
However, turbulent boundary layer control by means of oscillatory forcing, typically
spanwise, has been under research for many years. Spanwise forcing is generally much
easier to accomplish than streamwise forcing, because the mean velocity profile creates
difficulty in achieving streamwise forcing without significant vorticity and wall-normal
velocity introduction.
Research has led to numerous well documented experiments in which appropriate
three-dimensional disturbances introduced into the flow produce friction drag and
turbulence reductions. The spanwise force can be introduced either by spanwise wall
oscillation, electromagnetic Lorentz forcing, or a transverse travelling wave. Travelling
wave forcing results in a spatially periodic and time periodic oscillatory force of the
form:

 
y
−∆ 2π 2π
F1 = Ie sin z− t (6.1)
λz T

where I is the amplitude, λz is the wavelength in the span, ∆ is the wall-normal depth
of influence, T is the period, y is the wall-normal distance, z is spanwise position,
and t is time.
Turbulent flow control with spanwise-wall oscillation was originally carried out by
Jung et al. (1992), who demonstrated in their DNS study that the skin-friction drag
of the turbulent channel flow could be reduced by wall oscillation. They subjected a
turbulent channel flow to either to an oscillatory spanwise cross-flow or to spanwise
oscillatory motions of one of the channel walls. The oscillations gave rise to a 40%
reduction in the streamwise component of the Reynolds stress, with no significant
increase in the spanwise component, resulting in a weakened velocity gradient and

54
55

over 30% friction drag reduction. Similar drag reductions of up to 40% have been
found using Lorentz force actuation (Berger et al., 2000).
The friction drag and turbulence reduction predicted by Jung et al. (1992) were
experimentally confirmed by Laadhari et al. (1994). Laadhari et al conjectured that
the continuous shifting of the longitudinal vortices to different positions relative to
the wall velocity streaks weakens the intensity of the streaks, by injecting high-speed
fluid into low-speed streaks and low-speed fluid into high-speed regions. Further DNS
studies by Baron and Quadrio (1996) and experiments by Choi and Graham (1998)
confirmed the low speed streaks were indeed weakened and even stabilised by wall
oscillations.
Baron and Quadrio (1996) and Quadrio and Sibilla (2000) also demonstrated that
net energy savings are possible from this drag reduction technique when the wall-
oscillation velocity was less than a half of the free-stream velocity. Indeed, there was a
10% net energy saving by the spanwise-wall oscillation when the velocity was a quarter
of the free-stream velocity. The DNS studies were complemented by experimental
investigations by Laadhari et al. (1994), Choi and Clayton (2001), and Choi (2002b).
Choi et al. (2002) measured the streamwise variation of wall-shear stress over an
oscillating plate in the turbulent boundary layer, and showed that the skin-friction
coefficient is reduced by as much as 45% over the oscillating wall as compared to that
over the stationary wall.
Constant crossflow also affects the turbulent structures near the wall. Kiesow and
Plesniak (2003) found that the length of near-wall streaks was significantly reduced
by the introduction of constant spanwise crossflow in a flow visualisation experiment
using Particle Image Velocimetry (PIV). This modification was attributed to the
influence of the spanwise shear, where it was found that the shear increased the
number and strength of flow structures that interacted with the inner region of the
boundary layer. The resulting increase in momentum transfer results in a thickening
of this inner layer, which has also been documented in experiments using spanwise
oscillations (Choi, 2002b; Choi et al., 1998). In an similar experiment on the influence
of constant rotation of a pipe around its axis, Orlandi and Fatica (1997) found the
spanwise crossflow resulted in drag reduction. Their results showed that crossflow
lifted streamwise vortices away from the wall, resulting in a more stable boundary
layer where the number and intensity of ejection and sweep events was reduced.
Berlin (1998) successfully implemented spanwise oscillation control strategies for
oblique transition scenarios. Results showed that transition could be delayed signifi-
cantly by such oscillations. Furthermore, when oscillations were used in a transition
scenario initiated by a random disturbance, it was found that transition could be
prevented. These results suggest oscillations have a stabilising effect on boundary
56 CHAPTER 6. REVIEW OF OSCILLATORY FORCING

layer instability mechanisms.


The focus in many recent drag reduction techniques has been on controlling lo-
cally individual streamwise vortices based on sophisticated, but often rather complex,
closed-loop control strategies. Some recent research involving closed-loop control
strategies was discussed in Section 2.2.1. While effective, such approaches may be
inefficient or not even feasible in the high Reynolds number regime where length
scales are of the order of tens of microns. The guiding fundamental principle behind
such approaches has been the weakening of streamwise vortices and the simultaneous
enhancement of spanwise vorticity. This theme has been followed in many turbulence
control strategies that are comprehensively discussed by Gad-el Hak (2000). However,
the research performed on near-wall spanwise oscillations presents a promising large
scale forcing strategy, as detailed flow information is not required and micro-scale
actuation is unnecessary.

6.1 Mechanism of drag reduction


Choi (2002b) and Choi et al. (2002) related the mechanism of drag reduction by span-
wise wall oscillations to the spanwise vorticity generated by the periodic Stokes layer
over the oscillating wall. Dhanak and Si (1999) were able to demonstrate numer-
ically that the interaction between evolving, axially stretched, streamwise vortices
and a Stokes layer on the oscillating surface beneath them leads to reductions in
skin friction. Their work suggested that the reduction in turbulent skin-friction by
spanwise-wall oscillation is a result of the attenuation in the formation of stream-
wise streaks. The production of turbulence energy was significantly reduced by the
increased interaction between the quasi-streamwise vortices and the oscillating wall,
leading to a rapid viscous annihilation of the coherent structure.
For drag reduction to occur, the turbulence regeneration mechanism that is due
to low speed streaks and streamwise vortices must be inhibited. The cause of drag
reduction by spanwise oscillations is therefore the modification of the dynamics that
involve longitudinal vortices or steaks in the near-wall region. Streaks are typically
stabilised by spanwise near-wall oscillations, while travelling wave forcing results in
elimination of many streaks. In both cases, wall-normal vorticity has been reduced,
as evidenced in reduced spanwise velocity fluctuation. Thus, it is suggested that this
reduction in spanwise velocity fluctuation leads to a reduction of streamwise vorticity
generation through the STG mechanism.
It is possible that spanwise motions cause interference between the low-speed
streaks and the streamwise vortices, as the vortices are moved relative to the streaks,
resulting in a weakening of both. Furthermore, the oscillatory motion may directly
6.2. PREDICTION OF DRAG REDUCTION 57

50
Channel, Re = 100 (Choi et al, 2002)
Channel, Re = 200 (Choi et al, 2002)
40 Channel, Re = 400 (Choi et al, 2002)
Pipe, Re = 150 (Choi et al, 2002)
Pipe (Quadrio & Sibilla, 2000)
30 Pipe (Choi & Graham, 1998)
Channel (Baron & Quadrio, 1996)
1000 Vc2+50 Vc
Dr

20

10

0 0.05 0.1 0.15 0.2 0.25


Vc

Figure 6.1: Drag reduction correlation by Choi et al. (2002).

cause weakening of the streaks, through modification of the streak cross-sectional


profile. This modification of streak profile may directly interfere with the STG mech-
anism of streak instability growth.

6.2 Prediction of drag reduction


In the case of near-wall spanwise oscillations, it has been suggested that drag reduction
is a function of the peak nondimensional forcing velocity (Choi, 2002b). Choi et al.
(2002) performed detailed analysis of several studies on wall oscillations, and suggest
a complex equation for drag reduction at low Reynolds numbers, which is plotted in
Figure 6.1. The reader is referred to Choi et al. (2002) for details of the parameter
Vc .
Du et al. (2002) found a simple equation relating forcing amplitude, time period,
and wall-normal depth of influence for travelling wave excitation. This travelling wave
excitation was actuated by an array of electromagnetic tiles, they suggest that such
a force can be realised by active wall motion. Furthermore, a recent work by Vlachos
and Rediniotis (2003) aims to develop a shape-memory-alloy based smart skin, with
which they hope to achieve up to 50% drag reductions for Reynolds numbers above
those tested numerically.
Differences exist in the drag reduction due to wall oscillation and travelling wave
excitation. While wall oscillations reduce streak strength and occurrence, travelling
58 CHAPTER 6. REVIEW OF OSCILLATORY FORCING

waves eliminate most streaks, often resulting in very wide areas of lower-speed fluid,
rather than typical streaks. These differences suggest the two methods differ in the
mechanism of interference with the turbulence regeneration cycle, and consequently
different scaling equations may apply. It is believed travelling wave forcing will be
more effective at higher Reynolds numbers. The recent review by Karniadakis and
Choi (2003) on travelling wave excitation indicates some flexibility exists in the imple-
mentation of a travelling wave. For a given Reynolds number, the following equation
was generally found to hold true for a given drag reduction:

I × T+ × ∆ = C (6.2)

where I is the forcing amplitude, T + is the time period, ∆ is the wall-normal depth
of influence, and C is an arbitrary constant. For values of C = 1, drag reductions
of above 30% were consistently achievable for a variety of combinations of the other
variables. Work by Berger et al. (2000) supports the relationship between amplitude
and time period as suggested by Karniadakis and Choi (2003) with similar results.

6.3 Conclusions
A wealth of research indicates near-wall oscillations have potential to dramatically
alter boundary layer dynamics. Many studies have achieved significant drag reduction
as a result of spanwise oscillations. The studies mentioned have made significant
contribution to the understanding of oscillatory forcing on viscous boundary layer
flows.
However, the mechanism by which near-wall oscillation achieves this drag reduc-
tion has not be established. The present work suggests that near-wall oscillations
have a stabilising effect on the instability mechanisms which may govern turbulence
transition and regeneration. The time-dependent stability of Blasius flow subject to
near-wall oscillation is analysed herein, and provides insight into this possible mech-
anism. What follows is a linear stability analysis of laminar flows, with the objective
of understanding the possible stabilisation mechanism of oscillatory forcing.
Chapter 7

Linear stability analysis

7.1 Introduction
Early theoretical work on the stability of inviscid fluid flows was initiated by Rayleigh
(1880, 1887). Rayleigh demonstrated that there must be an inflection point in the
velocity profile of inviscid flows for instability. Fjørtoft (1950) qualified this criterion
later by demonstrating the inflection point must be a local maximum. The famous
viscous analyses of Orr (1907) and Sommerfeld (1908) extended the foundational
theory of Rayleigh and resulted in the well-known Orr-Sommerfeld equation. This
equation was complimented by the wall-normal vorticity equation derived by Squire
(1933). Tollmien (1929) and Schlichting (1933) were the first to discover instabilities
in viscous flows where inflection points do not occur, after whom TS instability is
named. These works form the basis of linear stability theory. Reed et al. (1996)
reviewed the linear-stability literature and discussed the importance of stability as it
relates to transition and aircraft skin-friction reduction.
The Orr-Sommerfeld equation is difficult to solve because it is very sensitive to
small errors (Henningson, 2004). The Orr-Sommerfeld differential operator must
be discretised to obtain solution, and high errors can be introduced through the
discretisation method. Methods include the method of compound matrices, iterative
shooting methods, and spectral methods. The benchmark paper by Orszag (1971)
showed the ‘exponential convergence’ property of spectral methods to obtain a highly
accurate solution of the Orr-Sommerfeld equation. Following this work, the solution
method for the linear stability equations shifted away from classical finite-difference-
shooting methods towards spectral techniques. A spectral method has been employed
in this work for a number of reasons.
Firstly, spectral methods exhibit exponential convergence. Thus, an accurate
solution is obtained using less than one third of the number of nodes required for finite
difference methods. Secondly, spectral methods efficiently provide high accuracy of

59
60 CHAPTER 7. LINEAR STABILITY ANALYSIS

derivatives, which is essential due to the extreme sensitivity of the Orr-Sommerfeld


equation. Finally, shooting methods require reasonably accurate initial estimates for
the starting values, or divergence may result. Thus, shooting methods may result in
modes being missed from the solution. In contrast, spectral methods yield all the
eigenvalues of both the discrete and infinite spectrum without the need for shooting
or iteration. However, due to matrices being full, rather than sparse as in finite-
differencing, the number of nodes is somewhat limited by computational power.
In the present work, a spectral method using Chebyshev polynomials has been
employed, which is typical of recent works (e.g. Bertolotti et al., 1992; Berlin et al.,
1994; Andersson et al., 1999; Brandt and Henningson, 2002; Brandt et al., 2003;
Schoppa and Hussain, 2002). To map the Blasius boundary layer to the Chebyshev
computational domain, the transformation used by Malik (1990) is employed.

7.2 Governing equations


The Navier-Stokes equations, in Cartesian tensor notation are defined:

∂Ui ∂Ui ∂p 1 2
= −Uj − + ∇ Ui (7.1a)
∂t ∂xj ∂xi Re
∂Ui
=0 (7.1b)
∂xi

where Ui represents the mean flow velocity, U , V , and W , in each coordinate direction,
x, y, and z respectively. The divergence operator, ∇, is defined by:

∂2 ∂2 ∂2
∇2 = + +
∂x2 ∂y ∗ 2 ∂z 2
(7.2)
∂4 ∂4 ∂4
∇4 = + +
∂x4 ∂y ∗ 4 ∂z 4

The nondimensional height, y ∗ , for the Blasius boundary layer is defined as the
p
similarity variable given by y ∗ = Reδ y ν/U∞ x. Equations 7.1 are supplemented
with boundary conditions. For solid walls, ui (xi , t) = 0 is often used. To derive
the nonlinear disturbance equations, two flows are considered. An undisturbed state
(Ui ,P ) and a disturbed state (Ui + ui ,P + p) are considered, both of which satisfy
the satisfy the Navier-Stokes equations (7.1). The in each coordinate direction is
represented by ui . By subtracting the equations for the undisturbed state from those
of the disturbed state, the nonlinear disturbance equations are easily derived.
7.2. GOVERNING EQUATIONS 61

∂u ∂ui ∂Ui ∂p 1 2 ∂ui


= −Uj − uj − + ∇ ui − u j (7.3a)
∂t ∂xj ∂xj ∂xi Re ∂xj
∂ui
=0 (7.3b)
∂xi

In typical linear stability analyses, a parallel two-dimensional flow is considered for


which Ui = U (y ∗ ) and V and W are zero. This assumption is usually made because
U is typically orders of magnitude larger than V and W . In this case, however W
will not be neglected. Thus, Equation 7.3 can be re-written in component form where
V = 0 as:

∂u ∂u ∂u ∂p 1 2
+U +W + vU 0 = − + ∇u (7.4a)
∂t ∂x ∂z ∂x Re
∂v ∂v ∂v ∂p 1 2
+U +W =− ∗ + ∇v (7.4b)
∂t ∂x ∂z ∂y Re
∂w ∂w ∂w ∂p 1 2
+U +W + vW 0 = − + ∇w (7.4c)
∂t ∂x ∂z ∂z Re

∂ui
Nonlinear terms involving uj ∂x j
are neglected because this is a linear analysis, and
therefore the disturbance uj is assumed to be infinitesimally small, making the non-
linear terms negligible. The prime (0 ) denotes a y ∗ -derivative. Taking the divergence
of these equations gives:

∂2u ∂2u ∂2u ∂v ∂2p 1 2 ∂u


+U 2 +W + U0 =− 2 + ∇ (7.5a)
∂t∂x ∂x ∂z∂x ∂x ∂x Re ∂x
∂2v ∂2v ∂2v ∂2p 1 2 ∂v

+ U ∗
+ W ∗
= − ∗ 2
+ ∇ (7.5b)
∂t∂y ∂x∂y ∂z∂y ∂y Re ∂y ∗
∂2w ∂2w ∂2w ∂v ∂2p 1 2 ∂w
+U + W 2 + W0 =− 2 + ∇ (7.5c)
∂t∂z ∂x∂z ∂z ∂z ∂z Re ∂z

Then adding all these equations together yields:


62 CHAPTER 7. LINEAR STABILITY ANALYSIS

     
∂ ∂u ∂v ∂w ∂ ∂u ∂v ∂w ∂ ∂u ∂v ∂w
+ ∗+ +U + + +W + +
∂t ∂x ∂y ∂z ∂x ∂x ∂y ∗ ∂z ∂z ∂x ∂y ∗ ∂z
2
 
0 ∂v 0 ∂v ∇ 1 2 ∂u ∂v ∂w
+U +W =− p+ ∇ + + (7.6)
∂x ∂z 2 Re ∂x ∂y ∗ ∂z

Making use of the continuity equation, many of the terms become zero and the equa-
tion simplifies.

∂v ∂v
∇2 p = −2U 0 − 2W 0 (7.7)
∂x ∂z

Differentiating Equation 7.7 with respect to y ∗ yields:

∂p ∂v ∂v
∇2 ∗
= −U 00 − W 00 (7.8)
∂y ∂x ∂z

Whereupon substitution into Equation 7.8 into Equation 7.5b yields an Equation 7.5b
yields an Orr-Sommerfeld-type

  
∂ ∂ ∂ 2 ∂00 00 ∂ 1 4
+U +W ∇ −U −W − ∇ v=0 (7.9)
∂t ∂x ∂z ∂x ∂z Re

However, another equation is required to fully describe the disturbance. Such an


equation can be derived by differentiating the first Navier-Stokes equation (7.4a) with
respect to z and the third Navier-Stokes equation (7.4c) with respect to z, giving the
result:

∂2u ∂2u ∂ 2 u ∂v 0 ∂2p 1 2 ∂u


+U +W 2 + U =− + ∇ (7.10a)
∂z∂t ∂z∂x ∂z ∂z ∂z∂x Re ∂z
∂2w ∂2w ∂2w ∂v 0 ∂2p 1 2 ∂w
+U 2 +W + W =− + ∇ (7.10b)
∂x∂t ∂x ∂x∂z ∂x ∂z∂x Re ∂x

Because the functions are continuous, the order of partial differentiation is arbitrary.
Thus, subtracting Equation 7.10b from Equation 7.10a yields:
7.2. GOVERNING EQUATIONS 63

∂2u ∂2w ∂2u ∂2w


 2
∂2w
    
∂ u
− +U − +W −
∂z∂t ∂x∂t ∂z∂x ∂x2 ∂z 2 ∂x∂z
 
∂v 0 ∂v 0 1 2 ∂u ∂w
+ U − W − ∇ − = 0 (7.11)
∂z ∂x Re ∂z ∂x

Then using the equation for wall-normal vorticity:

∂u ∂w
ηy = −
∂z ∂x

A Squire-type equation is derived.

 
∂ ∂ ∂ 1 2 ∂v ∂v 0
+U +W − ∇ ηy = − U 0 + W (7.12)
∂t ∂x ∂z Re ∂z ∂x

An Orr-Sommerfeld-Squire-type substitution for vertical disturbance velocity and


wall-normal disturbance vorticity can be made to find a find a solution to Equa-
tions 7.9 and 7.12, such that v and ηy are:

v (x, y ∗ , z, t) = ṽ (y ∗ ) ei(αx+βz−ωt) (7.13a)


ηy (x, y ∗ , z, t) = η̃y (y ∗ ) ei(αx+βz−ωt) (7.13b)

Thus, the governing equations can be re-written:

 
2 2
 00 00 1 2 2 2

(−iω + iαU + iβW ) D − k − iαU − iβW − D −k ṽ = 0 (7.14a)
Re
 
1 2 2
η̃y + iβU 0 ṽ − iαW 0 ṽ = 0 (7.14b)

(−iω + iαU + iβW ) − D −k
Re

where D represents differentiation with respect to y ∗ , and k 2 = α2 +β 2 . Equations 7.14


are subject to the boundary conditions ve = De v = ηey = 0 at solid walls and in the
free stream.
64 CHAPTER 7. LINEAR STABILITY ANALYSIS

7.3 Chebyshev spectral method

7.3.1 Chebyshev collocation and differentiation


The collocation method requires a set of discrete Gauss-Lobatto collocation points,
forming a set of points at which the stability equations must be simultaneously sat-
isfied. These Gauss-Lobatto points are defined:

πi
ξi = cos , i = 0, . . . , N (7.15)
N

where N is the number of collocation points used. The disturbance velocity, ṽ, is
defined at the discretised collocation points, and can be represented by a truncated
series of Chebyshev polynomials.

N
X
ṽ (ξi ) = v̂k Tk (ξi ) (7.16)
k=0

Disturbance in normal vorticity, η̃y , can be approximated in a similar fashion. The


disturbance velocity at each collocation point is therefore the sum of all Chebyshev
polynomials evaluated at that point, where each Chebyshev polynomial is weighted
by the corresponding spectral coefficient. The k th Chebyshev polynomial is defined:

Tk (ξ) = cos k cos−1 ξ



−1≤ξ ≤1 (7.17)

Note that the value of Tk (ξ) is limited to −1 ≤ Tk (ξ) ≤ 1. Derivatives can be


obtained by differentiating Equation 7.17. This is done by transforming the function,
ṽ (ξ), into a vector of spectral coefficients, v̂ (ξi ), using the Chebyshev transformation
matrix, T. This transformation matrix is defined:

 
ijπ
Tij = cos i, j = 0, . . . , N (7.18)
N

which makes Tv̂ (ξi ) = ṽ (ξ). Differentiation can then be performed accurately in
the spectral domain by operating on the vector of spectral coefficients, v̂ (ξi ). For
example, the second derivative is defined:
7.3. CHEBYSHEV SPECTRAL METHOD 65

N
∂ 2 ṽ (ξ) X
= D 2
(ξi ) Tṽ (ξi ) = v̂k Tk00 T (ξi ) (7.19)
∂ξ 2 k=0

The operator D is the strictly upper triangular Chebyshev differential matrix with
entries deduced from Equation 7.20 (Peyret, 2002):

N
2 X
vk0 = pvp k = 0, . . . , N − 1 (7.20)
ck p=k+1
(p+k) odd

where

2 if k = 0
ck = (7.21)
1 if k ≥ 1

For example, using Equation 7.20, the Chebyshev differential matrix for N = 7 is
written:

 
0 1 0 3 0 5 0 7
 
 0 4 0 8 0 12 0 
 

 0 6 0 10 0 14 
 
1
 0 8 0 12 0 
D (ξ) =   (7.22)

 0 10 0 14 
 

 0 12 0 


 0 14 

0

The boundary values of Chebyshev polynomials defined:

Tk (±1) = (±1)k
(7.23)
Tk0 (±1) = (±1)k+1 k 2

7.3.2 Transformations
To incorporate the semi-infinite physical domain of the Blasius boundary layer 0 ≤
y ∗ ≤ ymax to the Chebyshev interval −1 ≤ ξ ≤ 1 an algebraic mapping after that of
66 CHAPTER 7. LINEAR STABILITY ANALYSIS

Malik (1990) is used. Although other algebraic, logarithmic and exponential trans-
formations have been defined (Hanifi et al., 1996; Theofilis, 1994; Schmid and Hen-
ningson, 2001), this transformation gave the best results.

1+ξ yb − a
y∗ = a , ξ= (7.24)
b−ξ y∗ + a

where

yi ymax
a= (7.25a)
ymax − 2yi
2a
b=1+ (7.25b)
ymax

and ymax = 35, yi = 10. Using these values for a and b the derivatives over the
physical domain, y ∗ , are:

∂ξ ab + a
ξ 0 (y ∗ ) = = (7.26a)
∂y ∗ (y ∗ + a)2
∂2ξ −2 (ab + a)
ξ 00 (y ∗ ) = ∗ 2 = (7.26b)
∂y (y ∗ + a)3
∂3ξ 6 (ab + a)
ξ 000 (y ∗ ) = ∗ 3 = (7.26c)
∂y (y ∗ + a)4
∂4ξ −24 (ab + a)
ξ 0000 (y ∗ ) = ∗ 4 = (7.26d)
∂y (y ∗ + a)5

Thus, using the chain rule, the Chebyshev differentiation matrices can be defined:

T (y ∗ ) = T (ξ) (7.27a)
D (y ∗ ) = ξ 0 TD (ξ) (7.27b)
2
D2 (y ∗ ) = ξ 0 TD2 (ξ) + ξ 00 TD (ξ) (7.27c)
03
D3 (y ∗ ) = ξ TD3 (ξ) + 3ξ 0 ξ 00 TD2 (ξ) + ξ 000 TD (ξ) (7.27d)
04 0 2 00
D4 (y ∗ ) = ξ TD4 (ξ) + 6ξ ξ TD3 (ξ)
h i
00 2
+ 3ξ + 4ξ ξ TD2 (ξ) + ξ 0000 TD (ξ)
0 000
(7.27e)
7.4. TEMPORAL FORMULATION 67

where ξ 0 , ξ 00 , ξ 000 and ξ 0000 represent diagonalised vectors of the derivatives over the
physical domain given by Equations 7.26a-d at the Gauss-Lobatto points defined by
Equation 7.15, and T is the Chebyshev transformation matrix (7.18). For Poiseuille or
Couette flow a linear mapping can be chosen such that Dn (y ∗ ) = Dn (ξ), eliminating
the need for a transformation by setting the physical domain to lie in −1 ≤ y ∗ ≤ 1.
But in the Blasius case the flow is defined between 0 ≤ y ∗ ≤ ∞, and therefore the
transformations given in Equations 7.24 Equations 7.24 through

7.4 Temporal formulation


The temporal analysis assumes a real wavenumber and a complex frequency, and
yields the solution in terms of temporal growth. The frequency is found as the
solution to a first order nonlinear differential eigenproblem. Using eigenfunctions
containing spectral coefficients, v̂ and η̂y , Equation 7.14 can be written in temporal
form.

! ! ! !
LOS 0 v̂ D2 − k2 0 v̂
= iω (7.28)
βU 0 − αW 0 LSQ η̂y 0 1 η̂y

where

i 2
LOS = (αU + βW ) D2 − k 2 − αU 00 − βW 00 + D2 − k2

(7.29a)
Re
i
D2 − k2

LSQ = αU + βW + (7.29b)
Re

Equations 7.29a and 7.29b represent and Squire differential operators respectively.
Equations 7.28 are subject to the boundary conditions ve = D (y ∗ ) ve = ηey = 0
at solid walls at in the free stream. All matrix terms that are not multiplied by
a differentiation matrix, Di (y ∗ ), are multiplied by the Chebyshev transformation
matrix, T, because the eigenfunctions are spectral coefficients. U and W represent
matrices containing the diagonalised vectors of streamwise and spanwise velocity at
the Gauss-Lobatto collocation points.
The phase speed, c, is often more useful than the frequency, ω, and their relation-
ship is defined:
68 CHAPTER 7. LINEAR STABILITY ANALYSIS

ω = αc (7.30)

where c is made of cr and ci , the real and imaginary parts of the phase speed, re-
spectively. The coefficient, α, is the streamwise wavenumber which appears in the
approximated form of the disturbance in Equation 7.13.
Equations 7.29 represent two 2-by-2 block matrices, into which boundary condi-
tions are introduced as follows. The summation of all Chebyshev polynomials at ±1
must equal zero to satisfy the boundary conditions on Equation 7.28. Hence, making
use of the identities in (7.23) the two upper and two lower rows of the block matrices
for LOS and LSQ are modified:

 
··· BC1k ···
 

 ··· BC2k ··· 

 a
 3,1 a3,2 ··· a3,N a3,N +1  
 . .. .. .. . 
LOS = .. . . . ..  (7.31)
 
aN −1,1 aN −1,2 · · ·
 
 a N −1,N aN −1,N +1 


 · · · BC3 k · · · 

··· BC4k ···

where

BC1k = (1)k
BC2k = (−1)k
(7.32)
BC3k = (1)k+1 k 2
BC4k = (−1)k+1 k 2

and k = 1, 2, . . . , N + 1. The upper and lower rows were chosen due to reduce
discretisation errors, which are largest at the ends of the computational domain.
Corresponding upper and lower rows of other blocks of Equation 7.28 are set to zero,
thus satisfying the boundary conditions.

7.5 Spatial formulation


Temporal analysis provides growth rates in time, while spatial analysis gives the
growth of instabilities in the spatial domain. The spatial formulation is presented
7.5. SPATIAL FORMULATION 69

in this section for completeness. Here, the frequency is assumed to be real and the
complex wavenumber is solved as the solution to a fourth order differential eigenprob-
lem (7.14). First, the problem will be formulated as a fourth order eigenproblem. Due
to the difficulty of solving such a problem, however, a Haj-Hariri (1988) transforma-
tion will be used to halve the order of the equations.

7.5.1 4th order formulation

The governing equations (7.14) can be written as

   
i 4 3 2i 2 2
 2  2 2
 00

α − U α + ω − βW − D −β α + U D − β − U α ṽ
Re Re
 
2 2
 00 i 2 2 2

+ (βW − ω) D − β − βW + D −β ṽ = 0 (7.33a)
Re

  
i 2 i 2 2
η̃y + βU 0 ṽ − αW 0 ṽ = 0 (7.33b)

− α + αU + βW − ω + D −β
Re Re

where Di represents the ith -order transformed Chebyshev differentiation matrix with
respect to y ∗ . Defining z̃ = η̃ṽ , Equations 7.33 can be written in a more general
 

form.

A0 z̃ + αA1 z̃ + α2 A2 z̃ + α3 A3 z̃ + α4 A4 z̃ = 0 (7.34)

where

!
L0 0
A0 = (7.35a)
βU 0 S0
!
L1 0
A1 = (7.35b)
−W 0 U
70 CHAPTER 7. LINEAR STABILITY ANALYSIS

!
L2 0
A2 = i
(7.35c)
0 − Re
!
−U 0
A3 = (7.35d)
0 0
!
i
Re
0
A4 = (7.35e)
0 0

and

i 2
L0 = (βW − ω) D2 − β 2 − βW 00 + D2 − β 2

(7.36a)
Re
i 2 2

S0 = βW − ω + D −β (7.36b)
Re
L1 = U D2 − β 2 − U 00

(7.36c)
2i
D2 − β 2

L2 = ω − βW − (7.36d)
Re

Unfortunately, Equation 7.33 is difficult to solve accurately, and therefore the second
order formulation is usually used.

7.5.2 Second order formulation

It is common to use a Haj-Hariri (1988) transformation to halve the order of the


equations.

! !
ṽ Ṽ (−αy∗ )
= e (7.37)
η̃y Ẽ

where α is the streamwise wavenumber of the disturbance, which is assumed to be


of the form given in Equation 7.13. Note that this transformation of the disturbance
vectors ṽ and η̃y does not decrease the generality of the solutions to the governing
Equations (7.14). Using this transformation, transformation, Equations 7.13a and
7.13b are reduced to a second order nonlinear set of equations:
7.5. SPATIAL FORMULATION 71

(iω − iαU − iβW ) D2 − 2αD − β 2 Ṽ


 
 
00 00 1 2 2 2

+ iαU + iβW + D − 2αD − β Ṽ = 0 (7.38a)
Re

1
(iω − iαU − iβW ) Ẽ − iβU 0 Ṽ + iαW 0 Ṽ + D2 − 2αD − β 2 Ẽ = 0

(7.38b)
Re

Using the transformation of Equation 7.24, Equations 7.38 can be reduced to a linear
 T
problem by introducing the vector quantity αṼ , Ṽ , Ẽ , which leads to a linearised
matrix formulation.

     
−R1 −R0 0 αV̂ R2 0 0 αV̂
 I 0 0  V̂  = α 0 I 0  V̂  (7.39)
     

−Q −S −T0 Ê 0 0 T1 Ê

Where

4 2
R2 = D + 2iU D (7.40a)
Re
4 3 4 2
R1 = −2iωD − D + β D − iU D2 + iDβ 2 + iU 00 + 2iβD (7.40b)
Re Re
1 4 2 2 2 1 4
R0 = iωD2 − iωβ 2 − iβW D2 + iβ 3 W + iβW 00 + D − β D + β (7.40c)
Re Re Re
2
T1 = D + iU (7.40d)
Re
1 2 1 2
T0 = −iω − D + β − iβW (7.40e)
Re Re
S = iβU 0 (7.40f)
Q = −iW 0 (7.40g)

and Di represents the ith -order transformed Chebyshev differentiation matrix with
respect to y ∗ defined by Equations 7.27. As in the temporal formulation, all matrix
terms not multiplied by a Chebyshev derivative matrix, Di are multiplied by the
Chebyshev transformation matrix, T because the eigenfunctions are spectral coeffi-
cients. Matrices U and W again contain the diagonalised vectors of streamwise and
spanwise velocity at the Gauss-Lobatto collocation points. Boundary conditions are
72 CHAPTER 7. LINEAR STABILITY ANALYSIS

1.8

1.6
0.6 0.2

1.4

1.2 0.5

1 0.1
ε

0.4

0.8

0.3
0.6

0.2
0.4

0.1
0.2

0
−0.5 0 0.5 1 1.5 2
δ

Figure 7.1: Stability diagram for Mathieu’s equation, generated using 9 modes (k =
−4, −3, . . . , 3, 4).

introduced in the manner presented in Section 7.4.

7.6 Generalised time-dependent temporal formu-


lation
The temporal and spatial stability analyses presented in the preceding sections are
steady-state formulations. The assumptions introduced in Equations 7.13 are not
valid for time-dependent base flows. Hence, a different approach must be taken to
analyse oscillatory forcing, which is presented here.

7.6.1 Introduction to parametric forcing - Mathieu’s equa-


tion
To understand the following analysis, it is useful to use a simple model problem to
describe the basic analysis technique. Mathieu’s equation (Mathieu, 1868) occurs
in many areas of physics from free-surface waves to quantised electron energy states.
Due to its relative simplicity, it has become the principal model equation for studying
and illustrating the effects of parametric forcing. Mathieu’s equation is given by

d2 y
+ (δ +  cos t) y = 0 (7.41)
dt2
7.6. TIME DEPENDENT FORMULATION 73

15

10

5
imaginary part

−5

−10

−15
−0.1 −0.05 0 0.05 0.1 0.15
real part

Figure 7.2: Eigenvalues, λ, of Mathieu’s equation for δ = 0.56 and  = 0.8 using 21
modes (k = −10, −9, . . . , 9, 10).

It is useful to re-write this equation in complex-exponential form. By using conjugate


pairs of time dependent terms, the complex part can be eliminated.

1
 cos t =  eit + e−it

(7.42)
2

When  is nonzero, the term of period T = 2π introduces parametric forcing into


the equation. According to Floquet theory (Floquet, 1883; Kuchment, 1993), the
solutions are of the form:

X
y (t) = eλt ak eikt ak ∈ Z (7.43)
k

where ak is the multiplying coefficient for mode k, and λ is the complex Floquet
exponent. This Floquet exponent, λ, is found as the solution to an eigenproblem.
Equation 7.43 represents a decomposition into an exponential part and a purely oscil-
latory part. Substituting Equations 7.42 and 7.43 into Equation 7.41, then balancing
terms of equal harmonic dependence (eikt ) gives a nonlinear eigenproblem for λ.
74 CHAPTER 7. LINEAR STABILITY ANALYSIS

1
(ik + λ)2 ak + δak +  (ak−1 + ak+1 ) = 0 (7.44)
2

Appearance of ak−1 and ak+1 terms shows the coupling of all frequencies, and
therefore a single forcing frequency generates a range of response frequencies. The
response frequency may not be at the same frequency as that of the excitation. After
simple manipulations, Equation 7.44 can be reformulated as a linear eigenproblem.

" # " #
ak Ak
λ = (7.45)
Ak (k 2 − δ) ak − 2ikAk − 21  (ak−1 + ak+1 )

Writing Equation 7.45 in matrix form yields:

" # " #" #


a 0 I a
λ = (7.46)
A P Q A

where

.  . 
.. ..
   
a=
ak 
 A=
Ak 
 (7.47)
.. ..
. .
 
.
 .. ... ... ...
  

   
1 2 1
P= −2 k − δ −2 Q= −2ik (7.48)
  
  
.. .. ..  ..
 

 . . . .

and I and 0 represent identity and zero matrices, respectively. The series of Fourier
modes (7.43) must be truncated in order to numerically solve the problem.
The effect of truncating the infinite Floquet series of Equation 7.43 is shown in
Figure 7.2. The eigenvalues (Floquet exponents) near the ends of the truncated
series are erroneous and therefore ignored. These erroneous eigenvalues can be seen
in Figure 7.2, which appear as the only unstable eigenvalues for δ = 0.58 and  = 0.8.
It was found that not many modes are required to calculate accurate results.
The stability plot calculated for Figure 7.1 was made using only 9 modes (k =
−4, −3, . . . , 3, 4), and the inclusion of more modes did not result in any notable
7.6. TIME DEPENDENT FORMULATION 75

difference for the domain studied. Solutions within the solid lines are purely os-
cillatory, while solutions outside the regions of stability show exponential growth at
rates given by the contours. Because the range of  is relatively small, the stability
plot of Figure 7.1 agrees well with the known stability plot for the Mathieu equa-
tion (McLachlan, 1947; Arscott, 1964). However, increasing numbers of modes are
required to find stable solutions as  increases.

7.6.2 Linearised stability equations


The analysis herein follows that of Schmid and Henningson (2001). However, a similar
analysis was made by von Kerczek (1982), who studied the effect of oscillatory pressure
gradient on Poiseuille flow. Starting with the Orr-Sommerfeld and Squire Equations
of equations 7.9 and 7.12,

   
∂ 2 ∂ 2 00 ∂ 1 4 ∂ 2 00 ∂
∇v= − U ∇ +U + ∇ −W ∇ +W v (7.49a)
∂t ∂x ∂x Re ∂z ∂z
   
∂ ∂ 1 2 ∂v 0 ∂v 0 ∂
ηy = −U + ∇ ηy − U + W − W ηy (7.49b)
∂t ∂x Re ∂z ∂x ∂z

it is assumed that the streamwise base velocity U and the spanwise base velocity
W are composed of steady and sinusoidal components. Equation 7.49 represents a
parametrically forced equation, with similarities to the Mathieu equation described
in Section 7.6.1. Assuming time-periodic solutions according to Floquet theorem, the
following substitutions can be made:

AW  ∗

W (y ∗ , t) = W (y ∗ ) + W eiΩt + W e−iΩt eiφ (7.50a)
2
A U
 ∗

U (y ∗ , t) = U (y ∗ ) + U eiΩt + U e−iΩt (7.50b)
2X
v=e i(αx+βz) λt
e vk (y ∗ ) eikΩt (7.50c)
k
X
ηy = e i(αx+βz) λt
e ηk (y ∗ ) eikΩt (7.50d)
k

where Ω is the forcing frequency and φ is the phase difference between the streamwise
and spanwise oscillations. Coefficients AU and AW represent the peak amplitude of
oscillatory velocity in the streamwise and spanwise directions, respectively. The base
76 CHAPTER 7. LINEAR STABILITY ANALYSIS

flow profiles in the streamwise and spanwise directions are given by the vectors U
and W , and oscillatory profiles are defined by the vectors U and W , respectively. In
all cases studied, streamwise and spanwise oscillations were studied independently,
making this phase difference redundant, and thus φ was set to 0. As in the problem
modelled by the Mathieu equation, terms of equal harmonic dependence (i.e. of the
same eiCΩt where C = Z+ ) can be balanced.

 
2 00 2 00 2 i 4 2
− iλ∇ vk = αU − αU ∇ + βW − βW ∇ − ∇ − kΩ∇ vk +
Re

αAU  00  βA eiφ  00 
2 W 2
U − U∇ + W − W∇ vk+1
2 2

αAU  ∗ 00 ∗ 2
 βA eiφ  ∗ 00 ∗ 2

W
+ U −U ∇ + W −W ∇ vk−1 (7.51a)
2 2

 
i 2
− iληk = −αU − βW − ∇ − kΩ ηk + [−βU 0 + αW 0 ] vk
Re
βAW eiφ βAU 0 αAW eiφ 0
   
αAU
+ − U− W ηk+1 + − U + W vk+1
2 2 2 2
αAU ∗ βAW eiφ ∗ βAU ∗ 0 αAW eiφ ∗ 0
   
+ − U − W ηk−1 + − U + W vk−1 (7.51b)
2 2 2 2

An asterisk (∗ ) represents complex conjugation. Equations 7.51 represent an eigen-


problem in which each mode is coupled to all others. Due to the this coupling,
oscillatory response at frequencies different to that of the forcing frequency can be
excited. Again, the number of modes must be truncated and modes at the end of
the series ignored due to truncation error. Re-writing Equations 7.51 in a discretised
matrix form with eigenfunctions of spectral coefficients, v̂ and η̂, yields:

λAv̂k = Bk v̂k + Cv̂k+1 + Dv̂k−1 (7.52a)


λIη̂k = Fk η̂k + Gv̂k + Hη̂k+1 + Kv̂k+1 + Lη̂k−1 + Mv̂k−1 (7.52b)

where
7.6. TIME DEPENDENT FORMULATION 77

A = −i∇2 (7.53a)
I = −iI (7.53b)
i 4
Bk = αU 00 − αU ∇2 + βW 00 − βW ∇2 − ∇ − kΩ∇2 (7.53c)
Re
αAU  00  βA eiφ  00 
W
C= U − U ∇2 + W − W ∇2 (7.53d)
2 2
iφ 
αAU  ∗ 00 ∗ 2
 βA We ∗ 00 ∗ 2

D= U −U ∇ + W −W ∇ (7.53e)
2 2
i 2
Fk = −αU − βW − ∇ − kΩ (7.53f)
Re
G = −βU 0 + αW 0 (7.53g)

αAU βAW e
H=− U− W (7.53h)
2 2
βAU 0 αAW eiφ 0
K=− U + W (7.53i)
2 2
αAU ∗ βAW eiφ ∗
L=− U − W (7.53j)
2 2
βAU ∗ 0 αAW eiφ ∗ 0
M=− U + W (7.53k)
2 2

where ∇2 = D2 − k 2 and I is a 2 (N + 1) identity matrix. All matrix terms that are


not multiplied by a differentiation matrix, Di , are multiplied by Chebyshev trans-
formation matrix, T, because the eigenfunctions are spectral coefficients. U and W
represent matrices containing the diagonalised vectors of streamwise and spanwise
velocity at the Gauss-Lobatto collocation points. Again, the number of modes must
be truncated to solve the problem numerically. As in the Mathieu equation, modes
near the end of the truncated Floquet series are grossly erroneous due to truncation
errors and must be neglected. Defining the vector of spectral coefficients, ẑk :

" #
v̂k
ẑk =
η̂k

Equations 7.52 can then be rewritten.

" # " # " # " #


Ak 0 Bk 0 C 0 D 0
λ ẑk = ẑk + ẑk+1 + ẑk−1 (7.54)
0 I G Fk K H M L
78 CHAPTER 7. LINEAR STABILITY ANALYSIS

Rewriting Equation 7.54 as a single matrix eigenproblem yields:

λPk ẑk = Qk ẑk + Rẑk+1 + Sẑk−1 (7.55)

Thus, the entire truncated series of modes can be included in the full coupled system.

     
P0 ẑ0 Q0 R ẑ0
P1 ẑ1   S Q1 R ẑ1 
     

λ ..
 .  = 
 .   .. .. . .  .. 
  (7.56)

 .  .   . . .  . 
Pk ẑk S Qk ẑk

7.7 Solution method


The time-independent temporal and spatial equations (7.28,7.39) were solved using
an eigenvalue solver in Matlab. The computer code was modularised to allow Couette,
Poiseuille, and Blasius flows to be analysed.
Time-dependent analysis of Equation 7.56 was undertaken using a similar modular
Matlab code. The number of modes used was limited by computational power (Intel
Pentium 4, 3.2GHz, 2GB RAM). However, some studies have achieved reasonable
accuracy using only a few modes (Joslin and Morris, 1992). It should be noted that
the system under consideration (Equation
refeqn:timedep:stability:6) is a viscous fluid system subject to harmonic shearing,
which exhibits conjugate-translation (CT) symmetry (Or, 1997). In support of the
work by Or (1997), Schulze (1999) showed that such systems do not exhibit sub-
harmonic responses. The therefore been omitted from 7.56. Results Equation 7.56.
Results using a Floquet detuning constant yielded identical growth rates to those
with no detuning constant.
Chapter 8

Stability analysis results

A Matlab code was developed to solve the linearised stability equations (7.28,7.39
and 7.56). The programs used for the temporal stability calculations are presented
in Appendix A. A selection of standard results for time-independent non-oscillatory
flows are presented here. These results are in excellent agreement with the work of
Orszag (1971), Theofilis (1994), and Schmid and Henningson (2001), confirming the
validity of the time-dependent code. Spatial results presented in Section 8.2 also
agree well with previous studies. Results are also presented for the time-dependent
formulation of the equations, where parametric oscillatory forcing of the boundary
layer is shown to have a stabilising effect at low frequencies of oscillation.

8.1 Temporal formulation

8.1.1 Plane Couette flow

Plane Couette flow has a simple linear variation in streamwise velocity, and therefore
has the simplest eigenvalue spectrum. The Orr-Sommerfeld eigenvalue spectrum for
this flow is shown in Figure 8.1 for a Reynolds number of 1000, with α = 1 and
β = 1. In this example, 250 collocation points have been used (N = 250), and results
agree with those of Schmid and Henningson (2001). The flow has both rotational
symmetry about the channel center-point and reflective symmetry about the channel
center-line, resulting in a symmetrical eigenvalue spectrum. Each eigenvalue has a
corresponding eigenfunction, which gives the mode-shape of the disturbance for the
corresponding eigenvalue. In cases where the base flow is symmetrical about the flow
centreline, these eigenfunctions are either symmetrical or antisymmetrical about the
centreline.

79
80 CHAPTER 8. STABILITY ANALYSIS RESULTS

−0.2

−0.4

−0.6

−0.8
ci

−1

−1.2

−1.4

−1.6

−1.8

−2
−1 −0.5 0 0.5 1
c
r

Figure 8.1: Orr-Sommerfeld spectrum for plane Couette flow at Re = 1000, α = 1


and β = 1 using 250 collocation points.

8.1.2 Plane Poiseuille flow

Plane Poiseuille flow is also a benchmark flow profile for stability analysis. A typical
Orr-Sommerfeld eigenspectrum is shown in Figure 8.2, which has the same three-
branch structure as plane Couette flow. These branches are typically called A, P,
and S branches, as shown in Figure 8.2. Eigenvectors with an imaginary part greater
than 0 (shown by the grey area) are unstable and result in solutions that grow in
time according to eci t . It can be shown that the solutions to the Squire equation are
always damped (Schmid and Henningson, 2001). Thus, in cases where both β and
spanwise velocity are zero, the governing system of equations (7.28) is uncoupled and
only Orr-Sommerfeld modes need to be solved to determine stability.
Curves that define the boundary between areas in parameter space where expo-
nentially growing solutions exist, and where only stable solutions exist, are called
neutral curves. Figure 8.3 shows neutral curves for Poiseuille flow with β = 0, to-
gether with curves showing constant nonzero growth rates and phase speed of the
least stable mode, generally a TS wave. The point at the left-most tip of the neutral
stability curve, shown by the thick line, defines the lowest Reynolds number for which
unstable solutions exist, which is the critical Reynolds number, Recrit . The largest
growth rate in perfect plane Poiseuille flow is ωi = 0.007688 at a Reynolds number
of approximately 46950, with a phase speed of α = 0.782 (Schmid and Henningson,
8.1. TEMPORAL FORMULATION 81

0.1

−0.1

−0.2 P
−0.3 A
−0.4
ci

−0.5

−0.6

−0.7
S
−0.8

−0.9

−1
0 0.2 0.4 0.6 0.8 1
c
r

Figure 8.2: Orr-Sommerfeld spectrum for plane Poiseuille flow at Re = 10000, α = 1


and β = 0 using 250 collocation points.

2001). Thus, such a disturbance will grow in size according to:

A1
= eωi (t1 −t0 ) (8.1)
A0

and therefore the disturbance will double in approximately 90.2 time units. Similar
stability curves to that of Figure 8.3 can be drawn for nonzero values of β. Figure 8.4
shows growth rates and phase speeds of the most unstable mode of plane Poiseuille
flow at Re = 10000. The form of this plot is typical for all Reynolds numbers, with
growth rates decreasing with increasing values of β.

8.1.3 Blasius boundary layer flow


The Blasius boundary layer has, at most, one unstable mode, which is a TS wave. The
eigenvalues for Blasius flow have a branched structure that differs to those of Couette
or Poiseuille flow. The eigenvalues consist of a finite number of discrete eigenvalues
and a so-called continuous spectrum (Schmid and Henningson, 2001). This continuous
spectrum results from the discrete representation of the infinite domain. Such a
continuous spectrum does not exist for flows in bounded domains, like Poiseuille flow,
because all eigenvalues are discrete, although infinite in number.
82 CHAPTER 8. STABILITY ANALYSIS RESULTS

−0.05
1.8

−0.04
1.6
−0.03

1.4 −0.02

−0.01

1.2
0
α

0.8

0.6
−0.04
−0.05
−0.06
0.4 −0.0 −0.07
8

0.2
1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
Reynolds number

1.8
0.2
5

1.6
5
0.3

1.4

1.2 0.3
α

0.25
0.8

0.2
0.6

0.15
0.4

0.2
1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
Reynolds number

Figure 8.3: Contours of constant growth rate, ci (top); and constant phase velocity,
cr (bottom) of Poiseuille flow for β = 0.
8.1. TEMPORAL FORMULATION 83

1.8

1.6

1.4

1.2 −0.01 −0.0


2 −0. −0
03 .04
α

0.8

0.6

0.4

0.2
0 0.2 0.4 0.6 0.8 1 1.2 1.4
β

1.8

1.6

0.25
1.4

1.2
0.25
α

0.275

1
0.225

0.8
0.2 0.3

0.6
0.17
5

0.4 0.1 25
5 0.3

5
0.3
0.2
0 0.2 0.4 0.6 0.8 1 1.2 1.4
β

Figure 8.4: Maximum growth (top) and phase of the least stable mode (bottom) of
Poiseuille flow for Re = 10000 with non-zero β.
84 CHAPTER 8. STABILITY ANALYSIS RESULTS

0.1 0.1

0 0

−0.1 −0.1

−0.2 −0.2

−0.3 −0.3

−0.4 −0.4
ci

ci
−0.5 −0.5

−0.6 −0.6

−0.7 −0.7

−0.8 −0.8

−0.9 −0.9

−1 −1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
c c
r r

0.1 0.1

0 0

−0.1 −0.1

−0.2 −0.2

−0.3 −0.3

−0.4 −0.4
ci

ci

−0.5 −0.5

−0.6 −0.6

−0.7 −0.7

−0.8 −0.8

−0.9 −0.9

−1 −1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
c c
r r

0.1 0.1

0 0

−0.1 −0.1

−0.2 −0.2

−0.3 −0.3

−0.4 −0.4
ci

ci

−0.5 −0.5

−0.6 −0.6

−0.7 −0.7

−0.8 −0.8

−0.9 −0.9

−1 −1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
c c
r r

0.1

−0.1

−0.2

−0.3

−0.4
ci

−0.5

−0.6

−0.7

−0.8

−0.9

−1
0 0.2 0.4 0.6 0.8 1
c
r

Figure 8.5: Typical temporal eigenvalue distribution for the Blasius boundary layer.
Re = 500, α = 0.2, β = 0 and N = 25, 50, 75, 100, 125, 150, and 175, descending
from left to right respectively. The instability region lies in Im > 0, shown by the
greyed area.
8.1. TEMPORAL FORMULATION 85

0.5

−0
.0
0.45 25

−0 −0
0.4 .01 .03 −0.05
25 75

0.35
0

0.3
0.0125
α

0.25

0.2

0.15

0.1

0.05
500 1000 1500 2000 2500 3000 3500 4000 4500 5000
Reynolds number

0.5

0.
25
0.
2

0.45
75

0.4
0.375

0.35
0.35

0.3
5 7
0.3

5
α

0.3

5
32
0.25 0.
0.3

0.275
0.2

0.15
0.25

0.1

0.05
500 1000 1500 2000 2500 3000 3500 4000 4500 5000
Reynolds number

Figure 8.6: Maximum growth (top) and phase speed of the least stable mode (bottom)
for the Blasius boundary layer for β = 0. These results are almost identical to those
of Schmid and Henningson (2001).
86 CHAPTER 8. STABILITY ANALYSIS RESULTS

0.5

0.45
−0.0
6
−0.0
−0.0 5
0.4 −0.0 4
−0.0 3
2
−0.0
1
0.35 0

0.3 0.01
α

0.25

0.2

0.15

0.1

0.05
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
β

0.5

0.375

0.45

0.4
0.375
0.4
0.35

0.3
α

0.25 0.35

0.2

0.3
25 5
42
0.
0.15

0.3
5
0.1 0.27 0.4
5
5
0.47

0.05
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
β

Figure 8.7: Maximum growth (top) and phase of the least stable mode (bottom)
for the Blasius boundary layer for Re = 1000 with non-zero β. Note the region of
instability, bounded by the thick contour line.
8.2. SPATIAL FORMULATION 87

35

30

25

20
y

15

10

0
−0.5 0 0.5 1 1.5 2 2.5 3
magnitude

Figure 8.8: Eigenfunction corresponding to the least stable growth rate of the Blasius
boundary layer at Re = 500, α = 0.2 and β = 0.

It can be seen in Figure 8.5 that the continuous spectrum of Blasius flow converges
quickly to a near-vertical line below the real axis at cr = 1 as the number of collocation
points increases. The discrete eigenvalues that govern the stability of the flow appear
to the left of the continuous spectrum, which sweeps down to reveal more stable
discrete eigenvalues with increasing numbers of collocation points. For the case shown,
with a boundary-layer Reynolds number of 500 and a streamwise wavenumber of
α = 0.2, the supremum eigenvalue is stable. The corresponding eigenfunction for this
eigenvalue is shown in Figure 8.8.

Figure 8.6 shows neutral stability curves and contours of constant phase of the
least stable eigenvalue for Blasius flow for β = 0. The discontinuity in the phase plot
represents a change in identity of the most unstable eigenvalue. Similar plots for a
constant Reynolds number of 1000 and varying spanwise wavenumber are shown in
Figure 8.7. These are in exact agreement with the results of Schmid and Henningson
(2001) and also are also validated by matching results from previous studies (e.g.
Orszag, 1971; Theofilis, 1994).
88 CHAPTER 8. STABILITY ANALYSIS RESULTS

0.8

0.7

0.6

0.5

0.4
Im

0.3

0.2

0.1

−0.1
0 0.2 0.4 0.6 0.8 1
Re

Figure 8.9: Spatial eigenvalue spectrum of streamwise wavenumbers of Poiseuille flow


for Re = 2000, ω = 0.3, β = 0 and N = 200. Eigenvalues with negative imaginary
part are unstable.

8.2 Spatial formulation


8.2.1 Plane Poiseuille flow
The spatial eigenspectrum is more complicated than the temporal spectrum. The
spatial formulation provides an eigenspectrum of complex disturbance streamwise
wavenumbers, rather than an eigenspectrum of complex disturbance frequency. Thus,
unstable solutions to the spatial equation grow in space, and eigenvalues can appear in
all four quadrants of the imaginary plane. Eigenvalues of interest appear in the upper
right quadrant, and cause unstable spatial growth downstream when these discrete
eigenvalues cross into the lower right quadrant. Eigenvalues in the left quadrants affect
the upstream perturbation dynamics (Schmid and Henningson, 2001). An example
result for plane Poiseuille flow is presented in Figure 8.9 to show the form of the
eigenspectrum.

8.2.2 Blasius boundary layer flow


An illustrative selection of eigenspectra for Blasius flow are presented in Figure 8.10.
Similar trends can be seen to those of the temporal case, where increasing numbers
of collocation points reveal the discrete eigenvalues. Results shown here agree very
well with those presented in Schmid and Henningson (2001).
8.2. SPATIAL FORMULATION 89

0.8 0.8

0.7 0.7

0.6 0.6

0.5 0.5

0.4 0.4
Im

Im
0.3 0.3

0.2 0.2

0.1 0.1

0 0

−0.1 −0.1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Re Re

0.8 0.8

0.7 0.7

0.6 0.6

0.5 0.5

0.4 0.4
Im

Im

0.3 0.3

0.2 0.2

0.1 0.1

0 0

−0.1 −0.1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Re Re

0.8 0.8

0.7 0.7

0.6 0.6

0.5 0.5

0.4 0.4
Im

Im

0.3 0.3

0.2 0.2

0.1 0.1

0 0

−0.1 −0.1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Re Re

0.8 0.8

0.7 0.7

0.6 0.6

0.5 0.5

0.4 0.4
Im

Im

0.3 0.3

0.2 0.2

0.1 0.1

0 0

−0.1 −0.1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Re Re

Figure 8.10: Spatial eigenvalue spectrum for Blasius flow for Re = 1000, ω = 0.26,
β = 0 and N = 25, 50, 75, 100, 125, 150, 200, 250 descending from left to right,
respectively.
90 CHAPTER 8. STABILITY ANALYSIS RESULTS

8.3 Time-dependent temporal formulation

0.8

0.6

0.4

0.2
Im

−0.2

−0.4

−0.6

−0.8

−1
−1 −0.8 −0.6 −0.4 −0.2 0
Re

Figure 8.11: Typical appearance of the eigenvalue spectrum for the time-dependent
case between limits of -1 and 1, for Re = 804, α = 0.266, a spanwise forcing with
Ω = 0.06, 64 collocation points and 11 modes. Note the similarities in each spectrum
for each mode, and the similarities with the plot of eigenvalues for Mathieu equation
in Figure 7.2.

The time-dependent formulation results in a number of eigenspectra arrayed in


the imaginary plane, each corresponding to a different Floquet mode, as shown in
Figure 8.11. In this figure, 64 collocation points were used (N = 64) and 11 modes
were used (M = 10, k = −5, −4, . . . , 4, 5). The full range of eigenvalues has been
truncated for clarity. At the upper and lower extremities of the array lie spurious
values that are the result of truncating the infinite series of modes in Equation 7.56.
Note the similarity between the spectrum corresponding to each mode. This similarity
implies that the number of modes used is sufficient.
Most eigenvalues in Figure 8.11 represent the stable continuous spectrum, which
are concentrated to the immediate left of the imaginary axis. The major eigenvalues
from the discrete spectrum are of interest, which in this case are unstable and circled.
The eigenspectrum corresponding to each mode appears in series, with the continuous
spectrum of mode k = 0 beginning at cim = −1.
Typically, if a non-spurious growing solution existed, it appeared in the first mode
and was therefore between cim = −1 and cim = 1. To eliminate spurious unstable
eigenvalues (λs ), non-spurious eigenvalues (λn ) were chosen according to
8.3. TIME-DEPENDENT TEMPORAL FORMULATION 91

0.5

0.45

0.4

0.35

0.3
0.015
α

0.25 0.02

0.2 0.025

0.15

0.1

0.05
500 1000 1500 2000 2500 3000 3500 4000 4500
Reynolds Number

Figure 8.12: Stability curves of Blasius flow for streamwise forcing of amplitude 0.05
and Ω = 0.18 (N = 64,M = 10)

[
|λn | < |λs | |λn | <1 (8.2)

This proved effective for most of the range of parameters tested. For small amplitudes
(A < 0.04), however, when the imaginary part of some spurious eigenvalues could lie
between −1 and 1 the spurious eigenvalues were not always eliminated. In these
cases, the spectrums from each mode are very close together, and caused difficulty
in distinguishing spurious eigenvalues from non-spurious ones. By including these
spurious eigenvalues, results for these small amplitudes are believed to be worst-case
growth rates, and the true theoretical growth rates may be lower.

8.3.1 Streamwise forcing


Figure 8.12 shows stability curves for streamwise forcing of amplitude 0.05 and Ω =
0.18. The forcing frequency is given by Ω/2π. Due to the difficulty in determining
discrete, negative growth rates, regions of stability are not contoured. It can be
seen that the neutral curve, shown by the thick line, is not greatly affected by small
amplitude forcing.
92 CHAPTER 8. STABILITY ANALYSIS RESULTS

0.9

0.8 0.06

0.7
Amplitude, AU

0.6
0.05
0.5
0.04
0.4
0.03
0.3
0.02
0.2
0.01
0.1

0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Figure 8.13: Stability curves of Blasius flow for streamwise forcing of the profile shown
in Figure 8.14 at Re = 700 and α = 0.25. 64 collocation points and 11 modes were
used.

The oscillatory forcing profile used in Figure 8.12 is shown in Figure 8.14 (top
plot), and the resulting streamwise velocity profile (bottom plot) is shown for a larger
amplitude of 0.2 times the freestream velocity for clarity. Comparison with the stan-
dard stability curves indicates this forcing destabilises the flow slightly. Indeed, it was
found that forcing at frequencies higher than 0.06/2π typically resulted in increased
instability of the flow.
The effect of amplitude and oscillation for an arbitrary Reynolds number of 700
and α = 0.25 can be seen in Figure 8.13. The comparatively low number of grid
points results in the staggered appearance. The large computational cost involved in
solving Equation 7.56 limited the number of which could be solved in a reasonable
time. Islands of instability observed near the upper left of this plot are likely to be
spurious, as selected runs using greater collocation points and modes showed these
regions are in fact stable. Indeed, just as in the Mathieu equation (7.47), it was found
that increasing the number of modes used typically resulted in the solutions being
more stable.
Figure 8.15 shows the results of different numbers of collocation points and modes
used. The top plot shows one of the spurious islands of instability found in Figure 8.13.
This region is unstable when only 11 modes are used, but stable when 21 modes are
used. The lower plot of Figure 8.15 shows that for higher frequencies (greater than
8.3. TIME-DEPENDENT TEMPORAL FORMULATION 93

50
real part
45 imaginary part

40
nondimensional height, y

35

30

25

20

15

10

0
−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4
Streamfunction

50

45

40
nondimensional height, y

35

30

25

20

15

10

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
nondimensional spanwise velocity, W

Figure 8.14: Forcing profile A. Normalised complex function describing oscillatory


velocity profile (upper); resulting velocity profiles over 1 period for a streamwise
oscillation amplitude of 0.2 times freestream velocity (lower).
94 CHAPTER 8. STABILITY ANALYSIS RESULTS

−3
x 10
6
N = 32, M = 11
N = 64, M = 11
5 N = 80, M = 21

4
growth rate

0
0 0.2 0.4 0.6 0.8 1
Streamwise oscillation amplitude, A
U

0.035
N = 32, M = 11
N = 64, M = 11
0.03 N = 80, M = 21

0.025
growth rate

0.02

0.015

0.01

0.005

0
0 0.05 0.1 0.15 0.2 0.25 0.3

Figure 8.15: Comparison of different number of collocation points and modes used
for Re = 700 and Ω = 0.06 (upper) and streamwise amplitude of 0.1 (lower). The
forcing profile is that shown in Figure 8.14.
8.3. TIME-DEPENDENT TEMPORAL FORMULATION 95

Ω = 0.1), results using only 32 collocation points and 11 modes agree well with more
accurate results calculated using 64 points and 21 modes.
It can be seen from Figure 8.16 that some frequencies of forcing can result in
excitation of unstable modes of low streamwise wavenumber. This excitation is also
greatly influenced by the frequency and shape of the oscillatory component, as can
be seen in Figure 8.17. The upper plot in Figure 8.17 is using forcing profile A
(Figure 8.14) at a frequency of 0.4/2π. Clearly the critical Reynolds number is sig-
nificantly decreased, and the unforced instability curve is enlarged considerably. On
the other hand, forcing of a different profile (Figure 8.18) can result in significant
modification of the instability modes even at low amplitudes.
However, forcing of profile A at low frequencies of oscillation was found to stabilise
the Blasius boundary layer. Two example plots of stability are given in Figures 8.19
and 8.20. Figure 8.19 illustrates that increasing the number of modes used typically
resulted in increased stabilisation. It can be seen that the critical Reynolds number
can be increased more than twofold by suitable streamwise forcing. At Ω = 0.01
and forcing amplitude of 0.2, the entire Re-α domain studied was stabilised up to a
Reynolds number of 3600 (0.05 ≤ α ≤ 0.5). At this frequency, computations typically
became subject to at error higher Reynolds numbers due to an insufficient number of
modes. For Ω < 0.01, considerably more modes were required for accuracy over the
Re-α domain studied, and due to the considerable computational cost involved these
very low frequencies were not able to be accurately studied.
Also due to the computational expense involved in generating these plots, the
forcing parameters have not been optimised. A set of parameters may therefore exist
that results in complete stabilisation to higher Reynolds numbers using lower forcing
amplitudes. Furthermore, the use of more frequency modes in Equation 7.28 tend to
result in increased stabilisation, and thus even greater stabilisation may result than
that shown in Figure 8.20.

8.3.2 Spanwise forcing


In general, streamwise forcing parameters also stabilised modes with non-zero β when
applied in the spanwise direction. Trends were similar, with frequencies below ap-
proximately 0.04/2π having a stabilising effect and higher frequencies resulting in
destabilisation. Some example growth rates are shown in Figure 8.21 for constant
Reynolds number of 700 and α = 0.25. Due to the much greater computational cost
to solve Equation 7.56, only a few results have been computed.
It can be seen in Figure 8.21 that the effect of the oscillations increases with
increasing forcing amplitude. Hence, low-frequency oscillations which stabilise the
96 CHAPTER 8. STABILITY ANALYSIS RESULTS

0.5

0.45

0.4

0.35
0.01
0.3
α

0.25
0.02

0.025
0.2

0.15

0.1

0.05
500 1000 1500 2000 2500 3000 3500 4000 4500
Reynolds Number

0.5

0.45

0.4

0.35
0.015
0.3
α

0.25 0.03
5
0.03
0.2

0.15

0.1

0.05
500 1000 1500 2000 2500 3000 3500 4000 4500
Reynolds Number

Figure 8.16: Stability curves of Blasius flow for streamwise forcing of amplitude 0.05
(upper), and 0.1 (lower) with Ω = 0.06 using 64 collocation points and 11 modes.
The forcing profile is that shown in Figure 8.14.
8.3. TIME-DEPENDENT TEMPORAL FORMULATION 97

0.5

0.45
0.01
0.4
0.02
0.35
0.03
0.3
α

0.25 0.04

0.2

0.15

0.1

0.05
500 1000 1500 2000 2500 3000 3500 4000 4500
Reynolds Number

0.5

0.45

0.4 0.02
0.0
0.35 1

0.3
α

0.25

0.2

0.15

0.1

0.05
500 1000 1500 2000 2500 3000 3500 4000 4500
Reynolds Number

Figure 8.17: Stability curves of Blasius flow for streamwise forcing of amplitude 0.2
and frequency of 0.4 (upper); amplitude 0.03 and Ω = 0.1 (lower) using 64 collocation
points and 11 modes. The forcing profile is that shown in Figure 8.14 for the upper
plot, and Figure 8.18 for the lower plot.
98 CHAPTER 8. STABILITY ANALYSIS RESULTS

50
real part
45 imaginary part

40
nondimensional height, y

35

30

25

20

15

10

0
−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4
Streamfunction

50

45

40
nondimensional height, y

35

30

25

20

15

10

0
−0.2 0 0.2 0.4 0.6 0.8 1 1.2
nondimensional spanwise velocity, W

Figure 8.18: Forcing profile B. Normalised complex function describing oscillatory


velocity profile (upper) ; resulting velocity profiles over 1 period for a streamwise
oscillation amplitude of 0.2 times freestream velocity. (lower)
8.3. TIME-DEPENDENT TEMPORAL FORMULATION 99

0.5

0.45

0.4

0.35

0.3
α

0.25

0.2

0.15

0.1

0.05
500 1000 1500 2000 2500 3000 3500 4000 4500
Reynolds Number

Figure 8.19: Stability curves of Blasius flow for streamwise forcing of amplitude 0.3
and Ω = 0.04. The forcing profile is that shown in Figure 8.14.

boundary layer have greater effect at higher amplitudes. Upon inspection of the
governing equation (7.56), it can be seen that spanwise oscillations cannot influence
modes of the form (α, 0) where β is zero. However, it can be seen that oscillations
can limit instability to mainly streamwise modes, by stabilising modes with spanwise
wavenumbers.
100 CHAPTER 8. STABILITY ANALYSIS RESULTS

0.5

0.45

0.4

0.35

0.3
α

0.25

0.2

0.15

0.1

0.05
500 1000 1500 2000 2500 3000 3500 4000 4500
Reynolds Number

0.5

0.45

0.4

0.35

0.3
α

0.25

0.2

0.15

0.1

0.05
500 1000 1500 2000 2500 3000 3500 4000 4500
Reynolds Number

Figure 8.20: Stability curves of Blasius flow for streamwise forcing of amplitude 0.2
and Ω = 0.03, using 64 collocation points and (upper) 11 modes, (lower) 21 modes.
The forcing profile is that shown in Figure 8.14. Note that half the number of contour-
points (26-by-26) were used in the bottom plot due to computational time required,
and it therefore has less resolution than the top plot.
8.3. TIME-DEPENDENT TEMPORAL FORMULATION 101

0.018

0.016

0.014

0.012
growthrate

0.01

0.008 A = 0.00
W
Ω = 0.03, A = 0.05
W
0.006 Ω = 0.03, AW = 0.1
Ω = 0.03, AW = 0.2
0.004
Ω = 0.1, AW = 0.05
Ω = 0.1, AW = 0.1
0.002
Ω = 0.1, AW = 0.2
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
β

Figure 8.21: Growth rates of Blasius flow for spanwise forcing at Re = 700 and α =
0.25 using 40 collocation points and 13 modes. Forcing profile A is used (Figure 8.14).
102 CHAPTER 8. STABILITY ANALYSIS RESULTS

8.4 Discussion

8.4.1 Unforced flow results


The results obtained from the temporal and spatial steady-state linearised stability
codes agree very well with other sources. The Matlab code developed is therefore
an excellent tool for computation of the linear stability equations for plane Couette,
plane Poiseuille, and Blasius flows.
It was found that as few as 32 collocation points were required to determine
the least-stable Blasius eigenvalue with reasonable accuracy, but more points were
required to determine the more stable discrete eigenvalues that did not determine the
stability of the flow. The eigenfunctions were more sensitive than the eigenvalues to
the number of collocation points used, however. Although the absolute value of the
eigenfunctions remained much the same, phase throughout the domain was highly
dependent on the number of collocation points. This effect is likely due to the global
nature of the solution method, where all eigenvalues are solved simultaneously. In
contrast to local methods, such as the shooting method, the number of points greatly
affects the eigenfunctions of the continuous spectrum, which in turn may influence
the form of the discrete eigenfunctions.

8.4.2 Forced flow results


The linearised time-dependent stability equation (hereafter LTSE) solved represents
a nonlinear parameter space containing 7 independent variables. Forcing parameters
include the shape of the oscillatory profile, its scaling in y ∗ , frequency, and amplitudes
in the streamwise and spanwise directions. The solution is further affected by the
selection and number of collocation points, and the number of modes used in the
truncated infinite Fourier series for Floquet analysis. Owing to the large scope of the
parameter space, the results given here represent only illustrative cases.
The few cases for Blasius flows studied suggest the boundary layer can be sta-
bilised, destabilised, or relatively unaffected by a given set of forcing parameters.
Owing to large increase in computational cost for non-zero values of β or spanwise
forcing amplitude, cases of streamwise forcing were predominantly studied. The re-
sults of these analyses suggest streamwise forcing near the wall of the Blasius bound-
ary layer (y ∗ ≤ 20) can significantly influence the perturbation dynamics. In general,
forcing frequencies higher than about 0.2/2π were found to have a destabilising effect
on the flow. The amplitude required to significantly influence the flow was typically
above 5% - 10% of the free-stream velocity.
The diversity of the results demonstrate the diversity of solutions to parametric
8.4. DISCUSSION 103

forcing of fluids possible. Similar diversity of solutions has been found for the Faraday
instability (Faraday, 1831; Pozrikidis and Yon, 1998), Taylor-Couette flow subject to
axial oscillations (Marques and Lopez, 1997, 2000), periodically driven cavity flows
(Vogel et al., 2003), and the so-called elliptic instability (e.g. Waleffe, 1990).
The accurate solution of the parametrically forced Mathieu equation using only
a few modes agrees excellently with accepted solutions (McLachlan, 1947; Arscott,
1964), and provides support for the time-dependent stability analysis. Interestingly,
subharmonic instabilities were not found in solutions to the equation. This result is
in agreement with the analysis of Schulze (1999), who concluded fluid systems with
harmonic shearing do not exhibit subharmonic instability. In a numerical analysis
of a fluid bounded by a free surface and an oscillating plate, Or (1997) also noted
subharmonic response was absent.
Several previous studies have found that flow stabilisation is possible from oscil-
lations. Oscillatory pressure introduced to channel flow was found to significantly
increase the critical Reynolds number by von Kerczek (1982), although the pressure
amplitude required made such a technique impractical for application. Selvarajan
et al. (1999) studied the stability of wavy-walled channel flows and found that stabili-
sation could occur when suitable wall-motion was implemented. Murray et al. (1990)
analysed the stability of Taylor-Couette flow subject to axial oscillations, and found
increases in the critical Reynolds number were possible. Such a flow configuration
corresponds to spanwise oscillation in the present case. Further research on para-
metrically forced Taylor-Couette flow by Hu and Kelly (1995), Marques and Lopez
(1997), Marques and Lopez (2000), and Meseguer and Marques (2000) all confirm
that substantial stabilisation is possible via near-wall oscillations.
The author is unaware of previous similar research on the linear stability of a
Blasius boundary layer with distributed streamwise or spanwise forcing, and therefore
direct comparison of results is difficult. However, in light of the excellent agreement
with previous research of the time-independent analyses, potential sources of error
can be limited to four main areas.
The first, and most obvious, potential source of error arises from the truncation of
the infinite Floquet system. Although the stability of the Mathieu equation could be
assessed accurately with only a small number of modes, the far greater complexity of
the LTSE suggests a higher number of modes may be required for accuracy. However,
the size of the Floquet eigenproblem becomes prohibitively large very rapidly as more
modes are used, due to the required inversion of an increasingly large system matrix.
The several test runs performed indicate, however, that fewer modes tend to result
in spurious instability, rather than spurious stable solutions.
Naturally, an adequate number of collocation points is required for an accurate
104 CHAPTER 8. STABILITY ANALYSIS RESULTS

solution. The LTSE analyses were limited to about 64 points again due to compu-
tational limitations. It can be seen from comparison with the eigenspectra of the
steady-state cases that these 64 collocation points are hardly satisfactory. Ideally,
more than 200 points would be used to ensure accuracy. However, this number of
collocation points used with only 21 modes would result in an eigenproblem contain-
ing at least 201 × 21 = 4221 simultaneous equations. Using only 64 points, and 11
modes, the contour plots in Figure 8.20 took over 33 hours to compute using a 3.2GHz
processor with 2GB of RAM.
The third source of error arises from the postprocessing conditions used to remove
spurious eigenvalues which arise from truncation of the infinite Floquet series. The
postprocessing condition of Equation 8.2 is believed to result in worst-case results,
where spurious eigenvalues are difficult to identify.
Finally, it is clear that the linear analysis presented may be inadequate to repre-
sent the dynamics of the problem. Of particular note is the predicted stability for
extremely low frequencies. Obviously, a quasi-steady analysis becomes more appro-
priate in cases where the forcing oscillation is no longer significant relative to the
frequency of perturbations. In such cases, quasi-steady instabilities can grow during
certain phases of oscillation, as pointed out by Murray et al. (1990) and Marques and
Lopez (1997). It is not clear at what frequency the present analysis looses validity.
The stability plot presented for Ω = 0.06 in Figure 8.16 appears reasonable. How-
ever, the plots in the lower plot of Figure 8.20 seem unintuitive given the results of
Figure 8.16, and the highly irregular shape of stability curves casts doubt on their
validity. However, the significance of the broad stabilisation of the Blasius boundary
layer given by this analysis certainly warrants further study.
Curving lines of neutral stability are visible at constant Reynolds number in the
frequency-amplitude domain, ignoring spurious islands of instability resulting from
insufficient modal resolution. Figure 8.13 illustrates these stability curves for a con-
stant Reynolds number of 700. It appears that at low frequency, as the streamwise
forcing amplitude increases, the oscillatory component dominates the stability of the
flow and can result in stabilisation. Such stabilisation is possible even at oscillatory
peak velocities of less than 20% of the free-stream velocity.
Although different time-scales are present in turbulent boundary layers, this stabil-
isation mechanism may be at work in turbulent flows where spanwise oscillations are
implemented. It is suggested that such spanwise oscillations may dominate over the
instability of streak-forming instabilities (which involve spanwise velocities), thereby
stabilising or partially stabilising the streak-formation in the boundary layer. A
stabilisation of a streak-forming instability process would account for the complete
modification of turbulent streak structure observed due to spanwise oscillations (Kar-
8.4. DISCUSSION 105

Figure 8.22: Oblique transition delay found by Berlin (1998) using an oscillating-wall
forcing profile similar to Figure 8.14. Here, ω corresponds to Ω in the present study.
Transition time is shown against forcing frequency (left) and maximum wall velocity
(right).

niadakis and Choi, 2003).


It is clear from the results of the current study that streamwise forcing must be
relatively low frequency and of sufficient amplitude to cause noticeable stabilisation.
At frequencies above about 0.05/2π, the forcing can excite resonant modes in the
boundary layer, as shown in Figure 8.16. Higher frequencies tend to have a general
destabilising effect over the whole Re-α domain, as the upper plot of Figure 8.17
shows.
Spanwise oscillations stabilise modes with non-zero β at frequencies similar to the
frequencies at which streamwise forcing can stabilise the boundary layer. It is noted
that the influence of the oscillations increases with increases β. Thus, only relatively
small amplitude oscillations may be required to stabilise streak-forming (0, 1) modes
of instability. These results can be correlated with experimental results by Berlin
(1998), who studied oblique transition delay by spanwise forcing. Figure 8.14 shows
trends found by Berlin (1998). Note that maximum transition delay was found for
frequencies below 0.09/2π, and increasing amplitude tend to enhance stabilisation for
amplitudes below 0.6.
In the present study, it is observed that stabilisation begins as the frequency
decreases below approximately 0.05/2π, depending on forcing and flow parameters.
A frequency of 0.05/2π corresponds to a non-dimensional forcing period of T ∗ = 126.
It was observed in the present study that stabilisation tended to be increased with
decreasing frequency. However, it is believed low frequency oscillations are unlikely to
stabilise laminar Blasius flow - contrary to the computational results found - because
the flow may become quasi-steady, and a time-independent effects would begin to
106 CHAPTER 8. STABILITY ANALYSIS RESULTS

dominate. Similar conclusions were reached in a similar Floquet analysis by Marques


and Lopez (1997) of axially-forced Taylor-Couette flow. They found the limit Ω → 0
was singular, which is also the case in the present study.
Even so, the results here show stabilisation begins at similar frequencies of os-
cillation that can be correlated to those which induce drag reduction in turbulent
boundary layers due to spanwise oscillations. Optimal forcing periods of between
T + = 50 and T + = 100 were found experimentally by Choi and Graham (1998), Jung
et al. (1992), Quadrio and Ricco (2003)), and Di Cicca et al. (2002) among others,
where T + = T uτ 2 /ν, and uτ is the wall friction velocity. The effectiveness of drag
reduction was found to decay as the oscillatory period, T + , increased above 150. This
contrasts with results of the present study, where stabilisation increases as the time
period tends to infinity, due to the singular nature of the solution as Ω → 0. Choi and
Graham (1998) comments that drag reduction is better scaled with nondimensional
velocity of the oscillatory forcing rather than the frequency of oscillation, which is
in agreement with the present results. Figure 8.21 demonstrates that spanwise os-
cillation amplitude has a relatively linear affect when stabilisation occurs, but the
response of the flow to changes in frequency is highly nonlinear, as can be seen in the
lower plot of Figure 8.15.
It appears, then, from the seemingly broad stabilising influence of appropriate os-
cillation indicated by present results, that such forcing has a stabilising effect on the
instability mechanisms in the boundary layer. It is possible that this phenomenon
may occur in both laminar and turbulent boundary layers. Schoppa and Hussain
(2002) suggested that turbulence regeneration is primarily due to a transient insta-
bility mechanism, and they found appropriate spanwise oscillations have a stabilising
effect on this instability mechanism. A mounting body of evidence suggests the drag
reduction due to spanwise oscillations appears to be related to a fundamental sta-
bilisation of a spanwise instability mechanism, which is supported by the present
results.
Considering practical methods of implementing the present oscillatory forcing,
introduction of an oscillatory streamwise component from a spanwise slot seems fea-
sible. The chosen oscillatory profile was arbitrary, and yet stabilisation was observed.
Thus, perhaps the form of the oscillations need not be exact to result in stabilisa-
tion. Most research using spanwise slots has focussed on wall-normal oscillations, but
this work demonstrates the potential for oscillations in the streamwise direction. The
forcing frequency for stabilisation is relatively low, and it is known that low frequency
modulation of the flow persists longer than high-frequency oscillations downstream
of the actuator (Kim et al., 2003). Thus, oscillations from a single spanwise slot
may have in a stabilising effect over a relatively long streamwise length. This form is
8.5. CONCLUSIONS 107

considerably more difficult to implement, however.


However, the possibility of inducing flow oscillations by static surface deformation
may hold potential. Indeed, this possibility has been suggested by several previous
studies (Handler et al., 1992; Bechert et al., 2000). Sirovich and Karlsson (1997) found
that specialised distributed surface bumps can reduce drag. Though a clear mecha-
nism was not suggested, it is noted that such bumps would induce three-dimensional
oscillatory components into the boundary layer of a similar form to that of the present
study. Provided these oscillations do not excite resonant instabilities, they may be an
effective method of drag reduction. It is also noted that shark riblets are also convex
in shape, which may result in a streamwise oscillatory component. Given the high
hydrodynamic efficiency of sharks (Bechert et al., 1985), it is suggested these oscilla-
tory components may help either delay transition or reduce turbulence regeneration
mechanisms.
Near-wall spanwise oscillations may be induced into the flow by suitable surface
deformation in the form of a spanwise travelling wave. Such forcing has already be
experimentally shown to reduce drag (e.g. Laadhari et al., 1994; Choi and Graham,
1998; Choi and Clayton, 2001), perhaps due to a similar stabilisation mechanism
as that of the present results. However, technical difficulties in the manufacture of
suitable actuation remains an obstacle to cost-effective real-world implementation for
drag reduction.
Although a nonlinear analysis was beyond the scope of the present work, it is likely
that nonlinear effects are important in the stabilisation of boundary layer modes due
to spanwise or streamwise oscillations. At large disturbance magnitude, nonlinear
interaction with the oscillatory shear will undoubtedly result in significant phenom-
ena, possibly resulting in stabilisation. Indeed, given the unintuitive stabilising effect
of oscillations in this linear analysis, it seems likely that oscillations will also have a
significant effect on nonlinear, transient mechanisms thought to be involved in tur-
bulence regeneration.

8.5 Conclusions
An efficient and useful Matlab code has been developed to solve the linearised stability
equations. Matching results for the steady state cases validate the basic code. Using
the modular code, plots of eigenvalues, plots of eigenfunctions, and stability curves
can be readily generated for Couette, Poiseuille, and Blasius boundary layers. Each
module of the code can be changed independently, and a module for Falkner-Scan-
Cooke boundary layers could easily be added.
Stability results for the Blasius flow modulated by streamwise or spanwise oscil-
108 CHAPTER 8. STABILITY ANALYSIS RESULTS

latory forcing are extremely diverse and nonlinear. While these results for oscillatory
forcing are somewhat preliminary, they indicate the Blasius boundary layer can be
stabilised to significantly higher Reynolds numbers through appropriate streamwise
forcing. Using an arbitrary forcing profile, stabilisation occurred for low frequency
oscillations using a maximum oscillatory velocity of greater than 20% free-stream
velocity. However, it is noted that the linear time-dependent analysis presented here
may not remain valid at these low frequencies for laminar flows, where quasi-steady
analysis may be more applicable. Smaller forcing amplitudes and higher forcing fre-
quencies both resulted in destabilisation of the Blasius flow. It was also found that
some combinations of parameters have little effect on the growth rate of disturbances.
The results suggest that streamwise oscillations can be used to delay transition
to turbulence in Blasius flows. The relatively broad influence of the arbitrary forc-
ing profile indicates a relatively general stabilisation mechanisms may be at work.
Physically, it appears that at sufficient forcing amplitudes, the oscillatory component
dominates over the instability of the mean flow, thereby resulting in stable growth
rates of disturbance. If this is the case, it is suggested that this mechanism may also be
effective in turbulent boundary layers. Particularly of note is the drag-reducing effect
of spanwise oscillations on the turbulence regeneration mechanisms, which has been
confirmed experimentally. It is suggested that this drag-reduction stems from stabil-
isation of (transient) growth mechanisms of streamwise coherent structures through
a generic stabilisation mechanism similar to that of the present study.
Practically, the results indicate the potential of future research into streamwise
and spanwise oscillatory boundary layer control. Streamwise oscillation may be intro-
duced through a spanwise slot of appropriate design, and spanwise oscillations may
be achieved through a spanwise travelling wave of surface deformation.

8.6 Future work


The lack of computational power has limited the present study to a minimum num-
ber of modes for the time-dependent case. It is therefore desirable to perform more
accurate calculations of oscillatory forcing using a greater number of modes and col-
location points. Further research would also allow a greater area of the parameter
space to be analysed, finding more optimum forcing parameters.
Recent years have seen the development of the so-called Parabolised Stability
Equations (PSE) (Schmid and Henningson, 2001). Although a time-dependent for-
mulation of the PSE would be challenging, results may be more accurate than those
presented here.
Experimental validation of the results is also required. The stabilisation predicted
8.6. FUTURE WORK 109

by the time-dependent results appear unintuitive and therefore questionable for low-
frequency oscillations. However, in turbulent flows, where shorter time-scales exist,
an analysis similar to this one may provide evidence of stabilisation in turbulent
flows via a similar mechanism to that presented here. Experimental testing, though
difficult in practise for the streamwise forcing profile studied herein, would also give
understanding of the physical reasons for the stabilisation observed herein.
The effects of oscillatory forcing on transient growth in the boundary layer may
also be of interest, particularly because transient mechanisms are thought to be domi-
nant in turbulence regeneration. It is not clear how results of the present study would
relate to secondary instability of streaks. Indeed, the secondary instability of stream-
wise streaks subject to spanwise oscillations would be of great interest, although
mathematically and computationally formidable.
110 CHAPTER 8. STABILITY ANALYSIS RESULTS
Part IV

Actuation method

111
Chapter 9

Analytical solution

Three methods exist to provide the suitable near-wall forcing of the near-wall bound-
ary layer, wall oscillation, Lorentz forcing and travelling-wave wall actuation. These
methods are shown schematically in Figure 1.1. The first of these methods, wall os-
cillation, has been shown to reduce drag, but requires large amplitudes for effect. The
last two of these methods can provide a travelling wave of the form of Equation 6.1.
In this work, moving wall actuation has been chosen for investigation because it is
more generally applicable to fluids, as Lorentz forcing requires electrolytic fluid and
the generation of relatively strong electromagnetic fields.
Work by Vlachos and Rediniotis (2003) suggests that wall motion may prove
efficient. Because relatively small vertical wall motions induce motions in the fluid
of comparable magnitude, the only limiting factor is the actuation method for the
moving wall. Their results suggest large drag reductions (up to 70%) are possible,
and proof of concept work for a shape-memory-alloy (SMA) based deformable skin
has been initiated.
It is likely that shape-memory-alloy technology will be appropriate for travelling-
wave forcing (Vlachos and Rediniotis, 2003). Because the technology is still relatively
new, considerable further development will be required until a practical and robust
actuator can be manufactured. However, a detailed understanding of the flow field
generated by such forcing is essential for design and prediction of drag reduction.
Thus, the flow field arising from travelling-wave wall motion was investigated numer-
ically using a finite element computational scheme and verified by physical experi-
mentation.

9.1 Problem specification


The flow field is periodic in the spanwise domain, with the same wavelength as the
forcing wave. Accordingly, the problem was non-dimensionalised to a single wave-

112
9.1. PROBLEM SPECIFICATION 113

length, of physical length λ. The solution is expected to be similar to Stoke’s oscillat-


ing plate solutions, which are known to be relatively stable (Kerczek and Davis, 1974)
and therefore the flow is assumed to be laminar. The flow is characterised by the fluid
density, viscosity, wavelength, wave-speed, and wave amplitude (ρ, µ, λ, Uw , and āw
respectively). Thus, letting:

ū v̄ x̄ ȳ Ā p̄ t̄ λ
u= ; v= ; x= ; y= A= ; p= ; t= ; T =
Uw Uw λ λ λ ρUw 2 T Uw
(9.1)

λUw
Re = (9.2)
ν

The problem can be specified as the flow over a wall moving in the vertical direction
with displacement given by:

y = Aei(ax+Ωt) (9.3)

where 2π/a is the wavelength λ of the surface displacement and 2π/Ω is the
frequency. A mapping is therefore required to transform the curved physical domain
to the straight computational domain. McLean (1983) used a conformal mapping for
a numerical simulation over a wavy wall, while Shen et al. (2003) used an algebraic
method. In a three-dimensional simulation, De Angelis et al. (1997) used a non-
orthogonal mapping. In the present case, it was found a non-orthogonal exponential
mapping was sufficient, of the form

x
e=x
(9.4)
ye = y + Aei(ax+Ωt) e−cy

where c = 5, (e
x, ye) are coordinates in the physical space and (x, y) are coordinates
in the computational domain. Figure 9.1 shows this mapping for the case of c = 5.
Higher values cause the effect of the displacement to decay faster with y, while lower
values propagate the displacement of the surface further away from the wall (a lower
decay rate). In all cases considered A is small and therefore terms due to the non-
orthogonality of the transformation can be omitted in the analysis. Only laminar
flow is considered in the analytical formulation.
114 CHAPTER 9. ANALYTICAL SOLUTION

0.5 0.5
nondimensional height, y

nondimensional height, y
0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0

0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1


nondimensional spanwise distance, z nondimensional spanwise distance, z

Figure 9.1: Domain mapping of horizontal lines in the computational domain to the
curved lines in the physical domain with c = 5 (left) and c = 3 (right). Note the
difference near the top of the plots.

9.2 Nonlinear formulation


The non-dimensionalised 2-D Navier Stokes equations are:

1 ∂2u ∂2u
 
∂u ∂u ∂u ∂p
+u +v =− + + (9.5a)
∂t ∂x ∂y ∂x Re ∂x2 ∂y
1 ∂ v ∂2v
 2 
∂v ∂v ∂v ∂p
+u +v =− + + (9.5b)
∂t ∂x ∂y ∂y Re ∂x2 ∂y 2

A streamfunction can be substituted:

∂ψ ∂ψ
u= ;v = − (9.6)
∂y ∂x

According to Floquet theorem, because the forcing is sinusoidal in x and t, the solu-
tion, ψ, can be represented as a Fourier series where each term is of the same period
as the forcing.

X
ψ= fk eik(ax+Ωt) (9.7)
k

Using a Chebyshev collocation technique, differentials in y can also be represented by


Chebyshev operators (see Section 7.3). Equations 9.5 can then be written:
9.2. NONLINEAR FORMULATION 115

∂2
   
∂ ∂ ∂ 2 ∂ 1 3
D ψ + Dψ D ψ− ψD ψ = − p + D 2ψ + D ψ (9.8a)
∂t ∂x ∂x ∂x Re ∂x
∂2 ∂3
   
∂ ∂ ∂ ∂ 1 2 ∂
− ψ − Dψ 2 ψ + ψ D ψ = −Dp + − 3ψ − D ψ (9.8b)
∂x ∂t ∂x ∂x ∂x Re ∂x ∂x

Substituting Equation 9.7 into the Navier-Stokes equations (9.8) yields:

∂ 1
ikΩDψ + ika (Dψ)2 − ika (ψ) D2 ψ = − p + −k 2 a2 Dψ + D3 ψ
 
(9.9a)
∂x Re
1
k 2 aΩψ = −Dp + ik 3 a3 ψ − ikaD2 ψ

(9.9b)
Re

The pressure distribution can be represented by a Fourier series just as the stream-
function, ψ, was represented:

X
p= pk eik(ax+Ωt) (9.10)
k

Substituting Equation 9.10 into Equations 9.9, the first equation of 9.9 can be mul-
tiplied by D and the second equation of 9.9 can be multiplied by iak.

1
ikΩD2 ψ + ikaD (Dψ)2 − ikaD (ψ) D2 ψ − −k 2 a2 D2 ψ + D4 ψ = −ikaDp
 
Re
(9.11a)
1
ik 3 a2 Ωψ + k 4 a4 ψ − k 2 a2 D2 ψ = −ikaDp

Re
(9.11b)

Rearranging Equation 9.11 to eliminate pressure yields:

i 2
ΩD2 ψ + aD (Dψ)2 − aD (ψ) D2 ψ − k 2 a2 Ωψ + D 2 − k 2 a2 ψ = 0

(9.12)
k Re
116 CHAPTER 9. ANALYTICAL SOLUTION

The Fourier series representing the streamfunction given by Equation 9.7 can be
substituted into Equation 9.12. Balancing terms that have the same time-dependance
(terms that all contain eCΩt where C ∈ Z) yields a coupled nonlinear simultaneous
equation.

M
X M
X
ΩD2 fk + aD fm D2 fk−m

(Dfm ) (Dfk−m ) − aD
m=−M m=−M
i 2
− k 2 a2 Ωfk + D2 − k 2 a2 fk = 0 (9.13)
k Re

In the case of wave-like wall motion considered here, the boundary condition is defined:

A  i(ax+Ωt)
+ e−i(ax+Ωt)

v= e (9.14)
2

The boundary condition of Equation 9.14 leads to all but the following boundary
conditions of Equation 9.12 being set to zero :

A
f10 (0) = (9.15a)
2
0 A
f−1 (0) = (9.15b)
2

Although Equation 9.12 is an accurate representation of the flow-field, the solution


of this equation is difficult. An iterative method would be required, which may not
converge well to the correct solution because the solution may not be convex. Further,
the large number of unknowns may require a large time for iterative solution. Rather
than solving Equation 9.12, an approximated governing equation effectively using
two Fourier modes was derived. This approximated solution did not require iterative
methods, and was sufficiently accurate for low-amplitude travelling wave forcing.
9.3. DERIVATION OF APPROXIMATED GOVERNING EQUATION 117

9.3 Derivation of approximated governing equa-


tion
A nonlinear differential equation governing the flow field over the moving wall will
now be presented. Consider the non-dimensionalised two-dimensional Navier Stokes
equations:

1 ∂2u ∂2u
 
∂u ∂u ∂u ∂p
+u +v =− + + (9.16a)
∂t ∂x ∂y ∂x Re ∂x2 ∂y
1 ∂ v ∂2v
 2 
∂v ∂v ∂v ∂p
+u +v =− + + (9.16b)
∂t ∂x ∂y ∂y Re ∂x2 ∂y 2

The derivation will involve the assumption that the form of the solutions are approx-
imately sinusoidal. Following this, let a complex stream-function, which represents
the oscillatory component of the flow, be defined by

ψ = f (x, y, t) ei(ax+Ωt) (9.17)

where only the real parts are considered. Numerical results given in section 11 confirm
independence of the oscillatory and steady-state component (which increases with
time). Only the oscillatory components will be considered in the present analysis.
The function f can be complex to allow for phase shifting. In the computational
domain, let the flow components v and u be defined by the complex stream function:

∂ψ ∂ψ
u= ;v = − (9.18)
∂y ∂x

Upon substitution of the equations for the velocities u and v in 9.18, the Navier-Stokes
equations (9.16) can be written:

∂2ψ ∂ψ ∂ 2 ψ ∂ψ ∂ 2 ψ
 3
∂3ψ

∂p 1 ∂ ψ
+ − =− + + 3 (9.19a)
∂y∂t ∂y ∂x∂y ∂x ∂y 2 ∂x Re ∂x2 ∂y ∂y
2 2 2
 3
∂3ψ

∂ ψ ∂ψ ∂ ψ ∂ψ ∂ ψ ∂p 1 ∂ ψ
− − + =− + − 3 − 2 (9.19b)
∂x∂t ∂y ∂x2 ∂x ∂y∂x ∂y Re ∂x ∂y ∂x
118 CHAPTER 9. ANALYTICAL SOLUTION

Because the streamfunction ψ is assumed to be approximately sinusoidal, having


the form of equation 9.17, equations for the partial derivatives of u and v can be
derived. For example, the spanwise velocity u can be written as

∂ψ ∂   ∂f (x, y, t) i(ax+Ωt)
u= = f (x, y, t) ei(ax+Ωt) = e (9.20)
∂y ∂y ∂y

u = f 0 ei(ax+Ωt) (9.21)

where a prime denotes differentiation with respect to y. Similarly for vertical velocity
v,

∂ei(ax+Ωt) ∂f (x, y, t) i(ax+Ωt)


 
∂ψ ∂  i(ax+Ωt)

v=− =− f (x, y, t) e = − f (x, y, t) + e
∂x ∂x ∂x ∂x
(9.22)

i(ax+Ωt)
It will be assumed that ∂e ∂x  ∂f (x,y,t)
∂x
and thus ∂f (x,y,t)
∂x
is negligible. This
assumption can be justified because of the sinusoidal nature of the streamfunction
with x, which is modulated only slightly over the x domain. Numerical solutions
(given in section 11) confirm the magnitude of the modulation is considerably smaller
than the amplitude of the sinusoid for small amplitudes. This assumption will result
in an approximated solution, where velocity components are purely sinusoidal in x
and t. Similarly, higher partial differentials in the x direction of f (x, y, t) will be
neglected. Thus:

∂ei(ax+Ωt)
v = −f (x, y, t) = −iaf ei(ax+Ωt) (9.23)
∂x

As with the vertical velocity calculation in Equation 9.22, in all cases of partial
differentiation with respect to the x direction the differential of f will be assumed
to be much smaller than the differential of ei(ax+Ωt) , and thus negligible. Similarly,
differentiation with respect to time of f will be assumed negligible to the differential
of ei(ax+Ωt) . With these assumptions, the following functions of f for the partial
differentials of the streamfunction are substituted into the Navier-Stokes equations
(9.16):
9.3. DERIVATION OF APPROXIMATED GOVERNING EQUATION 119

∂ψ
= f 0 ei(ax+Ωt) (9.24a)
∂y
∂ψ
= iaf ei(ax+Ωt) (9.24b)
∂x
∂2ψ ∂ 
iΩf ei(ax+Ωt) = iΩf 0 ei(ax+Ωt)

= (9.24c)
∂y∂t ∂y
∂2ψ ∂ 
iΩf ei(ax+Ωt) = −aΩf ei(ax+Ωt)

= (9.24d)
∂x∂t ∂x
∂2ψ ∂2ψ ∂  0 i(ax+Ωt) 
= = fe = iaf 0 ei(ax+Ωt) (9.24e)
∂x∂y ∂y∂x ∂x
∂2ψ ∂ 2  i(ax+Ωt) 
= fe = f 00 ei(ax+Ωt) (9.24f)
∂y 2 ∂y 2

∂2ψ ∂ 2  i(ax+Ωt) 
= fe = −a2 f ei(ax+Ωt) (9.24g)
∂x2 ∂x 2

∂3ψ ∂ 3  i(ax+Ωt) 
= fe = f 000 ei(ax+Ωt) (9.24h)
∂y 3 ∂y 3
∂3ψ ∂ 3  i(ax+Ωt) 
= fe = −ia3 f ei(ax+Ωt) (9.24i)
∂x3 ∂x3
∂3ψ ∂ 2  0 i(ax+Ωt) 
= fe = −a2 f 0 ei(ax+Ωt) (9.24j)
∂x2 ∂y ∂x2
∂3ψ ∂2  i(ax+Ωt)
= iaf 00 ei(ax+Ωt)

= iaf e (9.24k)
∂y 2 ∂x ∂y 2

Numerical simulations (in section 11) confirm the pressure decays approximately
exponentially (see Figure 9.3).

p = P0 ei(ax+Ωt+θ) e−κy (9.25)

where κ is the rate of pressure decay in the y direction. The governing equations
(9.19) therefore become

2 1  2 0
iΩf 0 + ia (f 0 ) ei(ax+Ωt) − iaf f 00 ei(ax+Ωt) = −iaP0 eiθ e−κy + −a f + f 000 (9.26a)

Re
1
aΩf + a2 f f 0 ei(ax+Ωt) − a2 f f 0 ei(ax+Ωt) = κP0 eiθ e−κy + ia f − iaf 00
 3 
(9.26b)
Re

Noting the second and third terms in the second equation cancel, and multiplying
120 CHAPTER 9. ANALYTICAL SOLUTION

Log RMS pressure with height

10e2
0.5 10e3

nondimensional height, y
10e4
10e5
0.4

0.3

0.2

0.1

0
−4.5 −4 −3.5 −3 −2.5 −2 −1.5 −1
log nondimensional RMS pressure

Figure 9.2: Numerical simulation results showing the natural logarithm of pressure
with height for different Reynolds numbers with a nondimensional amplitude of 0.05.
Note the similarity in decay rate (gradient). Decay rates converge with decreasing
amplitudes.

the 2nd equation by ia/κ, the equation becomes:

ia2 a2  2
− iaP0 eiθ e−κy = − a f − f 00

Ωf − (9.27)
κ κRe

Substituting this pressure term into the first equation (9.26a) gives:

h
2
i ia2 a2  2 1  2 0
− ia (f 0 ) − f f 00 ei(ax+Ωt) = iΩf 0 + a f − f 00 − −a f + f 000
 
Ωf +
κ κRe Re
(9.28)

Now, defining the ratio s = a/Ω, the governing equation can be rewritten as

h
2
i a2 isa  2 i  2 0
s (f 0 ) − f f 00 ei(ax+Ωt) = −f 0 − f + a f − f 00 − −a f + f 000 (9.29)
 
κ κRe aRe

h
0 2 00
i i  3
ei(ax+Ωt) = −κf 0 − a2 f + sa f − saf 00 + af 0 − a−1 f 000 (9.30)

sκ (f ) − f f
Re
9.3. DERIVATION OF APPROXIMATED GOVERNING EQUATION 121

h i
2
f 000 = iasκRe (f 0 ) − f f 00 ei(ax+Ωt) − sa2 f 00 + iκaRe + a2 f 0 + ia3 Re + sa4 f
 

(9.31)

The boundary conditions on f (x, y, t) are:

f = f 0 = f 00 = f 000 = 0 as y →∞
u = f 0 ei(ax+Ωt) = 0 ∴ f 0 = 0 at y =0 (9.32)
A
v = −iaf ei(ax+Ωt) = Aei(ax+Ωt) ∴f =− at y =0
ia

The governing Equation 9.31 was solved numerically using a Matlab (v6.5) dif-
ferential boundary value problem solver, which uses a Chebyshev collocation method
to achieve spectral accuracy of f and f 0 . However, it should be noted that it is
often desirable to make a = 1 for the stability of the solution. In this case, the non-
dimensionalised distances are multiplied by 2π and the non-dimensional velocities are
also accordingly multiplied by 2π, thus making a and Ω equal to 1. The solution is
then calculated between 0 ≤ x ≤ 2π, 0 ≤ y ≤ 2π. At high Reynolds number, it is
desirable to lower s below unity by multiplying Ω by a constant C. The increases the
effective Reynolds number also by a factor of C, thereby requiring the equation to be
solved for an adjusted Reynolds number of ReC = Re/C in order to give the actual
solution for a given Re.
The differentials of f (x, y, t) are all in the vertical direction, and thus f (x, y, t)
can be evaluated as a complex differential equation in y for given x and t. Numerical
simulations confirm that κ is a primarily a function of wavelength, with contribution
from amplitude and playing a minor role the vertical decay of pressure.


κ≈1 (9.33)
a

It should be noted that the pressure field given in equation 9.27 was chosen without
loss of generality. So long as the pressure field decays exponentially with height, the
solution does not assume a form of the pressure field with respect to x. The actual
pressure field may be composed of a Fourier series in x, but all these terms will be
eliminated from the equation in the substitution step of Equation 9.28. Once the
streamfunction ψ is determined, the pressure field can be calculated by substitution
of ψ into the Navier-Stokes equations.
122 CHAPTER 9. ANALYTICAL SOLUTION

9.4 Stokes oscillating plate solution


By simply changing the boundary conditions of Equation 9.32 to those shown in
Equation 9.34, a more advanced solution to the stokes oscillating plate problem can
be found. In this case, the wall oscillation velocity may vary sinusoidally over the
spanwise domain with wavelength 2π a
. In this case, an oscillating vertical velocity is
observed which is similar in form to the solution of the spanwise velocity induced by
the vertically oscillating boundary. It is a trivial matter of letting the wavelength
tend to infinity to derive the standard Stokes oscillating plate solution.

f = f 0 = f 00 = f 000 = 0 as y → ∞
u = f 0 ei(ax+Ωt) = Aei(ax+Ωt) ∴ f 0 = Astokes at y = 0 (9.34)
v = −iaf ei(ax+Ωt) = 0 ∴ f = 0 at y = 0

In the case of large wavelength, a becomes very small, s = a/Ω also becomes
small, and κ also becomes very small. If a → 0 , all these variables are of the same
order as a. Upon inspection of Equation 9.31 it is clear that the lowest order terms
(2nd order) will dominate the right hand side of the equation. Thus, ignoring terms of
greater than second order, making use of relation 9.33, and without loss of generality
setting a = 1 the governing equation (9.31) becomes:

f 000 = (1 + iWo) f 0 (9.35)

Where Wo is the Womersly number. For Re  1, the term involving the Womersly
number dominates and the equation becomes f 000 = iWof 0 . Integration with respect
to y yields:

f 00 − iWof = B (9.36)

and from the boundary conditions (9.34) it can be seen that B = 0. Therefore
Equation 9.36 becomes the Stokes solution (Schlichting, 1960). Thus, it can be seen
that seen that the governing equation of 9.31 is a more generalised Stokes equation
which allows for not only spanwise surface oscillation, but also for small-amplitude
vertical surface oscillation.
Chapter 10

2-Dimensional numerical
simulation

10.1 Theory
10.1.1 Notation of tensors and derivatives
In this section, three forms of mathematical notation for equations that involve tensors
and/or their derivatives are briefly reviewed as reference for the theory in following
sections. The three forms are as follows:

(i) Standard tensor notation

(ii) Spatial derivative shorthand notation

(iii) Temporal derivative notation

Standard tensor notation uses subscripted indices to represent the order and com-
ponents of a tensor. Spatial derivative shorthand notation uses a subscripted comma
(,) to represent first-order spatial derivatives. Temporal derivative notation employs
either a standard derivative operator or a dot () to represent a derivative with respect
to time.

Standard tensor notation

In standard tensor notation, subscripted indices are used to identify the order and
components of a tensor. For example, the notation αij represents the elements of a
second-order tensor of dimension two. If i, j = 1, 2 the tensor has the form

" #
α11 α12
α21 α22

123
124 CHAPTER 10. 2-DIMENSIONAL NUMERICAL SIMULATION

Variables with repeated indices, such as αii , denote the summation over the range
of the repeated index followed by a reduction in the order of the tensor. The reduction
in the order of the tensor is equal to the number of repeated indices. For example,
αii represents a scalar quantity (zero-order tensor) that is the sum of the diagonal
elements of tensor αij . That is, for i = 1, 2:

αii = α11 + α22

Spatial derivative shorthand notation

In spatial derivative shorthand notation, the combination of a subscripted comma and


index - for example, (, i) - represents the first derivative with respect to the index.
In this manual, the subscripted index i typically represents the coordinate direction
xi . Therefore, for any variable, α, the notation α,i represents a first-order tensor the
components of which are of the form

∂α
∂xi

for i = 1, 2 for a two-dimensional problem. For example, in a two-dimensional flow


problem:

h i
∂α ∂α
α,j = ∂x1 ∂x2

If a tensor subscript contains two distinct indices separated by a comma - for


example (i, j) - the notation represents a second-order tensor each element of which
constitutes a first-order derivative with respect to direction xj . For example, for a
two-dimensional problem (i, j = 1, 2),
" #
∂u1 ∂u1
∂x1 ∂x2
ui,j = ∂u2 ∂u2
∂x1 ∂x2

If the subscripted indices are identical to each other - for example, (i, i) or (j, j)
- the notation represents a scalar quantity that constitutes the sum of the diagonal
elements.
10.1. THEORY 125

Temporal derivative notation

The notation for temporal derivatives employs either a derivative operator or a dot
() to represent the derivative with respect to time. For example, the first derivative
with respect to time of the variable y is represented in either of the following forms:

∂y
or ẏ
∂t

10.1.2 Conservation equations


The basic transport equations that govern the flow of a viscous fluid are mathematical
representations of conservation principles. Specifically, they represent the conserva-
tion of three physical quantities:

(i) Mass

(ii) Momentum

(iii) Energy

The following sections describe the basic transport equations as they apply to the
present analysis.

10.1.3 Mass conservation


The principle of mass conservation applied to a flowing fluid results in the equation

∂ρ
+ (ρuj ),j = 0 (10.1)
∂t

where j = 1, 2 for two-dimensional or axi-symmetric flows, j = 1, 2, 3 for three dimen-


sional flows, ρ is density, and uj is velocity. For the constant-density, incompressible
fluid assumed for the present analysis, Equation 10.1 reduces to the constraint:

uj,j = 0 (10.2)

which, for a two-dimensional flow problem, represents the equation


126 CHAPTER 10. 2-DIMENSIONAL NUMERICAL SIMULATION

∂u1 ∂u2
+ =0
∂x1 ∂x2

or, in a Cartesian reference frame,

∂u ∂v
+ =0
∂x ∂y

10.1.4 Momentum conservation


The principle of conservation of linear momentum dictates that, for any fluid element,
 
∂ui
ρ + uj ui,j = σij,j + ρfi (10.3)
∂t

where i, j = 1, 2 for two-dimensional or axi-symmetric flows, σij is the stress tensor,


and fi is the body force per unit mass. In most cases, ρfi represents the force due
to gravity, which is ignored in the present analysis. Assuming incompressibility and
constant viscosity, Equation 10.3 can also be written in Navier-Stokes form as:
 
∂ui
ρ + uj ui,j = −p,i + µ [ui,j ],j + ρfi (10.4)
∂t

The Navier-Stokes form was used for the present analysis.

10.1.5 Thermal energy conservation


The present analysis was assumed to be at constant temperature and thus the thermal
energy equation was not solved.

10.1.6 Formulation of the discrete problem


The continuum fluid region was divided into 4-noded quadrilateral elements. Within
each element, the dependent variables of velocity and pressure, ui and p, are interpo-
lated by functions of compatible order as given by:

ui (x, t) = ϕT Ui (t)
(10.5)
p (x, t) = ψ T P (t)
10.1. THEORY 127

where Ui and P are column vectors of element nodal point unknowns and ϕ and
ψ are column vectors of the interpolation functions. The same basis functions are
employed for all components of the velocity to minimise computational cost. Sub-
stitution of these approximations into the field equations and boundary conditions
yields equations for momentum and mass conservation as:

f1 (ϕ, ψ, Ui , P ) = R1
(10.6)
f1 (ϕ, Ui ) = R2

where R1 and R2 are the residuals resulting from the use of the approximations of
Equation 10.1.
The Galerkin form of the Method of Weighted Residuals was used to reduce these
errors to zero, by making the residuals orthogonal to the interpolation functions of
each element (that is, ϕ and ψ). These orthogonality conditions are expressed by:

(f1 , ϕ) = (R1 , ϕ) = 0
(10.7)
(f2 , ψ) = (R2 , ψ) = 0

R
where (a, b) denotes the inner product, defined by (a, b) = V
a · b dV , and V is the
volume of the element.

10.1.7 Derivation of matrix coefficients


Using the definition of the Galerkin procedure, Equation 10.4, and the finite element
approximations (10.1), the following integral equations can be written:

∂ϕT
Z  Z 
T dUi
ρ0 ϕϕ dV + ρ0 ϕuj dV Ui
V dt V ∂xj
∂ϕ ∂ϕT
Z  Z 
∂ϕ T
− ψ dV P + µ dV Ui
V ∂xj V ∂xi ∂xj
∂ϕ ∂ϕT
Z  Z Z
+ µ dV Uj = σi ϕdS + ρ0 fi ϕdV (10.8)
V ∂xj ∂xj S V

∂ϕT
Z 
ψ dV Ui = 0 (10.9)
V ∂xi
128 CHAPTER 10. 2-DIMENSIONAL NUMERICAL SIMULATION

In deriving the above equations, the Green-Gauss theorem was used to reduce
the second-order diffusion terms in the momentum and energy equations and the
pressure term to first-order terms plus a surface integral. The appearance of the
surface integrals containing the applied surface stresses corresponds to the ”natural”
boundary conditions for the problem.
Once the form of the interpolation functions ϕ and ψ are specified, the integrals
defined in equations 5.3.4 and 5.3.5 may be evaluated to produce the required coeffi-
cient matrices. Combining the momentum and energy equations into a single matrix
equation for two-dimensional, isothermal incompressible problems produces a system
of the form:

     
M 0 0 U̇1 2K11 + K22 K12 −C1 U1
 0 M 0 U̇2  +  K21 K11 + K22 −C2  U2 
     

0 0 0 Ṗ −C1T −C2T 0 P
    
A1 (U1 ) + A2 (U2 ) 0 0 U1 F1
+ 0 A1 (U1 ) + A2 (U2 ) 0 U2  = F2  (10.10)
    

0 0 0 P 0

where:

Z
M= ρ0 ϕϕT dV
ZV
Mp = ρ0 ψψ T dV
ZV
∂ϕ ∂ϕT
Kij = µ dV
V ∂xj ∂xi
Z (10.12)
∂ϕ T
Ci = ψ dV
V ∂xi
∂ϕT
Z
Ai (Uj ) = ρ0 ϕuj dV
V ∂xi
Z Z
Fi = σi ϕ dS + ρ0 fi ϕ dV
S V

h i h i
with U = U1 U2 and setting V = U1 U2 P , these may be combined into the
single matrix equation,

" #" # " #" # " #


M 0 U̇ A (U ) + K −C U F
+ = (10.13)
0 0 Ṗ −C T 0 P 0
10.2. SOLUTION METHOD 129

The matrix equation above represents the discrete analogue of the equations of
motion for an individual fluid element. Equations for each element were combined
using the direct stiffness approach to give a system of matrix equations of the form:

M V̇ + K (U ) V = F (10.14)

10.2 Solution method


10.2.1 Non-dimensionalisation
The problem was again non-dimensionalised to a single wavelength, and laminar flow
was assumed as in the analytical approximation. The flow is characterised by the fluid
density, viscosity, wavelength, wave-speed, and wave amplitude (ρ, µ, λ, Uw , and Ā
respectively). Thus, letting:

ū v̄ x̄ ȳ Ā p̄ t̄ λ
u= ; v= ; x= ; y= A= ; p= ; t= ; T =
Uw Uw λ λ λ ρUw 2 T Uw
(10.15)

Equation 10.4 then becomes:

∂ui 1
+ uj ui,j = −p,i + [ui,j ],j + fi (10.16)
∂t Re

where

λUw
Re = (10.17)
ν

10.2.2 Domain and boundary conditions


The two-dimensional, non-dimensionalised domain of 1 unit used contained 722 linear,
quadrilateral elements, as shown in Figure 10.1 (except for cases of Re = 10, where
the domain was extended to a non-dimensional height of 2). The top boundary
130 CHAPTER 10. 2-DIMENSIONAL NUMERICAL SIMULATION

Figure 10.1: Computational mesh for numerical simulation.

was modelled as a plate, while no-slip was enforced on the lower, travelling wave
interface. Only one wavelength was modelling, and degrees of freedom at the left and
right boundaries were set to be equal to create a spatially periodic system.
In order to eliminate any effects due to initial conditions, the amplitude of the interface
displacement was varied according to Equation 10.18:

A = A0 1 − exp −t3
 
(10.18)

where t is non-dimensional time, and A0 is the asymptotic amplitude. This smooth


increase of amplitude, shown in Figure 10.2, was required to eliminate spurious pres-
sures. A MATLAB code was used to generate the script read-file for FIDAP, au-
tomating the entry of the periodic boundary conditions unique for each node. Grid
independence was confirmed using a 482 grid. The height-to-wavelength aspect ra-
tio of unity was confirmed accurate below y = 0.5 for all except very low Reynolds
number. These errors are discussed in Section 11.3 of this work.

10.2.3 Numerical scheme


A segregated finite element solver was used, using diagonal preconditioning and the
conjugate gradient squared (CGS) solution method derived from biconjugate gradient
method by Sonneveld et al. (1985) to deal with non-symmetric, non-positive real
systems of equations. A pressure projection method was used in determination of
10.2. SOLUTION METHOD 131

0.9

0.8

0.7

0.6

0
A/A
0.5

0.4

0.3

0.2

0.1

0
0 0.5 1 1.5 2 2.5 3
time

Figure 10.2: Smooth ramp-up of amplitude used in simulations, given by Equa-


tion 10.18.

the transient pressure field. A time step of 0.01 seconds with backward integration
was used over the time domain, which typically was greater than 10 periods to allow
the flow field to reach steady state. Streamlined upwinding was combined with a
hybrid relaxation scheme, where relaxation was adjusted as a function of residuals to
minimise computational cost.
To allow for distortion of the flow domain, a Lagrangian pseudo-elastostatic remesh-
ing scheme was employed. Routines programmed in the fluid-structure-interaction
component of the FIDAP code minimised element deformation.
Chapter 11

Actuation results

The results of both analytical and numerical solutions are presented in this chapter.
Section 11.1 presents the solution to Equation 9.31 with the Stokes oscillating-plate
boundary conditions of Equations 9.34. Section 11.2 presents a discussion of results
for travelling wave forcing over a range of Reynolds numbers and wave amplitudes,
where Equation 9.31 was solved with the boundary conditions given in Equation 9.32.

11.1 Stokes oscillating plate solution

Velocity profiles in y over one cycle


1

0.9

0.8

0.7
nondimensional height, y

0.6

0.5

0.4

0.3

0.2

0.1

0
−1 −0.5 0 0.5 1
nondimensional spanwise velocity, W

Figure 11.1: Stokes oscillating-plate


√ solution obtained using the boundary conditions
in Equations 9.34 at Wo = 10 with the general governing Equation 9.31.

The Womersly number defines the Stokes oscillating plate solution. It is defined:

132
11.2. TRAVELLING-WAVE FORCING 133

Streamfunction profile in y over one cycle


1
real part
0.9 imaginary part

0.8

0.7

nondimensional height, y
0.6

0.5

0.4

0.3

0.2

0.1

0
−0.02 0 0.02 0.04 0.06 0.08 0.1 0.12
Streamfunction

Figure 11.2: Stream √


function of Equation 9.34 for a Stokes oscillating plate at a
Womersly number of 10.

s
ρf
Ωn = A (11.1)
µ

where f is the oscillation frequency and A is the amplitude. Figure 11.1 shows the

Stokes solution of an oscillating plate, at a Womersly number of 10, using the
analytical nonlinear differential governing equation derived in Sections 9.3 and 9.4.
Here the nonlinear terms of Equation 9.31 were neglected according to the long-
wavelength approximation made in Section 9.4. The stream function calculated for
this Stokes flow is shown in Figure 11.2. Errors are discussed in Section 11.3.

11.2 Travelling-wave forcing


The analytical solution was found to match the numerical results at low wave ampli-
tudes. This result provides validation of both the analytical and numerical methods
used for low amplitudes. Qualitatively, these results agree with those of Shen et al.
(2003), who investigated flow over a wavy wall with different phase-speeds of stream-
wise wave velocity. The analytical and numerical solutions are shown and compared
in Figures 11.3, 11.4, 11.5 and in the additional figures presented in Appendix B.
The top plots show contours of constant spanwise velocity, the center plots give con-
tours of constant vertical velocity and the lower plots show contours of pressure. The
numerical solution is shown by solid lines for positive values, and dashed lines for
negative values; the analytical solution is shown by dot-dashed lines where values
134 CHAPTER 11. ACTUATION RESULTS

are positive and dotted lines for negative values. The vertical contour lines represent
zero-values. For low amplitudes, the numerical and analytical solutions coincide, as
can be seen in Figure 11.3a.
The spanwise oscillatory profile due to travelling-wave wall motion is character-
istic of an oscillatory Stokes layer. Figure 11.3 shows the solution is approximately
sinusoidal, with the first and second Floquet modes being the predominant modes ex-
cited at low Reynolds number and amplitude. Vertical velocity decays with increasing
distance away from the wall, in approximately exponential fashion. Phase-shifting of
the response flow field can be seen near the wall. The pressure is nearly symmetrical,
and also decays exponentially away from the wall.
The numerical and analytical solutions of spanwise and vertical velocity are gen-
erally in good agreement, and the pressure field is well-represented by the approxima-
tion made in Equation 9.25. As the amplitude increases, spanwise velocity becomes
increasingly nonsymmetric in the numerical solution. The analytical solution also ex-
hibits similar asymmetry, but it does not coincide as well with the numerical solution
once the wave amplitude and Reynolds number become comparatively large, as can
be seen in Figures 11.4 and 11.5.
As amplitude and Reynolds number increase, the numerically determined span-
wise velocity in the trough is lower than the corresponding theoretical value. The
differences between the analytical approximation and the more accurate numerical
solution arise from the effective inclusion of only two Floquet modes of the more ac-
curate Equation 9.13, because the approximate Equation 9.31 was solved. It can be
seen from numerical results that higher-frequency Floquet components become signif-
icant near the wall at higher Reynolds numbers. These higher-frequency components
decay much faster than the fundamental-mode response, which is why the analytical
solution always closely matches the numerical solution further away from the wall
(y > 0.2).
It is notable that at higher Reynolds numbers and amplitudes the variation of the
streamfunction with x is larger. This variation is shown in Figure 11.6, and is due
to the nonlinear component of Equation 9.31 that involves the term ei(ax+Ωt) . This
variation corresponds to the second nonlinear mode of Equation 9.13. It should also
be noted in many cases that because the variations are a small percentage of the time-
averaged values over the entire domain, a reasonable approximation is still achieved
if this non-linear term is neglected. In some cases neglecting the nonlinear term
may be desirable, as it is this term that may cause computational difficulties when
solving the equation in some cases. The left plots of Figure 11.6 show this non-linear
term causes fluctuation in the streamfunction of approximately 10% over the domain
for a Reynolds number of 102 . This fluctuation level increases with increasing wave
11.2. TRAVELLING-WAVE FORCING 135

(a) contour spacing = 0.005 (d) contour spacing = 0.02


0.5 0.5
nondimensional height, y

nondimensional height, y
0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0

0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1

(b) contour spacing = 0.005 (e) contour spacing = 0.02


0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0

0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1

(c) contour spacing = 0.005 (f) contour spacing = 0.02


0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0

0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1


nondimensional spanwise distance, z nondimensional spanwise distance, z

Figure 11.3: Flow field for Re = 102 at wave amplitudes of 0.005 (left) and 0.020
(right), shown by contours of spanwise velocity (a,d), vertical velocity (b,e) and pres-
sure (c,f). Numerical solution is shown in solid (positive) and dashed (negative) lines,
the analytical solution is shown by dot-dashed (positive) and dotted (negative) lines.
Wave motion is from left to right.
136 CHAPTER 11. ACTUATION RESULTS

(a) contour spacing = 0.04 (d) contour spacing = 0.06


0.5 0.5
nondimensional height, y

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0

0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1

(b) contour spacing = 0.04 (e) contour spacing = 0.06


0.5 0.5
nondimensional height, y

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0

0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1

(c) contour spacing = 0.025 (f) contour spacing = 0.05


0.5 0.5
nondimensional height, y

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0

0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1


nondimensional spanwise distance, z nondimensional spanwise distance, z

Figure 11.4: Flow field for Re = 102 at wave amplitudes of 0.035 (left) and 0.050
(right), shown by contours of spanwise velocity (a,d), vertical velocity (b,e) and pres-
sure (c,f). Numerical solution is shown in solid (positive) and dashed (negative) lines,
the analytical solution is shown by dot-dashed (positive) and dotted (negative) lines.
Wave motion is from left to right.
11.2. TRAVELLING-WAVE FORCING 137

(a) contour spacing = 0.06 (d) contour spacing = 0.06


0.5 0.5
nondimensional height, y

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0

0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1

(b) contour spacing = 0.06 (e) contour spacing = 0.06


0.5 0.5
nondimensional height, y

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0

0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1

(c) contour spacing = 0.04 (f) contour spacing = 0.04


0.5 0.5
nondimensional height, y

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0

0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1


nondimensional spanwise distance, z nondimensional spanwise distance, z

Figure 11.5: Flow field for Re = 103 (left) and Re = 104 (right) at wave amplitude
of 0.050, shown by contours of spanwise velocity (a,d), vertical velocity (b,e) and
pressure (c,f). Numerical solution is shown in solid (positive) and dashed (negative)
lines, the analytical solution is shown by dot-dashed (positive) and dotted (negative)
lines. Wave motion is from left to right.
138 CHAPTER 11. ACTUATION RESULTS

(a) (b)
nondimensional height, y

0.5 0.5

nondimensional height, y
0.4 0.4

0.3 0.3

0.2 0.2

0.1
0.1
0
0
−1 0 1 2 3 4 5 6 7
0 0.2 0.4 0.6 0.8 1
−3
x 10

(c) (d)
0.5 0.5
nondimensional height, y

nondimensional height, y
0.4 0.4

0.3
0.3
0.2
0.2
0.1
0.1
0
0
−0.05 −0.04 −0.03 −0.02 −0.01 0 0 0.2 0.4 0.6 0.8 1
Streamfunction nondimensional spanwise distance, z

Figure 11.6: Profiles of the analytical streamfunction ψ in y over one period at fixed
x position, and contours of the deviation of ψ from the time-averaged mean (right)
for a Reynolds number of 102 and wave amplitude of 0.050. The real part is given in
plots (a) and (b), and the imaginary part is shown in plots (c) and (d). Wave motion
is from left to right in (b) and (d).

amplitude and increasing Reynolds number, which upon inspection of the nonlinear
governing Equation 9.13 is due to greater coupling of Floquet response modes.

It can be seen that as Reynolds number increases, the phase shift in velocity con-
tours over the y-domain decreases. In Figure 11.5 this decrease is apparent as vertical
lines in the vertical velocity and pressure contour plots. Phase shifting is typical of
oscillatory flows, and is a function of the nondimensional parameters governing the
flow.

The results presented in Figures 11.3 through 11.5 are of only the oscillatory com-
ponent of the flow, but numerical results show that a mean flow is gradually induced
in the direction of the wave motion. It was found that the oscillatory component was
largely independent of the mean flow. The development of this secondary flow over
time is shown in Figure 11.7 for Reynolds numbers of Re = 102 (left) and Re = 105
(right). Note that spanwise velocity was subject to a zero-velocity boundary condition
at the wave surface and at the top of the y-domain. The mean velocity component
had a more fuller profile at lower Reynolds numbers, as can be seen from Figure 11.8.
11.2. TRAVELLING-WAVE FORCING 139

1 1

0.9 0.9

0.8 0.8
nondimensional height, y

nondimensional height, y
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 0.01 0.02 0.03 0.04 0.05 0.06 −0.05 0 0.05 0.1 0.15 0.2 0.25 0.3
mean spanwise velocity mean spanwise velocity

Figure 11.7: Profiles of mean (quasi-steady) spanwise velocity induced in the numer-
ical solution for Re = 102 (left) and Re = 105 (right), at t = 5, 10, 20, 30, 40, 50, 60.
The direction of the mean flow is in the same direction as wave-motion.

1
Re = 102
0.9
Re = 103
0.8 Re = 104
Re = 105
nondimensional height, y

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
0 0.05 0.1 0.15 0.2 0.25
mean spanwise velocity

Figure 11.8: Profiles of mean (quasi-steady) spanwise velocity induced in the numer-
ical solution at t = 60. The direction of the mean flow is in the same direction as
wave-motion.
140 CHAPTER 11. ACTUATION RESULTS

11.3 Errors

There are two sources of error in the analytical approximation. First, the most
significant error arises from the low number of Floquet modes used in the analytical
approximation of Equation 9.31, compared to Equation 9.13. Differences between
the more accurate numerical solution and the analytical approximation are most
pronounced in the trough of the wave.
The analytical and numerical solutions converge at low amplitudes. With in-
creasing amplitude, however, the analytical approximation exhibits a more time-
symmetrical solution, and the more accurate numerical solution shows less positive
spanwise velocity in the wave trough. If several Floquet modes were included in the
nonlinear Equation 9.13 and this nonlinear equation were solved, it is expected the
solution would more closely match the numerical answer.
The second source of error in the analytical approximation arises from the as-
sumptions made in the derivation of the governing Equation 9.31. In particular, this
equation requires the assumption of exponential pressure decay. The decay rate of
the fundamental (synchronous) sinusoidal pressure component is well approximated
by Equation 9.33, but the fundamental reason for this equation to hold is not clear.
It should be noted, however, that at high Reynolds number a harmonic pressure
component could be seen in the numerical results near the wall. The nonlinear Equa-
tion 9.13 does not assume any form of the pressure field, and therefore represents a
more general formulation.
The two predominant sources of possible error in the numerical solution are due
to insufficient grid resolution near the wall and the errors due to the upper boundary
condition. Grid independence of the solution was confirmed for Reynolds numbers of
up to 1000, but it can be seen that at the highest Reynolds number tested (Re = 105 )
that large velocity gradients are present near the wall. These high gradients suggest
the use of a finer grid near the wall at high Reynolds number for increased accuracy.
Second, the vertical location of the upper boundary is important to the solution
at very low Reynolds numbers. It was found that a very low Reynolds number of
10, the solution near the upper boundary was slightly more accurate with the upper
boundary moved to y = 2, rather than y = 1. However, this error was less than 10%.
For Reynolds number of 100 and above, increasing the vertical height of the domain
had a negligible effect on the oscillatory solution, thereby confirming that the vertical
domain was large enough for reasonable accuracy of the oscillatory flow. However,
the condition of zero spanwise velocity imposed on the top boundary influenced the
mean (quasi-steady) spanwise velocity profile, as can be seen in Figure 11.8.
11.4. CONCLUSIONS 141

11.4 Conclusions
The analytical technique outlined provides a good approximation to the oscillatory
flow over a wavy wall for low wave amplitudes. The approximated governing equation
derived can also be used to find Stokes oscillating plate solutions, when appropriate
boundary conditions are applied. It can be seen from the results that travelling-wave
excitation is suitable for boundary layer control applications.
The response flow decays exponentially away from the wall, and the major velocity
and pressure components are of the same frequency as the driving wall motion. At
low Reynolds number, the oscillatory profile induced is relatively smooth in velocity
gradient. However, higher Reynolds numbers lead to very high velocity gradients
immediately next to the wall, which correspond to the excitation of higher, coupled
response harmonics.
It is noted that for a boundary layer control application the amplitude and velocity
of the wave would need to be considerable, with the wave travelling at a velocity com-
parable to or even greater than the free-stream velocity. Considerable technological
research is therefore required for development of a practical actuation method.

11.5 Future work


The boundary conditions herein have been kept simple, using a sinusoidal waveform.
However, this form of analysis could have broad application for many two-dimensional
flows forced by wall motion. The results have shown that the response flow is pre-
dominantly of the same frequency as the forcing flow, with only small influence due
to higher harmonic modes.
It is therefore suggested that the nonlinear governing equation for such flow given
by Equation 9.13 would yield an accurate solution to flows over wavy-walls. Indeed,
more complex wall-shapes could be incorporated into the boundary conditions. Us-
ing the approximated solution given by Equation 9.31 could be used as an initial
guess, thereby assisting convergence to the final nonlinear solution using an iterative
numerical method.
Experimental verification of the findings elucidated here would be required if sim-
ilar travelling-wave excitation was to be used for boundary layer control. Before
implementation, three-dimensional effects in particular must be determined. In light
of the stabilisation mechanism presented in Part III which such forcing may produce
in boundary layer flows, such experimentation may lead to an effective method of
boundary layer flow control.
Bibliography

Acarlar, M. S. and Smith, C. R. (1987). A study of hairpin vortices in a laminar


boundary layer. part ii: hairpin vortices generated by fluid injection. Journal of
Fluid Mechanics, 175:43 – 83.

Adrian, R. J., Meinhart, C. D., and Tomkins, C. D. (2000). Vortex organization


in the outer region of the turbulent boundary layer. Journal of Fluid Mechanics,
422:1–54.

Alfredsson, P. H. and Johansson, A. V. (1984). Time scales in turbulent channel flow.


Physics of Fluids, 27(1974 - 1981).

Andersson, P., Berggren, M., and Henningson, D. (1999). Optimal disturbances and
bypass transition in boundary layers. Physics of Fluids, 11(1):134–150.

Arscott, F. M. (1964). Periodic differential equations. International series of mono-


graphs in pure and applied mathematics. Pergamon Press, Poland.

Asai, M., Minagawa, M., and Nishioka, M. (2002). The instability and breakdown of
a near-wall low-speed streak. Journal of Fluid Mechanics, (455):289–314.

Bake, S., Ivanov, A. V., Fernholz, H. H., Neemann, K., and Kachanov, Y. S. (2002).
Receptivity of boundary layers to three-dimensional disturbances. European Jour-
nal of Mechanics, B/Fluids, 21(1):29–48.

Baker, J., Myatt, J., and Christofides, P. D. (2002). Drag reduction in flow over a flat
plate using active feedback control. Computers and Chemical Engineering, 26(7-8
Aug 15 2002):1095–1102.

Balakumar, P. and Widnall, S. E. (1986). Application of unsteady aerodynamics to


large-eddy breakup devices in a turbulent flow. Physics of Fluids, 29(6):1779–1787.

Baron, A. and Quadrio, M. (1996). Turbulent drag reduction by spanwise oscillations.


Appl. Sci. Res., 55:311 – 326.

142
BIBLIOGRAPHY 143

Bayly, B. J. and Orszag, S. A. (1988). Instability mechanisms in shear-flow transition.


Annual Review of Fluid Mechanics, 20:359 – 391.

Bechert, D. W. and Bartenwerfer, M. (1989). The viscous flow on surfaces with


longitudinal ribs. Journal of Fluid Mechanics, 206:105–129.

Bechert, D. W., Bruse, M., Hage, W., and Meyer, R. (2000). Fluid mechanics of bi-
ological surfaces and their technological application. Naturwissenschaften, 87:157–
171.

Bechert, D. W., Hoppe, G., and Reif, W.-E. (1985). On the drag reduction of the
shark skin. Technical report, AIAA paper No. 85-0546.

Benhalilou, M., Anselmet, F., and Fulachier, L. (1994). Conditional reynolds stress
on a v-grooved surface. Physics of Fluids, 6(6):2101–2117.

Berger, T. W., Kim, J., Lee, C., and Lim, J. (2000). Turbulent boundary layer control
utilizing the lorentz force. Physics of Fluids, 12(3):631.

Berlin, S. (1998). Oblique waves in boundary layer transition. Doctoral thesis, Royal
Institute of Technology.

Berlin, S., Lundbladh, A., and Henningson, D. S. (1994). Spatial simulations of


oblique transition in a boundary layer. Physics of Fluids, 6(6):1949 – 1951.

Berlin, S., Wiegel, M., and Henningson, D. S. (1999). Numerical and experimental
investigations of oblique boundary layer transition. Journal of Fluid Mechanics,
393:23 – 57.

Bernard, T. W., Kim, J., Lee, C., and Lim, J. (1993). Vortex dynamics and the
production of reynolds stress. Jounal of Fluid Mechanics, 253:385–419.

Bertolotti, F. (2000). Receptivity of three-dimensional boundary-layers to localized


wall roughness and suction. Physics of Fluids, 12(7 Jul):1799–1809.

Bertolotti, F. P., Herbert, T., and Spalart, P. R. (1992). Linear and nonlinear stability
of the blasius boundary layer. Journal of Fluid Mechanics, 242:441 – 474.

Blackwelder, R. F. and Eckelmann, H. (1979). Streamwise vortices associated with


the bursting phenomenon. Journal of Fluid Mechanics, 94:577 – 594.

Blackwelder, R. F. and Haritonidis, J. H. (1983). Scaling of the bursting frequency


in turbulent boundary layers. Journal of Fluid Mechanics, 132:87 – 103.
144 BIBLIOGRAPHY

Blasius, H. (1908). Grenzschicheten in flussigkeiten mit kleiner reibung. Z. Math.


Phys., 56:1 – 37.

Brandt, L., Cossu, C., Chomaz, J.-M., Huerre, P., and Henningson, D. S. (2003). On
the convectively unstable nature of optimal streaks in boundary layers. Journal of
Fluid Mechanics, (485):221–242.

Brandt, L. and Henningson, D. S. (2002). Transition of streamwise streaks in zero-


pressure-gradient boundary layers. Journal of Fluid Mechanics, (472):229–261.

Braslow, A. L. (1999). A history of suction-type laminar-flow control with emphasis


on flight research, volume 13 of Monographs in Aerospace History. NASA History
Division, NASA Headquarters, Washington, DC.

Brooke, J. W. and Hanratty, T. J. (1993). Origin of turbulence-producing eddies in


a channel flow. Physics of Fluids, A 5:1011.

Brown, G. L. and Thomas, A. S. W. (1977). Large structure in a turbulent boundary


layer. Physics of Fluids, 20:S243.

Cantwell, B. J., Coles, D., and Dimotakis, P. E. (1978). Structure and entrainment in
the plane of symmetry of a turbulent spot. Jounal of Fluid Mechanics, 87:641–672.

Carpenter, P. (1993). Optimization of multiple-panel compliant walls for delay of


laminar-turbulent transition. AIAA Journal, 31:1187 – 1188.

Chang, Y., Collis, S. S., and Ramakrishnan, S. (2002). Viscous effects in control of
near-wall turbulence. Physics of Fluids, 14(11):4069–4080.

Choi, H. (2002a). Active control methods for drag reduction in flow over bluff bod-
ies. In Proceedings of the 2002 ASME Joint U.S.-European Fluids Engineering
Conference,Jul 14-18 2002, volume 257 of American Society of Mechanical Engi-
neers, Fluids Engineering Division (Publication) FED, pages 1431–1436, Montreal,
Que.,United States. American Society of Mechanical Engineers.

Choi, H., Moin, P., and Kim, J. (1991). On the effect of riblets in fully developed
laminar channel flows. Physics of Fluids A: Fluid Dynamics, 3(8):1892–1896.

Choi, J.-I., Xu, C.-X., and Sung, H. J. (2002). Drag reduction by spanwise wall
oscillation in wall-bounded turbulent flows. AIAA Journal, 40(5):842–850.

Choi, K.-S. (2002b). Near-wall structure of turbulent boundary layer with spanwise-
wall oscillation. Physics of Fluids, 14(7 July 2002):2530–2542.
BIBLIOGRAPHY 145

Choi, K.-S. and Clayton, B. R. (2001). Mechanism of turbulent drag reduction with
wall oscillation. International Journal of Heat and Fluid Flow, 22(1):1–9.

Choi, K.-S., DeBisschop, J.-R., and Clayton, B. R. (1998). Turbulent boundary layer
control by means of spanwise-wall oscillation. AIAA Journal, 36:1157.

Choi, K.-S. and Graham, M. (1998). Drag reduction of turbulent pipe flows by
circular-wall oscillation. Physics of Fluids, 10(1):7.

Chu, D. C. and Karniadakis, G. E. (1993). A direct numerical simulation of laminar


and turbulent flow over riblet-mounted surfaces. Journal of Fluid Mechanics, 250:1–
42.

Clark III, H. and Deutsch, S. (1991). Microbubble skin friction reduction on an


axisymmetric body under the influence of applied axial pressure gradients. Physics
of Fluids A: Fluid Dynamics, 3(12):2948–2954.

Corrino, E. R. and Brodkey, R. S. (1969). A visual investigation of the wall region in


turbulent flow. Journal of Fluid Mechanics, 37:1 – 30.

Corrsin, S. (1943). Investigation of flow in an axially symmetric heated jet of air.


Technical report, NACA Adv. Conf. Rep. 3123.

Corrsin, S. and Kistler, A. (1954). The free-stream boundaries of turbulent flows.


Technical report, NACA TN-3133.

De Angelis, V., Lombardi, P., and Banerjee, S. (1997). Direct numerical simulation
of turbulent flow over a wavy wall. Physics of Fluids, 9(8 Aug):2429–2442.

Deutsch, S. and Castano, J. (1986). Microbubble skin friction reduction on an ax-


isymmetric body. Physics of Fluids, 29(11):3590–3597.

Dhanak, M. and Si, C. (1999). On reduction of turbulent wall friction through span-
wise wall oscillations. Journal of Fluid Mechanics, 383(Mar):175–195.

Di Cicca, G. M., Iuso, G., Spazzini, P. G., and Onorato, M. (2002). Particle image
velocimetry investigation of a turbulent boundary layer manipulated by spanwise
wall oscillations. Journal of Fluid Mechanics, (467 Sep 25 2002):41–56.

Djenidi, L., Anselmet, F., Liandrat, J., and Fulachier, L. (1994). Laminar boundary
layer over riblets. Physics of Fluids, 6(9):2993–2999.

Doligalski, T. L. and Walker, J. D. A. (1984). The boundary layer induced by a


convected two-dimensional vortex. Journal of Fluid Mechanics, 139:1.
146 BIBLIOGRAPHY

Du, Y., Symeonidis, V., and Karniadakis, G. (2002). Drag reduction in wall-bounded
turbulence via a transverse travelling wave. Journal of Fluid Mechanics, (457):1–34.

Faraday, M. (1831). On a particular class of acoustical figures; and on certain forms


assumed by groups of particles upon vibrating elastic surfaces. Philos. Trans. R.
Soc. Lond., 121:299–340.

Favre, A. J., Gaviglio, J. J., and Dumas, R. (1957). Space-time double correlations
and spectra in a turbulent boundary layer. Journal of Fluid Mechanics, 2:313 –
341.

Fjørtoft, R. (1950). Application of integral theorems in deriving criteria for instability


for laminar flows and for the baroclinic circular vortex. Geofys. Publ., Oslo, 17(6):1–
52.

Floquet, G. (1883). Sur les équations différentielles linéaries à coefficients périodiques.


Ann. École Norm. Sup., 12:47–88.

Gad-el Hak, M. (1996a). Complaint coatings: A decade of progress. Applied Mechan-


ics Reviews, 49(10 - 2):S147 – S157.

Gad-el Hak, M. (1996b). Modern developments in flow control. Applied Mechanics


Reviews, 49(7):365–380.

Gad-el Hak, M. (2000). Flow control : passive, active, and reactive flow management.
Cambridge University Press, Cambridge.

Goldstein, M. E. and Hultgren, L. S. (1989). Boundary-layer receptivity to long-wave


disturbances. Annual Review of Fluid Mechanics, 21:137–166.

Grant, J. L. (1958). The large eddies of turbulent motion. Journal of Fluid Mechanics,
4:149 – 190.

Hagiwara, Y., Hana, H., M., T., and Murai, S. (2000). Numberical simulation of the
interactions of highly entangled polymers with coherent structure in a turbulent
channel flow. International Journal of Heat and Fluid Flow, 21:589 – 598.

Haj-Hariri, H. (1988). Transformations reducing the order of the parameter in differ-


ential eigenvalue problems. Journal of Computational Physics, pages 472 – 484.

Hall, P. and Horseman, N. J. (1991). The linear inviscid secondary instability of


longitudinal vortex structures in boundary layers. Journal of Fluid Mechanics,
232:357.
BIBLIOGRAPHY 147

Hamilton, J. M., Kim, J., and Waleffe, F. (1995). Regeneration mechanisms of near-
wall turbulence structures. Journal of Fluid Mechanics, 287:317.

Han, M., Huh, J., Lee, S., and Lee, S. (2002). Micro-riblet film for drag reduc-
tion. In Proceedings of the Pacific Rim Workshop on Transducers and Micro/Nano
Technologies, pages 47–50, Xiamen, China.

Handler, R. A., Levich, E., and Sirovich, L. (1992). Drag reduction in turbulent
channel flow by phase randomization. Physics of Fluids A, 5(3):686–694.

Hanifi, A., Schmid, P. J., and Henningson, D. S. (1996). Transient growth in com-
pressible boundary layer flow. Physics of Fluids, 8(3):826–837.

Head, M. and Bandyopadhyay, P. (1981). New aspects of a turbulent boundary-layer


structure. Journal of Fluid Mechanics, 107:297.

Henningson, D. (2004). Private communication.

Henningson, D., Lundbladh, A., and Johansson, A. V. (1993). A mechanism for


bypass transition from localized disturbances. Jounal of Fluid Mechanics, 250:169–
207.

Herbert, T. (1983). Secondary instability of plane channel flow to subharmonic three-


dimensional disturbances. Physics of Fluids, 26(4):871–874.

Herbert, T. (1988). Secondary instability of boundary layers. Annual Review of Fluid


Mechanics, 20:487 – 526.

Hetsroni, G., Zakin, J. L., and Mosyak, A. (1997). Low-speed streaks in drag-reduced
turbulent flow. Physics of Fluids, 9(8):2397 – 2404.

Ho, C.-M. and Tai, Y.-C. (1996). Review: MEMS and its applications for flow
control. Journal of Fluids Engineering, 118:437 – 447.

Ho, C.-M. and Tai, Y.-C. (1998). Micro-electro-mechanical systems (MEMS) and
fluid flows. Annual Review of Fluid Mechanics, 30:579 – 612.

Högberg, M. and Henningson, D. S. (2002). Linear optimal control applied to in-


stabilities in spatially developing boundary layers. Journal of Fluid Mechanics,
(470):151–179.

Hu, H.-C. and Kelly, R. E. (1995). Effect of a time-periodic axial shear flow upon the
onset of taylor vortices. Physical Review E, 51(4):3242–3251.
148 BIBLIOGRAPHY

Hultgren, L. S. and Gustavsson, L. H. (1981). Algebraic growth of disturbances in a


laminar boundary layer. Physics of Fluids, 24(6):1000 – 1004.

Jeong, J., Hussain, F., Schoppa, W., and Kim, J. (1997). Coherent structures near
the wall in a turbulent channel flow. Journal of Fluid Mechanics, 332:185–214.

Jiménez, J. and Moin, P. (1991). The minimal flow unit in near-wall turbulence.
Journal of Fluid Mechanics, 225:213 – 240.

Jiménez, J. and Pinelli, A. (1999). The autonomous cycle of near-wall turbulence.


Journal of Fluid Mechanics, 389:335 – 359.

Joslin, R. D. and Morris, P. J. (1992). The effect of compliant walls on secondary


instabilities in boundary-layer transition. AIAA Journal, 30(2):332–339.

Jung, W., Mangiavacchi, N., and Akhavan, R. (1992). Suppression of turbulence in


wall-bounded flows by high-frequency spanwise oscillations. Physics of Fluids A
(Fluid Dynamics), 4(8 Aug):1605.

Karniadakis, G. and Choi, K.-S. (2003). Mechanisms on transverse motions in tur-


bulent wall flows. Annual Review of Fluid Mechanics, 35:45–62.

Kawahara, G., Jimenez, J., Uhlmann, M., and Pinelli, A. (2003). Linear instability of
a corrugated vortex sheet - a model for streak instability. Jounal of Fluid Mechanics,
483:315–342.

Kawahara, G., Jimnez, J., Uhlmann, M., and Pinelli, A. (1998). The instability of
streaks in near-wall turbulence. Center for Turbulence Research Annual Research
Briefs, 1998:155 – 170.

Kelvin (1910). Mathematical and physical papers, volume 4 of Hydrodynamics


and General Dynamics. Cambridge University Press, Cambridge. Lord Kelvin
(Sir William Thomson).

Kerczek, C. v. and Davis, S. H. (1974). Linear stability theory of oscillatory stokes


layers. Jounal of Fluid Mechanics, 62:689–703.

Kiesow, R. O. and Plesniak, M. W. (2003). Near-wall physics of a shear-driven three-


dimensional turbulent boundary layer with varying crossflow. Journal of Fluid
Mechanics, (484):1–39.

Kim, C., Jeon, W.-P., Park, J., and Choi, H. (2003). Effect of a localized time-periodic
wall motion on a turbulent boundary layer flow. Physics of Fluids, 15(1 January
2003):265–268.
BIBLIOGRAPHY 149

Kim, H., Kline, S. J., and Reynolds, W. C. (1971). The production of turbulence near
a smooth wall in a turbulent boundary layer. Journal of Fluid Mechanics, 50:133.

Kim, J., Moin, P., and Moser, R. (1987). Turbulence statistics in fully-developed
channel flow at low reynolds number. Journal of Fluid Mechanics, 177:133–166.

Kim, J. and Spalart, P. R. (1987). Scaling of the bursting frequency in turbulent


boundary layers at low reynolds numbers. Physics of Fluids, 30(11):3326–3328.

King, R. A. (2000). Receptivity and growth of two- and three-dimensional distur-


bances in a blasius boundary layer. PhD Thesis, February 11, 2000(Department of
Aeronautics and Astronautics at the Massachusetts Institute of Technology).

Klebanoff, P. S., Tidestrom, K. D., and Sargent, L. M. (1962). The three-dimensional


nature of boundary-layer instability. Jounal of Fluid Mechanics, 12:1–34.

Kline, S. J., Reynolds, W. C., Schraub, F. A., and Runstadler, P. W. (1967). The
structure of turbulent boundary layers. Journal of Fluid Mechanics, 30(4):741 –
773.

Kline, S. J. and Runstadler, P. W. (1959). Some preliminary results of visual studies


of the flow model of the wall layers of the turbulent boundary layer. Trans. ASME,
Ser., E 2:166–170.

Kodama, Y., Kakugawa, A., Takahashi, T., and Kawashima, H. (2000). Experimental
study on microbubbles and their applicability to ships for skin friction reduction.
International Journal of Heat and Fluid Flow, 21(5):582–588.

Kodama, Y., Kakugawa, A., Takahashi, T., Nagaya, S., and Kawamura, T. (2001).
Drag reduction of ships by microbubbles. Technical report, 24th UJNR/MFP Joint
Meeting, Marine Facilities Panel of the U.S./Japan Cooperative Program in Natural
Resources.

Koosh, V., Gupta, B., Babcock, D., Goodman, R., Jiang, F., Tai, Y., Tungy, S., and
Hoy, C. (1999). Analog VLSI system for active drag reduction. In Cauwenberghs,
G. and Bayoumi, M. A., editors, Learning on Silicon: Adaptive Vlsi Neural Systems,
volume 512 of Kluwer International Series in Engineering and Computer Science.
Kluwer Academic Publishers.

Kral, L. D. (2001). Active flow control technology. ASME fluids engineering division
technical brief, Washington University.
150 BIBLIOGRAPHY

Kravchenko, A. G., Choi, H., and Moin, P. (1993). On the relation of near-wall
streamwise vorticies to wall skin friction in turbulent boundary layers. Physics of
Fluids, 5:3307 – 3309.

Kuchment, P. (1993). Floquet theory for partial differential equations, volume 60 of


Operator Theory Advances and Applications. Birkhäuser Verlag, Boston.

Laadhari, F., Skandaji, L., and Morel, R. (1994). Turbulence reduction in a boundary
layer by a local spanwise oscillating surface. Physics of Fluids, 6(10):3218–3220.

Landahl, M. T. (1977). Dynamics of boundary layer turbulence and the mechanism


of drag reduction. Physics of Fluids, 20(10 (part 11)):S55–S63.

Lee, C., Kim, J., Babcock, D., and Goodman, R. (1997). Application of neural
networks to turbulence control for drag reduction. Physics of Fluids, 9(Jun 6):1740–
1747.

Lee, T., Fisher, M., and Schwartz, W. H. (1995). Investigation of the effects of a
compliant surface on boundary-layer stability. Jounal of Fluid Mechanics, 288:37–
58.

Li, F. and Malik, M. R. (1995). Fundamental and subharmonic secondary instabilities


of grtler vortices. Journal of Fluid Mechanics, 297:77 – 100.

Liu, K. N., Christodoulou, C., Riccius, O., and Joseph, D. D. (1990). Drag reduction
in pipes lined with riblets. AIAA Journal, 28(10):1697–1698.

Luchini, P. (2000). Reynolds-number-independent instability of the boundary layer


over a flat surface: optimal perturbations. Journal of Fluid Mechanics, 404:289 –
309.

Lumley, J. L. (1969). Drag reduction by additives. Annual Review of Fluid Mechanics,


1:367 – 384.

Lumley, J. L. (1977). Drag reduction in two phase and polymer flows. Physics of
Fluids, 20(10, part II):S64 – S71.

MacCormack, W., Tutty, O., Rogers, E., and Nelson, P. (2002). Stochastic optimi-
sation based control of boundary layer transition. Control Engineering Practice,
10(3):243–260.

MacManus, D. G. and Eaton, J. A. (2000). Flow physics of discrete boundary layer


suction - measurements and predictions. Jounal of Fluid Mechanics, 417:47–75.
BIBLIOGRAPHY 151

Madavan, N. K., Deutsch, S., and Merkle, C. L. (1984). Reduction of turbulent skin
friction by microbubbles. Physics of Fluids, 27(2):356–363.

Malik, M. R. (1990). Numerical methods for hypersonic boundary layer stability.


Journal of Computational Physics, 86(376-413).

Marques, F. and Lopez, J. M. (1997). Taylor-couette flow with axial oscillations of


the inner cylinder: Floquet analysis of the basic flow. Jounal of Fluid Mechanics,
348:153–175.

Marques, F. and Lopez, J. M. (2000). Spatial and temporal resonances in a periodi-


cally forced hydrodynamic system. Physica D, 136:340–352.

Mathieu, E. (1868). Mémoire sur le mouvement vibratoire d’une membrane de forme


elliptique. Jour. de. Math. Pures et Appliquées (Jour. de Loiuville), 13:137.

McLachlan, N. W. (1947). Theory and application of Mathieu functions. Oxford


University Press, London.

McLean, J. W. (1983). Computation of turbulent flow over a moving wavy wall.


Physics of Fluids, 26(8):2065 – 2073.

McMichael, J. M. (1996). Progress and prospects for active flow control using micro-
fabricated electromechanical systems (mems). Technical report, AIAA paper no.
96-0306.

Meseguer, A. and Marques, F. (2000). On the competition between centrifugal and


shear instability in spiral couette flow. Jounal of Fluid Mechanics, 402:33–56.

Murlis, J., Tsai, H. M., and Bradshaw, P. (1982). The structure of turbulent boundary
layers at low reynolds numbers. Journal of Fluid Mechanics, 122:13 – 56.

Murray, B. T., McFadden, G. B., and Coriell, S. R. (1990). Stabilisation of taylor-


couette flow due to time-periodic outer cylinder oscillation. Physics of Fluids A,
2(12):2147–2156.

Offen, G. R. and Kline, S. J. (1974). Combined dye-streak and hydrogen-bubble visual


observations of a turbulent boundary layer. Journal of Fluid Mechanics, 62:223 –
239.

Offen, G. R. and Kline, S. J. (1975). A proposed model of the bursting process in


turbulent boundary layers. Journal of Fluid Mechanics, 70(2):209–228.

Or, A. C. (1997). Finite-wavelength instability in a horizontal liquid layer on an


oscillating plate. Jounal of Fluid Mechanics, 335:213–232.
152 BIBLIOGRAPHY

Orlandi, P. and Fatica, M. (1997). Direct simulations of turbulent flow in a pipe


rotating about its axis. Journal of Fluid Mechanics, 343(Jul 25):43–72.

Orr, W. M. F. (1907). The stability or instability of the steady motions of a perfect


liquied and of a viscous liquid. part i: A perfect liquid; part ii: A viscous liquid. In
Proceedings of the Royal Irish Academy, volume 27, pages 9–138.

Orszag, S. A. (1971). Accurate solution of the orr-sommerfeld equation. Jounal of


Fluid Mechanics, 50:689 – 703.

Panton, R. L., e. (1997). Self-sustaining mechanisms of wall turbulence. Computa-


tional Mechanics Publications, Southampton, Great Britain.

Patterson, G. K., Chosnek, J., and Zakin, J. L. (1977). Turbulence structure in drag
reducing polymer solutions. Physics of Fluids, 20(10, part II):S89 – S99.

Peyret, R. (2002). Spectral methods for incompressible viscous flow, volume 148 of
Applied Mathematical Sciences. Springer-Verlag, New York.

Phillips, W., Wu, Z., and Lumley, J. (1996). On the formation of longitudinal vortices
in a turbulent boundary layer over wavy terrain. Journal of Fluid Mechanics,
326:321–341.

Pozrikidis, C. and Yon, S. A. (1998). Numerical simulation of parametric instability in


two and three-dimensional fluid interfaces. In Proceedings of the Fourth Micrograv-
ity Fluid Physics and Transport Phenomena Conference, pages 390–393, Cleveland,
Ohio.

Prandtl, L. (1904). Uber flussigkeitsbewegung bei sehr kleiner reibung. In Verh. III.
Intern. Math. Kongr., pages S. 484 – 491, Heidelberg. Teubner, Leipzig, 1905.

Quadrio, M. and Ricco, P. (2003). Critical assessment of turbulent drag reduction


through spanwise wall oscillaitons.

Quadrio, M. and Sibilla, S. (2000). Numerical simulation of turbulent flow in a pipe


oscillating around its axis. Journal of Fluid Mechanics, 424:217–241.

Rayleigh, L. (1880). On the stability of certain fluid motions. Proc. Math. Soc. Lond.,
11:57 – 70.

Rayleigh, L. (1887). On the stability of certain fluid motions. Proc. Math. Soc. Lond.,
19:67 – 74.
BIBLIOGRAPHY 153

Reddy, S. C., Schmid, P. J., Baggett, J. S., and Henningson, D. S. (1998). On stability
of streamwise streaks and transition thresholds in plane channel flows. Journal of
Fluid Mechanics, 365:269–303.

Reed, H. L., Saric, W. S., and Arnal, D. (1996). Linear stability theory applied to
boundary layers. Annual Review of Fluid Mechanics, page 389.

Reshotko, E. (2001). Transient growth: a factor in bypass transition. Physics of


Fluids, 13(5):1067–1075.

Robinson, S. K. (1991). Coherent motions in the turbulent boundary layer. Annual


Review of Fluid Mechanics, 23:601–639.

Runstadler, P. G., Kline, S. J., and Reynolds, W. C. (1963). An experimental investi-


gation of flow structure of the turbulent boundary layer. report no. md-8. Technical
report, Department of Mechanical Engineering, Stanford University.

Sahlin, A., Johansson, A. V., and Alfredsson, P. H. (1988). The possibility of drag
reduction by outer layer manipulators in turbulent boundary layers. Physics of
Fluids, 31(10):2814–2820.

Saric, W. S., Reed, H. L., and Kerschen, E. J. (2002). Boundary-layer receptivity to


freestream disturbances. Annual Review of Fluid Mechanics, 34:291–319.

Satake, S.-i. and Kasagi, N. (1996). Turbulence control with wall-adjacent thin layer
damping spanwise velocity fluctuations. International Journal of Heat and Fluid
Flow, 17(3):343–352.

Schlichting, H. (1933). Zur entstehung der turbulenz bei der plattenstomung. Nachr.
Ges. Wiss. Gottingen, pages 182–208.

Schlichting, H. (1960). Boundary Layer Theory. McGraw-Hill, New York, 4th edition.

Schmid, P. J. and Henningson, D. (1992). A new mechanism for rapid transition


involving a pair of oblique waves. Physics of Fluids A, 4(9):1986–1989.

Schmid, P. J. and Henningson, D. S. (2001). Stability and transition in shear flows,


volume 142 of Applied Mathematical Sciences. Springer-Verlag, New York.

Schoppa, W. and Hussain, F. (1998). Large-scale control strategy for drag reduction
in turbulent boundary layers. Physics of Fluids, 10(5 May):1049.

Schoppa, W. and Hussain, F. (2002). Coherent structure generation in near-wall


turbulence. Journal of Fluid Mechanics, (453):57–108.
154 BIBLIOGRAPHY

Schulze, T. P. (1999). A note on subharmonic instabilities. Physics of Fluids,


11(12):3573 – 3576.

Selvarajan, S., Tulapurkara, E., and Ram, V. V. (1999). Stability characteristics of


wavy walled channel flows. Physics of Fluids, 11(3):579–589.

Shen, L., Zhang, X., Yue, D. K., and Triantafyllou, M. S. (2003). Turbulent flow over
a flexible wall undergoing a streamwise travelling wave motion. Journal of Fluid
Mechanics, (484):197–221.

Singer, B. A., Reed, H. L., and Ferziger, J. H. (1989). The effects of streamwise
vortices on transition in the plane channel. Physics of Fluids A, 1(12):1960–1971.

Sirovich, L. and Karlsson, S. (1997). Turbulent drag reduction by passive means.


Nature, 338:753–755.

Smith, C. R. and Walker, J. D. A. (1998). Mechanisms of turbulent boundary layers:


vortex development and interactions, paper no. aiaa-98-2959. In 29th AIAA Fluid
Dynamics Conference, Albuquerque, NM.

Smits, A. J. and Delo, C. (2002). Self-sustaining mechanisms of wall turbulence.


In Reguera, D., Bonilla, L. L., and Rubi, J. M., editors, Coherent structures in
complex systems, Lecture notes in physics, pages 17–38. Springer, New York.

Sommerfeld, A. (1908). Ein beitrag zur hydrodynamischen erklarung der turbulenten


flussigkeitbewegungen. In Atti. del 4. Congr. Internat. dei Mat. III, pages 116 –
124, Roma.

Sonneveld, P., Wesseling, P., and de Zeeuw, P. M. (1985). Multigrid and conjugate
gradient methods as convergence acceleration techniques. In Paddon, D. J. and
Holstein, H., editors, Multigrid Methods for Integral and Differential Equations,
pages 117–167. Clarendon Press, Oxford.

Squire, H. B. (1933). On the stability for three-dimensional disturbances of viscous


fluid flow between parallel walls. Proc. Roy. Soc. Lond. Ser. A, 142:621–628.

Sugiyama, K., Kawamura, T., Takagi, S., and Matsumoto, Y. (2002). Numerical
simulations on drag reduction mechanism by microbubbles. In Proc. of the 3rd
Symp. on Smart Control of Turbulence, 2002, pages 129–138.

Sureshkumar, R., Beris, A. N., and Handler, R. A. (1997). Direct numerical simulation
of the turbulent channel flow of a polymer solution. Physics of Fluids, 9(3):743 –
755.
BIBLIOGRAPHY 155

Swearingen, J. D. and Blackwelder, Ron, F. (1987). The growth and breakdown of


streamwise vortices in the presence of a wall. Journal of Fluid Mechanics, 182:255
– 290.

Tardu, S. (2001). Active control of near-wall turbulence by local oscillating blowing.


Journal of Fluid Mechanics, (439):217–253.

Theodorsen, T. (1952). Mechanism of turbulence. Proc. Midwest. Conf. Fluid Mech.,


2nd, Columbus, Ohio, pages 1 – 18.

Theofilis, V. (1994). The discrete temporal eigenvalue spectrum of the generalised


hiemenz flow as solution of the orr-sommerfeld equation. Journal of Engineering
Mathematics, 28:241 – 259.

Tollmien, W. (1929). Uber die entstehung der turbulenz. Nachr. Ges. Wiss. Gottin-
gen, pages 21 – 44. (English translation NACA TM 609, 1931).

Toms, B. A. (1948). Some observations on the flow of linear polymer solutions through
straight tubes at large reynolds numbers. In Proc. First Int. Cong. Rheol., volume 2,
pages 135 – 141, North-Holland, Amsterdam.

Townsend, A. A. (1956). The structure of turbulent shear flow. Cambridge: Univ.


Press. 315 pp. 1st ed.

Vlachos, P. P. and Rediniotis, O. K. (2003). Active skin for turbulent drag reduction.
Technical report, NASA Langley Research Center, Flow Modelling and Control
Branch.

Vogel, M., Hirsa, A. H., and Lopez, J. M. (2003). Spatio-temporal dynamics of a


periodically driven cavity flow. Jounal of Fluid Mechanics, 478:197–226.

von Helmholtz, H. L. F. (1868). Monatsberichte der königlichen akademie der wis-


senschaften zu berlin. 215.

von Kerczek, C. (1982). Instability of oscillatory plane poiseuille flow. Jounal of Fluid
Mechanics, 116:91–114.

Waleffe, F. (1990). On the three-dimensional instability of strained vortices. Physics


of Fluids A, 2(1):76–80.

Waleffe, F. (1995). Hydrodynamic stability and turbulence: Beyond transients to a


self-sustaining process. Studies in Applied Mathematics, 95:319 – 343.

Waleffe, F. (1997). On a self-sustaining process in shear flows. Physics of Fluids,


9(4):883.
156 BIBLIOGRAPHY

Waleffe, F. and Kim, J. (1997). How streamwise rolls and streaks self-sustain in a
shear flow. In Panton, R. L., editor, Self-Sustaining Mechanisms of Wall Turbu-
lence, Advances in Fluid Mechanics, pages 309–332. Computational Mechanics Inc,
Billerica, MA, USA, 15 edition.

Weier, T., Fey, U., Gerbeth, G., Mutschke, G., and Avilov, V. (2000). Boundary
layer control by means of electromagnetic forces. ERCOFTAC Bulletin, 44:36–40.

Wright, M. and Nelson, P. (2001). Wind tunnel experiments on the optimization


of distributed suction for laminar flow control. In Proceedings of the Institution
of Mechanical Engineers, Part G: Journal of Aerospace Engineering, volume 215,
pages 343–354. Professional Engineering Publishing Ltd.

Wu, X. and Luo, J. (2003). Linear and nonlinear instabilities of a blasius boundary
layer perturbed by streamwise vortices. part 1. steady streaks. Journal of Fluid
Mechanics, (483):225–248.

Xu, J., Maxey, M. R., and Karniadakis, G. E. (2002). Numerical simulation of turbu-
lent drag reduction using micro-bubbles. Journal of Fluid Mechanics, (468):271–
281.

Xu, S., Rempfer, D., and Lumley, J. (2003). Turbulence over a compliant surface:
numerical simulation and analysis. Journal of Fluid Mechanics, 478:11–34.

Yu, X. and Liu, J. T. C. (1991). The secondary instability in goertler flow. Physics
of Fluids A, 3(8):1845 – 1847.

Zhou, J., Adrian, R., Balachandar, S., and Kendall, T. (1999). Mechanisms for
generating coherent packets of hairpin vortices in channel flow. Journal of Fluid
Mechanics, 387:353–396.
Appendix A

Matlab codes

A.1 stability

A.1.1 stability.m
function [growthrate] = stability(N, Re, alpha, beta, profile)

% Input conditions if run on own


if nargin == 0
N = 250; % Number of collocation points
Re = 10000; % Reynolds number... B = 1.7207876573;
alpha = 1; % Streamwise wavenumber
beta = 0; % Spanwise wavenumber
profile = ’poiseuille’; % Profile type: ’blasius’, ’poiseuille’
% or ’couette’

% make results directories as required


[sdir,messdir,messiddir] = mkdir(’results’);
[sdir,messdir,messiddir] = mkdir([’results/’, profile]);
end

% Define filename
filename = ([’results/’, profile,’/’, num2str(Re), ’_’, ...
num2str(alpha), ’_’, num2str(beta),’_’, num2str(N)]);

% Compute results only if they are not saved already


if exist([filename,’_eigvals.mat’]) == 0 % results not computed

157
158 APPENDIX A. MATLAB CODES

if strcmpi(profile,’blasius’) | strcmpi(profile,’cooke’)
% Convert Reynolds number accordingly
Re = Re/1.7207876573;
end

% set up spectral domain


k = [0:1:N]’;
j = k’;
etak = cos(pi*k/N); % Chebyshev collocation points in
% mapped domain
T = cos(k*pi*j/N); % Chebyshev transformation matrix (D0)
if beta == 0
SQ = 0;
else
SQ = 1;
end

% Solve eigenvalue problem in spectral space


% Make A and B matrices
[A, B, yk, uk, u1k, u2k] = makematrix(N, Re, ...
etak, T, alpha, beta, profile,SQ);

[eigenvectors, eigenvalues] = eig(A,B); % Solve eigenproblem

% Transform eigenvectors and eigvenvalues to physical space


z = zeros(N+1);
if SQ % both OS and SQ modes solved for
bigT = [T z
z T];
eigenvec = bigT*eigenvectors;
else % only OS modes solved for
eigenvec = T*eigenvectors;
end

% Create vector of eigenvalues


eigvals = diag(eigenvalues);

% Save eigenvalues to results file


A.1. STABILITY 159

save([filename, ’_eigvals.mat’],’eigvals’,’yk’,’uk’,’u1k’,’u2k’)

else % results already computed


eigenvec = []; % eigenvectors are not saved
load([filename,’_eigvals.mat’]) % load results from file
end

% Postprocessing of results
if nargin == 0
format long e % display results in double precision
[growthrate, majoreigs] = postprocess(N, yk, ...
eigvals, eigenvec,profile, uk, u1k, u2k, filename);
format % reset display precision to default
else
[growthrate, majoreigs] = postprocessnoplot(N, ...
yk, eigvals, eigenvec,profile, uk, u1k, u2k, filename);
end

A.1.2 makematrix.m

function [A, B, yk, uk, u1k, u2k] = ...


makematrix(N, Re, etak, T, alpha, beta, profile,SQ);

% Create velocity profiles and diff matrixes


% ===========================================================
switch profile
case ’blasius’
y_i = 10; ymax = 34.99;
[uk, u1k, u2k, yk, D1, D2, D3, D4] = ...
blasius(N,y_i,ymax,etak,T);
wk = zeros(N+1,1); w1k = wk; w2k = wk;
case ’cooke’
y_i = 7; ymax = 30; m = 0.2;
[uk, u1k, u2k, wk, w1k, w2k, yk, D1, D2, D3, D4] = ...
cooke(N,y_i,ymax,etak,T,m);
case ’poiseuille’
[uk, u1k, u2k, yk, D1, D2, D3, D4] = poiseuille(N,etak,T);
wk = zeros(N+1,1); w1k = wk; w2k = wk;
160 APPENDIX A. MATLAB CODES

case ’couette’
[uk, u1k, u2k, yk, D1, D2, D3, D4] = couette(N,etak,T);
wk = zeros(N+1,1); w1k = wk; w2k = wk;
otherwise
display(’Profile type unknown. Please check spelling.’)
end

% SETUP A and B MATRICES


% ===========================================================
% create U matrices
U = diag(uk);
U1 = diag(u1k);
U2 = diag(u2k);

% create W matrices
W = diag(wk);
W1 = diag(w1k);
W2 = diag(w2k);

% create A and B components


% ===========================================================
z = zeros(N+1);
k2 = alpha^2 + beta^2;
nabla2 = D2-k2*T;
nabla2squared = D4 - 2*k2*D2 + k2^2*T;

Los = (alpha*real(U)+beta*real(W))*nabla2-alpha*real(U2)*T ...


-beta*real(W2)*T+i/Re*nabla2squared;
Lsq = alpha*real(U)*T+beta*real(W)*T+i/Re*nabla2;

coupling = beta*real(U1)*T-alpha*real(W1)*T;
Imatrix = diag(ones(N+1,1))*T;

% Apply Boundary Conditions


% ===========================================================

% Create boundary condition matrix


BC = ones(N+1)*1.111111; % arbitrary constant
A.1. STABILITY 161

BC(1,:) = [ones(1,N+1)]; % v = 0 at y = 1
BC(N+1,:) = [2*mod(1:1:N+1,2)-1]; % v = 0 at y = -1
BC(2,:) = 1.^(1:N+1).*(0:N).^2; % dv/dy = 0 at y = 1
BC(N,:) = (-1).^(1:N+1).*(0:N).^2; % dv/dy = 0 at y = -1
BCpositions = find(BC~=1.111111); % indices BC values

% incorperate BCs into A and B components


% A components
Los(BCpositions)=BC(BCpositions);
Lsq(BCpositions)=BC(BCpositions);
coupling(BCpositions)=BC(BCpositions)*0;

% B components
nabla2(BCpositions)=BC(BCpositions)*0;
Imatrix(BCpositions)=BC(BCpositions)*0;

% compile matrices
% ===========================================================
if SQ
% if solving both OS and SQ modes
A = [ Los z
coupling Lsq];
B = alpha*[nabla2 z
z Imatrix];
else
% if only solving OS modes
A = [Los];
B = alpha*[nabla2];
end

A.1.3 postprocess.m

function [growthrate, majoreigs] = ...


postprocess(N, yk, eigvals, eigenvec,profile, uk, u1k, u2k, ...
filename)

% Plotting velocity profile


figure(1)
162 APPENDIX A. MATLAB CODES

set(gcf,’position’,[300, 500, 400,300],’paperunits’, ...


’centimeters’,’paperposition’,[1 1 10 8])
axes(’position’,[0.1300 0.1300 0.7750 0.8150],’box’,’on’, ...
’fontsize’,8)
hold on
plot(real(uk),yk,’-k’,’linewidth’,1.5)
plot(real(u1k),yk,’--k’,’linewidth’,1)
plot(real(u2k),yk,’:k’,’linewidth’,0.5)
xlabel(texlabel(’U/U_inf’))
ylabel(’nondimensional height’)
temp = axis; axis([temp(1), temp(2)-0.01*temp(2),temp(3), temp(4)])
% above line is required sometimes to tidy plot
legend(texlabel(’U’),texlabel(’dU/dy’),texlabel(’d^{2}U/dy^2’),2)
% saving

savename = [’results/’, profile, ’/profile.eps’];


print(’-deps2’,’-noui’,’-loose’,savename);

% Set relevant plots limits on eigenvalues


if strcmpi(profile,’blasius’) | strcmpi(profile,’cooke’) | ...
strcmpi(profile,’poiseuille’)
axislimits = [0 1.1 -1.0 0.1];
elseif strcmpi(profile,’couette’)
axislimits = [-1.1 1.1 -2.0 0.1];
else
display(’Profile type unknown. Please check spelling.’)
end

% plotting eigenvalues
% ===========================================================
figure(2)
set(gcf,’position’,[300, 500, 400,300],’paperunits’, ...
’centimeters’,’paperposition’,[1 1 10 8])
axes(’position’,[0.1300 0.1100 0.7750 0.8150],’box’,’on’, ...
’fontsize’,8)
hold on
patch([axislimits(1) axislimits(2) axislimits(2) ...
axislimits(1)],[0 0 axislimits(4) axislimits(4)], ...
A.1. STABILITY 163

[0.9 0.9 0.9],’EdgeColor’,’none’)


line([-100 100],[0 0],’linestyle’,’--’,’color’,’k’)
plot(real(eigvals),imag(eigvals),’k.’)
axis(axislimits)
xlabel(texlabel(’c_r’))
ylabel(texlabel(’c_i’))

% Find unstable eigenmodes and plot major eigenvectors


% ====================================================
growthrate = 0;
majoreigs = 0;
for index = 1:length(eigvals)
if abs(imag(eigvals(index))) < 50 & real(eigvals(index)) > 0 ...
& real(eigvals(index)) < 0.9
majoreigstemp(index,1) = eigvals(index);
if imag(eigvals(index)) > 0
figure(2)
plot(real(eigvals(index)),imag(eigvals(index)),’ko’)
end
end
end
figure(2)
savename = [filename, ’_eigenvalues.eps’];
print(’-deps2’,’-noui’,’-loose’,savename);

% clean up
majoreigstemp = majoreigstemp(find(majoreigstemp ~= 0 ));
majoreigssortedi = flipud(sort(imag(majoreigstemp)));
for index = 1:length(majoreigstemp)
index2 = find(imag(majoreigstemp) == majoreigssortedi(index,1));
index2 = index2(1);
majoreigs(index,1) = majoreigstemp(index2,1);
end
majoreigs % display major eigenvalues

% find maximum growthrate


growthrate = majoreigs(find(imag(majoreigs)==max(imag(majoreigs))));
164 APPENDIX A. MATLAB CODES

if isempty(eigenvec) == 0
% plot most unstable eigenvector
figure(3)
set(gcf,’position’,[300, 500, 400,300],’paperunits’, ...
’centimeters’,’paperposition’,[1 1 10 8])
axes(’position’,[0.1300 0.1100 0.7750 0.8150],’box’,’on’, ...
’fontsize’,8)
hold on
plot(real(eigenvec(1:N+1,find(eigvals == majoreigs(1)))),yk,’k-’)
plot(imag(eigenvec(1:N+1,find(eigvals == majoreigs(1)))),yk,’k--’)
ylabel(’y’);
xlabel(’magnitude’);
% saving
savename = [filename, ’_unstable_eigenvetors.eps’];
print(’-deps2’,’-noui’,’-loose’,savename);
end

A.1.4 cheb.m

% CHEB compute D = Chebyshev differentiation matrix

function [D] = cheb(N)


c = [2; ones(N-1,1); 2].*(-1).^(0:N)’;
D=zeros(N+1,N+1);
for i = 0:N
for j = 1:N
if i == 0
D(i+1,2*j)=2*j-1;
else
A.1. STABILITY 165

D(i+1,i+2*j)=2*(i+2*j-1);
end
end
end
D = D(1:N+1,1:N+1);

A.1.5 couette.m
function [uk, u1k, u2k, yk, D1, D2, D3, D4] = couette(N,etak,T);
% Couette profile

% Compute velocity profiles


yk = etak;
uk = -yk;
u1k = -1*ones(N+1,1);
u2k = zeros(N+1,1);

% Compute Spectral Chebyshev Differention matrices


D = cheb(N);
D1 = T*D;
D2 = T*D^2;
D3 = T*D^3;
D4 = T*D^4;

A.1.6 poiseuille.m
function [uk, u1k, u2k, yk, D1, D2, D3, D4] = poiseuille(N,etak,T);
% Poiseuille

% Compute velocity profiles


yk = etak;
uk = -(yk).^2+1;
u1k = -2*(yk);
u2k = -2*ones(N+1,1);
166 APPENDIX A. MATLAB CODES

% Compute Spectral Chebyshev Differention matrices


D = cheb(N);
D1 = T*D;
D2 = T*D^2;
D3 = T*D^3;
D4 = T*D^4;

A.1.7 blasius.m
function [uk, u1k, u2k, yk, D1, D2, D3, D4] = blasius(N,y_i,ymax,etak,T);
% Blasius boundary layer profile

% mapping for Blasius


a = y_i*ymax/(ymax-2*y_i);
b = 1+2*a/ymax;
yk = a*(1+etak)./(b-etak);

% Create Spectral Chebyshev differentiation matrices


[D1,D2,D3,D4] = chebtransform(N,a,b,yk,T);

% Compute solution;
solution = solveblasiusODE(ymax);
uk = deval(solution,yk,2)’;
u1k = deval(solution,yk,3)’;
u2k = D1*inv(T)*u1k;

A.1.8 chebtransform.m
% CHEB compute D = Chebyshev differentiation matrix

function [D1,D2,D3,D4] = chebtransform(N,a,b,yk,T)


% Make basic D matrix
c = [2; ones(N-1,1); 2].*(-1).^(0:N)’;
D=zeros(N+1,N+1);
for i = 0:N
for j = 1:N
if i == 0
D(i+1,2*j)=2*j-1;
A.1. STABILITY 167

else
D(i+1,i+2*j)=2*(i+2*j-1);
end
end
end
D = D(1:N+1,1:N+1);

% Calculate derivatives due to mapping


eta1 = diag(a.*(b+1)./(yk+a).^2);
eta2 = diag(-2.*a.*(b+1)./(yk+a).^3);
eta3 = diag(6.*a*(b+1)./(yk+a).^4);
eta4 = diag(-24.*a.*(b+1)./(yk+a).^5);

% Create derivative matrices


D1 = eta1*T*D;
D2 = eta1^2*T*D^2 + eta2*T*D;
D3 = eta1^3*T*D^3 + 3*eta1*eta2*T*D^2 + eta3*T*D;
D4 = eta1^4*T*D^4 + 6*eta1^2*eta2*T*D^3 + ...
(3*eta2^2 + 4*eta1*eta3)*T*D^2 + eta4*T*D;

A.1.9 solveblasiusODE.m

function [sol] = solveblasiusODE(ymax)


% blasius equation solver

% Solve for the boundary layer profile


solinit = bvpinit(linspace(0,ymax+0.01,30),[1 0 0]);
sol = bvp4c(@BlasiusODE,@BlasiusBC,solinit);

% ODE system
function dydx = BlasiusODE(x,y)
dydx = [ y(2)
y(3)
-y(1)*y(3)];
168 APPENDIX A. MATLAB CODES

% Boundary conditions
function res = BlasiusBC(ya,yinf)
res = [ ya(1)
ya(2)
yinf(2)-1];

A.2 FIDAP read-file generator


% FIDAP readfile generator for travelling wave simulation
% Reuben Rusk, 21-07-03

% Domain conditions ====================================


amp = 0.05;
freq = 1;
period = 1/freq;
time = 20;
dt = 0.01;
numsteps = time / dt;
intervals = 24*3;

interface_nodes = [1 3:1:(intervals+1) 2];


nodes = length(interface_nodes);

% % Readfile generation ====================================


%
fid=fopen(’wavemake.FDREAD’,’wt’);
fprintf(fid,’FI-GEN( ELEM = 1, POIN = 1, CURV = 1, SURF = 1, ...
NODE = 0, MEDG = 1, MLOO = 1, MFAC = 1, BEDG = 1, SPAV = 1, ...
MSHE = 1, MSOL = 1, COOR = 1 )’);
fprintf(fid,’\nPOINT( ADD, COOR, X = 0, Y = 0 )’);
fprintf(fid,’\nPOINT( ADD, COOR, X = 0, Y = 1 )’);
fprintf(fid,’\nPOINT( ADD, COOR, X = 1, Y = 0 )’);
fprintf(fid,’\nPOINT( ADD, COOR, X = 1, Y = 1 )’);
fprintf(fid,’\nPOINT( SELE, ID, WIND = 1 )’);
fprintf(fid,’\n1’);
A.2. FIDAP READ-FILE GENERATOR 169

fprintf(fid,’\n3’);
fprintf(fid,’\nCURVE( ADD )’);
fprintf(fid,’\nPOINT( SELE, ID, WIND = 1 )’);
fprintf(fid,’\n2’);
fprintf(fid,’\n4’);
fprintf(fid,’\nCURVE( ADD )’);
fprintf(fid,’\nPOINT( SELE, ID, WIND = 1 )’);
fprintf(fid,’\n1’);
fprintf(fid,’\n2’);
fprintf(fid,’\nCURVE( ADD )’);
fprintf(fid,’\nPOINT( SELE, ID, WIND = 1 )’);
fprintf(fid,’\n3’);
fprintf(fid,’\n4’);
fprintf(fid,’\nCURVE( ADD )’);
fprintf(fid,’\nCURVE( SELE, ID, WIND = 1 )’);
fprintf(fid,’\n2’);
fprintf(fid,’\n1’);
fprintf(fid,’\nSURFACE( ADD, CURV )’);
fprintf(fid,’\nCURVE( SELE, ID, WIND = 1 )’);
fprintf(fid,’\n1’);
fprintf(fid,’\n2’);
fprintf(fid,[’\nMEDGE( ADD, SUCC, INTE = ’, num2str(intervals),’, ...
RATI = 0, 2RAT = 0, PCEN = 0 )’]);
fprintf(fid,’\nCURVE( SELE, ID, WIND = 1 )’);
fprintf(fid,’\n4’);
fprintf(fid,’\n3’);
fprintf(fid,[’\nMEDGE( ADD, SUCC, INTE = ’, num2str(intervals),’, ...
RATI = 1.025, 2RAT = 0, PCEN = 0 )’]);
fprintf(fid,’\nCURVE( SELE, ID, WIND = 1 )’);
fprintf(fid,’\n2’);
fprintf(fid,’\n4’);
fprintf(fid,’\n1’);
fprintf(fid,’\n3’);
fprintf(fid,’\nMLOOP( ADD, PAVE )’);
fprintf(fid,’\nSURFACE( SELE, ID, WIND = 1 )’);
fprintf(fid,’\n1’);
fprintf(fid,’\nUTILITY( HIGH = 9 )’);
fprintf(fid,’\nMLOOP( SELE, ID, WIND = 1 )’);
170 APPENDIX A. MATLAB CODES

fprintf(fid,’\n1’);
fprintf(fid,’\nUTILITY( HIGH = 3 )’);
fprintf(fid,’\nMFACE( ADD )’);
fprintf(fid,’\nMFACE( SELE, ID, WIND = 1 )’);
fprintf(fid,’\n1’);
fprintf(fid,’\nMFACE( MESH, MAP, ENTI = "fluid" )’);
fprintf(fid,’\nEND( )’);
fprintf(fid,’\nFI-BC( )’);
fprintf(fid,’\nBGADD( SELE, EDGE, INCL, ID, WIND = 1 )’);
fprintf(fid,’\n1’);
fprintf(fid,’\nBGADD( ADD, EDGE, ENTI = "upperbound", INCL )’);
fprintf(fid,’\nBGADD( SELE, EDGE, INCL, ID, WIND = 1 )’);
fprintf(fid,’\n4’);
fprintf(fid,’\nBGADD( ADD, EDGE, ENTI = "leftbound", INCL )’);
fprintf(fid,’\nBGADD( SELE, EDGE, INCL, ID, WIND = 1 )’);
fprintf(fid,’\n2’);
fprintf(fid,’\nBGADD( ADD, EDGE, ENTI = "rightbound", INCL )’);
fprintf(fid,’\nBGADD( SELE, EDGE, INCL, ID, WIND = 1 )’);
fprintf(fid,’\n3’);
fprintf(fid,’\nBGADD( ADD, EDGE, ENTI = "interface", INCL )’);
fprintf(fid,’\nEND( )’);

fprintf(fid,’\nTITLE’);
fprintf(fid,’\n2D flow induced by a travelling wave’);

fprintf(fid,’\nFIPREP’);

% PROBLEM SETUP

fprintf(fid,’\nPROBLEM (2-D, INCOMPRESSIBLE, TRANSIENT, ...


LAMINAR, NONLINEAR, NEWTONIAN, MOMENTUM, ISOTHERMAL, ...
STRUCTURAL, REMESHING)’);
fprintf(fid,’\nEXECUTION (NEWJOB)’);
fprintf(fid,’\nDATAPRINT (CONTROL)’);
fprintf(fid,’\nPRINTOUT (NONE)’);
fprintf(fid,’\nSOLUTION (SEGREGATED = 200, VELCONV = 0.1e-2, ...
CGS = 400, NCGCONV = 0.1e-2, PRECONDITION = 11, NOLINESRCH, ...
PPROJECTION, NORMALSTRESS = 30, SURFCONV = 0.001, SCHA = 0 )’);
A.2. FIDAP READ-FILE GENERATOR 171

fprintf(fid,’\nSTRUCTURALOPTIONS (LARG, UNSTEADY_TERM, VELOCITY)’);


fprintf(fid,[’\nTIMEINTEGRATION (BACK, NSTE = ’, ...
num2str(numsteps),’, TSTA = 0, DT = ’, num2str(dt),’, ...
FIXE, NEWM = 0.5 )’]);
fprintf(fid,’\nPRESSURE (MIXED, CONTINUOUS )’);
fprintf(fid,’\nPOSTPROCESS (NBLOCKS = 1)’);
fprintf(fid,[’\n1 ’, num2str(numsteps),’ 2’]);
fprintf(fid,’\nOPTIONS ( STRESS-DIVERGENCE, UPWINDING )’);
fprintf(fid,’\nNONDIMENSION ( ON, LENGTH = 1, VELOCITY = 1, ...
MDENS= "fluid", MVISC= "fluid" )’);
fprintf(fid,’\nUPWINDING (STREAMLINE)’);
fprintf(fid,’\n1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1’);
fprintf(fid,’\nRELAXATION (HYBRID)’);

% Boundary conditions
% ======

% physical constraints
fprintf(fid,’\nBCNODE (DX, ENTI = "leftbound", ZERO)’);
fprintf(fid,’\nBCNODE (DX, ENTI = "rightbound", ZERO)’);
fprintf(fid,’\nBCNODE (DX, ENTI = "upperbound", ZERO)’);
fprintf(fid,’\nBCNODE (DY, ENTI = "upperbound", ZERO)’);
fprintf(fid,’\nBCNODE (UX, ENTI = "upperbound", ZERO)’);
fprintf(fid,’\nBCNODE (UX, ENTI = "interface", ZERO)’);
fprintf(fid,’\nBCNODE (DX, ENTI = "interface", ZERO)’);
fprintf(fid,[’\nBCPERIODIC (ALL, ENTI, REFE = "leftbound", PERI = ...
"rightbound", EXCL,R1NO = 1, R2NO = ’, num2str(intervals+2), ’, ...
P1NO = 2, P2NO = ’, num2str(intervals+3),’ )’]);
fprintf(fid,’\nDENSITY (SET = "fluid", CONS = 1 )’);
fprintf(fid,’\nVISCOSITY (SET = "fluid", CONS = 0.01, TWO-EQUATION )’);

%entities
fprintf(fid,’\nENTITY(NAME = "fluid", FLUI, MDEN = "fluid", ...
MVIS = "fluid" )’);
fprintf(fid,’\nENTITY(NAME = "upperbound", PLOT )’);
fprintf(fid,’\nENTITY(NAME = "leftbound", PLOT )’);
fprintf(fid,’\nENTITY(NAME = "rightbound", PLOT )’);
fprintf(fid,’\nENTITY(NAME = "interface", MEMBRANE, MOVING )’);
172 APPENDIX A. MATLAB CODES

for i = 1:length(interface_nodes)
x = (i-1)/(nodes-1);
if i < length(interface_nodes)
fprintf(fid,[’\nTMFU (SET =’,num2str(i),’, SINU, ACOE = ’, ...
num2str(amp),’, T0 =’, num2str(x*period), ...
’ , REFV = ’, num2str(period),’)’]);
fprintf(fid,[’\nBCNODE (DY, NODE = ’, ...
num2str(interface_nodes(i)),’, FSUBROUTINE, CURV = ’, ...
num2str(i),’, FACT = 1.0)’]);
else
fprintf(fid,[’\nBCNODE (DY, NODE = ’, ...
num2str(interface_nodes(i)),’, FSUBROUTINE, CURV = ’, ...
num2str(1),’, FACT = 1.0)’]);
end

end

fprintf(fid,’\nEND’);
fprintf(fid,’\nCREATE(FISOLV)’);
fclose(fid);
Appendix B

Additional wave-forcing results

173
174 APPENDIX B. ADDITIONAL WAVE-FORCING RESULTS

(a) (b)
nondimensional height, y

0.5 0.5

nondimensional height, y
0.4 0.4

0.3 0.3

0.2 0.2

0.1
0.1
0
0
−1 0 1 2 3 4 5 6 7
0 0.2 0.4 0.6 0.8 1
−4
x 10

(c) (d)
0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3
0.3
0.2
0.2
0.1
0.1
0
0
−5 −4 −3 −2 −1 0 0 0.2 0.4 0.6 0.8 1
Streamfunction −3
x 10 nondimensional spanwise distance, z

(e) (f)
0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
−0.03 −0.02 −0.01 0 0.01 0.02 0.03 −0.04 −0.03 −0.02 −0.01 0 0.01 0.02 0.03 0.04
nondimensional spanwise velocity, W nondimensional vertical velocity, V

Figure B.1: Re = 102 , wave amplitude is 0.005. Profiles of the analytical streamfunc-
tion ψ in y over one period at fixed x position (a,c), and contours of the deviation of
ψ from the time-averaged mean (b,d) for a Reynolds number of 102 . The real part
is given in plots (a) and (b), and the imaginary part is shown in plots (c) and (d).
Plots (e) and (f) show profiles over time of spanwise and vertical velocity, respectively.
Wave motion is from left to right.
175

(a) (b)
nondimensional height, y

0.5 0.5

nondimensional height, y
0.4 0.4

0.3 0.3

0.2 0.2

0.1
0.1
0
0
−0.5 0 0.5 1 1.5 2 2.5
0 0.2 0.4 0.6 0.8 1
−3
x 10

(c) (d)
0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3
0.3
0.2
0.2
0.1
0.1
0
0
−0.02 −0.015 −0.01 −0.005 0 0 0.2 0.4 0.6 0.8 1
Streamfunction nondimensional spanwise distance, z

(e) (f)
0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
−0.1 −0.05 0 0.05 0.1 −0.2 −0.15 −0.1 −0.05 0 0.05 0.1 0.15
nondimensional spanwise velocity, W nondimensional vertical velocity, V

Figure B.2: Re = 102 , wave amplitude is 0.020. Profiles of the analytical streamfunc-
tion ψ in y over one period at fixed x position (a,c), and contours of the deviation of
ψ from the time-averaged mean (b,d) for a Reynolds number of 102 . The real part
is given in plots (a) and (b), and the imaginary part is shown in plots (c) and (d).
Plots (e) and (f) show profiles over time of spanwise and vertical velocity, respectively.
Wave motion is from left to right.
176 APPENDIX B. ADDITIONAL WAVE-FORCING RESULTS

(a) (b)
nondimensional height, y

0.5 0.5

nondimensional height, y
0.4 0.4

0.3 0.3

0.2 0.2

0.1
0.1
0
0
−1 0 1 2 3 4 5
0 0.2 0.4 0.6 0.8 1
−3
x 10

(c) (d)
0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3
0.3
0.2
0.2
0.1
0.1
0
0
−0.035 −0.03 −0.025 −0.02 −0.015 −0.01 −0.005 0 0 0.2 0.4 0.6 0.8 1
Streamfunction nondimensional spanwise distance, z

(e) (f)
0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
−0.2 −0.15 −0.1 −0.05 0 0.05 0.1 0.15 0.2 −0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3
nondimensional spanwise velocity, W nondimensional vertical velocity, V

Figure B.3: Re = 102 , wave amplitude is 0.035. Profiles of the analytical streamfunc-
tion ψ in y over one period at fixed x position (a,c), and contours of the deviation of
ψ from the time-averaged mean (b,d) for a Reynolds number of 102 . The real part is
given in plots (a) and (b), and the imaginary part is shown in plots (c) and (d). Plots
(e) and (f) show profiles over time of spanwise and vertical velocity, respectively.
177

(a) contour spacing = 0.03


0.5
nondimensional height, y

0.4

0.3

0.2

0.1

0 0.2 0.4 0.6 0.8 1

(b) contour spacing = 0.06


0.5
nondimensional height, y

0.4

0.3

0.2

0.1

0 0.2 0.4 0.6 0.8 1

(c) contour spacing = 0.1


0.5
nondimensional height, y

0.4

0.3

0.2

0.1

0 0.2 0.4 0.6 0.8 1


nondimensional spanwise distance, z

Figure B.4: Flow field for Re = 101 and wave amplitude of 0.050, shown by contours
of spanwise velocity (a), vertical velocity (b) and pressure (c). Numerical solution is
shown in solid (positive) and dashed (negative) lines, the analytical solution is shown
by dot-dashed (positive) and dotted (negative) lines. Wave motion is from left to
right.
178 APPENDIX B. ADDITIONAL WAVE-FORCING RESULTS

(a) (b)
nondimensional height, y

0.5 0.5

nondimensional height, y
0.4 0.4

0.3 0.3

0.2 0.2

0.1
0.1
0
0
−1 0 1 2 3 4 5 6 7
0 0.2 0.4 0.6 0.8 1
−3
x 10

(c) (d)
0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3
0.3
0.2
0.2
0.1
0.1
0
0
−0.05 −0.04 −0.03 −0.02 −0.01 0 0 0.2 0.4 0.6 0.8 1
Streamfunction nondimensional spanwise distance, z

(e) (f)
0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
−0.2 −0.15 −0.1 −0.05 0 0.05 0.1 0.15 −0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4
nondimensional spanwise velocity, W nondimensional vertical velocity, V

Figure B.5: Re = 101 , wave amplitude is 0.050. Profiles of the analytical streamfunc-
tion ψ in y over one period at fixed x position (a,c), and contours of the deviation of
ψ from the time-averaged mean (b,d) for a Reynolds number of 102 . The real part is
given in plots (a) and (b), and the imaginary part is shown in plots (c) and (d). Plots
(e) and (f) show profiles over time of spanwise and vertical velocity, respectively.
179

(a) (b)
nondimensional height, y

0.5 0.5

nondimensional height, y
0.4 0.4

0.3 0.3

0.2 0.2

0.1
0.1
0
0
−1 0 1 2 3 4 5 6 7
0 0.2 0.4 0.6 0.8 1
−3
x 10

(c) (d)
0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3
0.3
0.2
0.2
0.1
0.1
0
0
−0.05 −0.04 −0.03 −0.02 −0.01 0 0 0.2 0.4 0.6 0.8 1
Streamfunction nondimensional spanwise distance, z

(e) (f)
0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
−0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 −0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4
nondimensional spanwise velocity, W nondimensional vertical velocity, V

Figure B.6: Re = 102 , wave amplitude is 0.050. Profiles of the analytical streamfunc-
tion ψ in y over one period at fixed x position (a,c), and contours of the deviation of
ψ from the time-averaged mean (b,d) for a Reynolds number of 102 . The real part is
given in plots (a) and (b), and the imaginary part is shown in plots (c) and (d). Plots
(e) and (f) show profiles over time of spanwise and vertical velocity, respectively.
180 APPENDIX B. ADDITIONAL WAVE-FORCING RESULTS

(a) (b)
nondimensional height, y

0.5 0.5

nondimensional height, y
0.4 0.4

0.3 0.3

0.2 0.2

0.1
0.1
0
0
−0.5 0 0.5 1 1.5 2 2.5 3 3.5 4
0 0.2 0.4 0.6 0.8 1
−3
x 10

(c) (d)
0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3
0.3
0.2
0.2
0.1
0.1
0
0
−0.05 −0.04 −0.03 −0.02 −0.01 0 0 0.2 0.4 0.6 0.8 1
Streamfunction nondimensional spanwise distance, z

(e) (f)
0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
−0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4 −0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4
nondimensional spanwise velocity, W nondimensional vertical velocity, V

Figure B.7: Re = 103 , wave amplitude is 0.050. Profiles of the analytical streamfunc-
tion ψ in y over one period at fixed x position (a,c), and contours of the deviation of
ψ from the time-averaged mean (b,d) for a Reynolds number of 102 . The real part is
given in plots (a) and (b), and the imaginary part is shown in plots (c) and (d). Plots
(e) and (f) show profiles over time of spanwise and vertical velocity, respectively.
181

(a) (b)
nondimensional height, y

0.5 0.5

nondimensional height, y
0.4 0.4

0.3 0.3

0.2 0.2

0.1
0.1
0
0
0 0.5 1 1.5 2
0 0.2 0.4 0.6 0.8 1
−3
x 10

(c) (d)
0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3
0.3
0.2
0.2
0.1
0.1
0
0
−0.05 −0.04 −0.03 −0.02 −0.01 0 0 0.2 0.4 0.6 0.8 1
Streamfunction nondimensional spanwise distance, z

(e) (f)
0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
−0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4 −0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4
nondimensional spanwise velocity, W nondimensional vertical velocity, V

Figure B.8: Re = 104 , wave amplitude is 0.050. Profiles of the analytical streamfunc-
tion ψ in y over one period at fixed x position (a,c), and contours of the deviation of
ψ from the time-averaged mean (b,d) for a Reynolds number of 102 . The real part is
given in plots (a) and (b), and the imaginary part is shown in plots (c) and (d). Plots
(e) and (f) show profiles over time of spanwise and vertical velocity, respectively.
182 APPENDIX B. ADDITIONAL WAVE-FORCING RESULTS

(a) contour spacing = 0.06


0.5
nondimensional height, y

0.4

0.3

0.2

0.1

0 0.2 0.4 0.6 0.8 1

(b) contour spacing = 0.06


0.5
nondimensional height, y

0.4

0.3

0.2

0.1

0 0.2 0.4 0.6 0.8 1

(c) contour spacing = 0.02


0.5
nondimensional height, y

0.4

0.3

0.2

0.1

0 0.2 0.4 0.6 0.8 1


nondimensional spanwise distance, z

Figure B.9: Flow field for Re = 105 and wave amplitude of 0.050, shown by contours
of spanwise velocity (a), vertical velocity (b) and pressure (c). Numerical solution is
shown in solid (positive) and dashed (negative) lines, the analytical solution is shown
by dot-dashed (positive) and dotted (negative) lines. Wave motion is from left to
right.
183

(a) (b)
nondimensional height, y

0.5 0.5

nondimensional height, y
0.4 0.4

0.3 0.3

0.2 0.2

0.1
0.1
0
0
−2 0 2 4 6 8 10
0 0.2 0.4 0.6 0.8 1
−4
x 10

(c) (d)
0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3
0.3
0.2
0.2
0.1
0.1
0
0
−0.05 −0.04 −0.03 −0.02 −0.01 0 0 0.2 0.4 0.6 0.8 1
Streamfunction nondimensional spanwise distance, z

(e) (f)
0.5 0.5
nondimensional height, y

nondimensional height, y

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
−0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4 −0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4
nondimensional spanwise velocity, W nondimensional vertical velocity, V

Figure B.10: Re = 105 , wave amplitude is 0.050. Profiles of the analytical stream-
function ψ in y over one period at fixed x position (a,c), and contours of the deviation
of ψ from the time-averaged mean (b,d) for a Reynolds number of 102 . The real part
is given in plots (a) and (b), and the imaginary part is shown in plots (c) and (d).
Plots (e) and (f) show profiles over time of spanwise and vertical velocity, respectively.

View pu

You might also like