You are on page 1of 274

Double-Averaged Open-Channel

Flow over Regular Rough Beds

Lorna J Campbell
MEng. (Aberdeen)

A thesis presented for the degree of


Doctor of Philosophy
at the

DEPARTMENT OF ENGINEERING
UNIVERSITY OF ABERDEEN

August 2005
DECLARATION

This work has been entirely composed by the author and has not been
accepted in any previous application for a degree. The work has been
undertaken by the candidate and all sources of information have been
acknowledged.

Lorna Campbell

PhD Thesis i Declaration


L J Campbell
ABSTRACT
Much research effort has been invested in clarifying the velocity and turbulence
characteristics of hydraulically rough open-channel flows at levels beyond the direct
influence of the bed topography. The majority of this analysis has been founded on
the temporally-averaged momentum equations (the Reynolds-Averaged Navier-
Stokes equations, RANS). However, the structure of shallow open-channel flows
(i.e. those in which the roughness height is of the same order of magnitude as the
flow depth) over hydraulically rough beds, together with the near-bed region of
deeper open-channel flows with rough beds awaits clarification. This is because in
these regions the effect of surface roughness on the flow properties cannot be
neglected, so that the information gained from a RANS approach is dependent on the
measurement location relative to the bed topography. To overcome this difficulty,
recent advances in the analysis of such flows have suggested that temporal averaging
of the momentum equations should be supplemented with spatial averaging (in
planes or volumes orientated parallel to the mean bed), thus yielding the double-
averaged momentum equations.

The purpose of this project was to conduct an extensive programme of laboratory-


based experiments to gauge the applicability of the double-averaged momentum
equations for shallow open-channel flows over geometrically simple rough beds. The
bed roughness comprised periodically positioned, transverse square bars that
extended across the whole width of the hydraulic flume. The centre-to-centre bar
pitch was varied between experiments, thus generating a broad range of near-bed
flow patterns for each of the three flow depths studied.

Proper assessment of the double-averaged equations for rough open-channel flow


requires comprehensive measurement of fluid velocities as they vary in both time
and space. Therefore, detailed Particle Image Velocimetry (PIV) measurements have
been obtained of the streamwise and bed-normal velocity components, including
their variation in both temporal and spatial domains which permits accurate
assessment of individual terms in the double-averaged flow equations.

The results show that double-averaging is a powerful tool for the analysis of
hydraulically rough flows. For example, for a range of isolated flow types, the
vertical distribution of the double-averaged streamwise velocity follows a simple
linear trend between the bars. Quadrant analysis has been successfully applied to the
spatial, rather than temporal, fluctuations of velocity components for the first time,
and double-averaging analysis has revealed areas of intense local momentum transfer
despite negligible global momentum exchange over the averaging window. The
effect of areas of flow separation and reattachment, especially the location of the
reattachment point, is evident throughout the results. In particular, this thesis reports
the discovery of an instability at the transition between wake interference and
isolated roughness flows at which the overall properties of the flow are dramatically
altered by the interaction between the fluid and the wall roughness.

The results form a comprehensive dataset that may be used for the development and
validation of numerical models over simple roughness types, whilst the analysis
significantly advances understanding of the fundamental mechanisms underlying the
interaction of a fluid flow and a rough surface.

PhD Thesis ii Abstract


L J Campbell
ACKNOWLEDGEMENTS
Jim Adamson
Denis Bain
Philip Benzie
Sylvia Carnie
Graeme Clubb
Peter Davies
Jeff Doucette
EPSRC Loan Pool
Heather Flett
Rab Fraser
Mike Gallagher
Derek Goring
Yakun Guo
John Heald
Angela Henderson
Nial Horton
Ranko Lazic
Bill Lowson
Ian McEwan
Iain McKinnon
Costa Manes
Ravi Manoharan
Alan Melville
Nigel Money
Claire Mulcock
Tim Nicholls
Vlad Nikora
Tom ODonoghue
Dubravka Pokrajac
John Polanski
Ann Rice
Kate Rudman
Anne Shipley
Keith Simpson
Eddie Stephen
Simon Tait
Phil Waterworth
John Watson
Brian Willetts
Scott Wright

THANK YOU

PhD Thesis iii Acknowledgements


L J Campbell
TABLE OF CONTENTS
List of Figures ............................................................................................................vii
List of Tables.............................................................................................................xiii
Nomenclature ............................................................................................................xiv
Chapter Marker Information ....................................................................................xvii

1.0 INTRODUCTION ............................................................................................... 1


1.1 AIMS AND OBJECTIVES ..................................................................................... 3
1.2 THESIS OUTLINE ............................................................................................... 4

2.0 LITERATURE REVIEW ................................................................................... 5


2.1 INTRODUCTION ................................................................................................. 5
2.2 BOUNDARY LAYER THEORY ............................................................................. 6
2.2.1 Hydraulically Smooth and Hydraulically Rough Flow.......................... 7
2.2.2 Open-Channel Flow ............................................................................. 10
2.3 FLUID FLOW OVER TRANSVERSE SQUARE BARS............................................. 15
2.3.1 General Overview of Square Bar Roughness Studies .......................... 16
2.3.1.1 Research Themes ..................................................................... 27
2.3.1.2 Closed-Channel versus Open-Channel Flows......................... 27
2.3.1.3 Velocity Measurement Techniques .......................................... 28
2.3.1.4 Velocity Measurement Locations............................................. 29
2.3.1.5 Roughness Spacing .................................................................. 30
2.3.2 Transverse Strip Roughness Classifications ........................................ 31
2.3.2.1 The Strip Roughness Terminology of Morris (1955) ............... 31
2.3.2.2 The Strip Roughness Terminology of Perry et al. (1969) ........ 34
2.3.3 Streamlines and Reattachment Length for Flows over Square Bars .... 36
2.3.4 The Resistance of Transverse Square Bar Roughness ......................... 39
2.4 DOUBLE-AVERAGING METHODOLOGY ........................................................... 43
2.4.1 The Development of Double-Averaging Theory ................................. 44
2.4.2 Double-Averaged Open-Channel Flow................................................ 46
2.4.2.1 Layers in Double-Averaged Open-Channel Flow ................... 46
2.4.3 Distribution of Double-Averaged Velocity in the Near-Bed Region... 50
2.4.4 Distribution of Double-Averaged Fluid Stress..................................... 53
2.5 CHAPTER CLOSURE ........................................................................................ 55

3.0 ANALYTICAL METHODOLOGY:


THE DOUBLE-AVERAGED FLOW EQUATIONS..................................... 57
3.1 INTRODUCTION ............................................................................................... 57
3.2 THE NAVIER-STOKES EQUATIONS .................................................................. 58
3.3 THE REYNOLDS-AVERAGED NAVIER-STOKES EQUATIONS (RANS) .............. 59
3.4 THE SPATIALLY-AVERAGED RANS EQUATIONS ............................................ 61
3.4.1 Flow Above the Roughness Crests....................................................... 63
3.4.2 Flow Below the Roughness Crests....................................................... 64
3.4.2.1 Pressure Field about a Series of Fences:
Introducing the Form Drag Term............................................ 65
3.4.3 Derivation of the Volume-Averaged RANS Equations ....................... 67
3.4.4 The Double-Averaged Continuity Equation ........................................ 70
3.5 FLUID STRESS IN DOUBLE-AVERAGED TERMS ............................................... 71

PhD Thesis iv Table of Contents


L J Campbell
3.6 SIMPLIFIED EQUATIONS FOR TWO-DIMENSIONAL FLOWS ............................... 71
3.7 CHAPTER CLOSURE ........................................................................................ 73

4.0 EXPERIMENTAL METHODOLOGY........................................................... 74


4.1 INTRODUCTION ............................................................................................... 74
4.2 PRINCIPLES OF PIV MEASUREMENT ............................................................... 75
4.2.1 PIV Correlation Algorithms by Example............................................. 77
4.2.1.1 Applying the Autocorrelation Function (ACF)........................ 79
4.2.1.2 Applying the Cross-Correlation Function (CCF).................... 82
4.2.2 PIV Correlation Peak Location: Sub-Pixel Accuracy .......................... 84
4.3 EXPERIMENTAL APPARATUS........................................................................... 86
4.3.1 Hydraulic Flume................................................................................... 86
4.3.2 PIV Equipment ..................................................................................... 87
4.3.2.1 Digital Camera and PC Components ...................................... 88
4.3.2.2 The Illumination Source........................................................... 89
4.3.2.3 PIV Seeding.............................................................................. 90
4.3.2.4 PIV System Timing................................................................... 90
4.3.2.5 PIV Processing Parameters..................................................... 91
4.3.3 Other Parameters .................................................................................. 92
4.3.4 Averaging Procedure............................................................................ 92
4.4 BED ROUGHNESS ............................................................................................ 94
4.5 HYDRAULIC CONDITIONS ............................................................................... 97
4.6 CHAPTER CLOSURE: TABLE OF EXPERIMENTS ................................................ 99
4.7 OVERVIEW OF RESULTS AND DISCUSSION CHAPTERS................................... 103

5.0 DOUBLE-AVERAGED FLOW


CHARACTERISTICS IN THE ROUGHNESS LAYER............................. 105
5.1 INTRODUCTION ............................................................................................. 105
5.2 WHY IS DOUBLE-AVERAGING NECESSARY?................................................. 106
5.2.1 Spatial Variability in the Roughness Layer........................................ 107
5.3 THE DOUBLE-AVERAGED VELOCITY DISTRIBUTION
OVER TRANSVERSE SQUARE BARS ............................................................... 112
5.3.1 Above the Roughness Layer:
The Potential for a Logarithmic Velocity Profile............................... 112
5.3.1.1 Log-Law Parameters ............................................................. 114
5.3.2 Velocity Distribution in the Roughness Layer ................................... 118
5.3.2.1 Skimming Flow ...................................................................... 119
5.3.2.2 Wake Interference Flow......................................................... 121
5.3.2.3 Isolated Roughness Flow ....................................................... 123
5.4 THE NATURE OF FLUID SHEAR STRESS ......................................................... 127
5.4.1 Fluid Stress for Skimming Flow ........................................................ 128
5.4.2 Fluid Stress for Wake Interference Flow ........................................... 131
5.4.3 Fluid Stress for Isolated Roughness Flow.......................................... 133
5.4.4 Form-Induced Stress Levels............................................................... 136
5.5 CHAPTER CLOSURE ...................................................................................... 138

6.0 LOCAL FLOW


CHARACTERISTICS IN THE ROUGHNESS LAYER............................. 140
6.1 INTRODUCTION ............................................................................................. 140

PhD Thesis v Table of Contents


L J Campbell
6.2 LOCAL FLUID MOTION IN THE ROUGHNESS LAYER ...................................... 140
6.2.1 A Preliminary Research Result .......................................................... 140
6.2.2 The Form-Induced Velocity
Components close to Square Bar Roughness ..................................... 141
6.2.2.1 Periodic Behaviour of the
Form-Induced Velocity Components ..................................... 146
6.3 QUADRANT MAPPING ................................................................................... 151
6.3.1 Conventional Quadrant Analysis ....................................................... 151
6.3.2 Spatial Quadrant Analysis: Quadrant Mapping ................................. 152
6.3.2.1 The Zone of the Rings ............................................................ 153
6.3.2.2 The Zone of Irregular Orbits ................................................. 159
6.3.2.3 Evolution of the Quadrant Orbit............................................ 165
6.3.2.4 The Effect of Bar Spacing on Quadrant Maps....................... 165
6.3.2.5 The Effect of Flow Depth on Quadrant Maps........................ 167
6.3.3 Re-assessing the Extent of the Roughness Layer............................... 169
6.4 CHAPTER CLOSURE ...................................................................................... 170

7.0 THE TRANSITION PHENOMENON .......................................................... 173


7.1 INTRODUCTION ............................................................................................. 173
7.2 WHAT IS THE TRANSITION PHENOMENON? ................................................... 174
7.3 FUNDAMENTAL MECHANISMS
UNDERLYING THE TRANSITION PHENOMENON ............................................. 175
7.3.1 Indications from the Temporally-Averaged Flow Field .................... 176
7.3.2 Turbulent Structure at the d-k-type Roughness Transition ................ 180
7.4 EFFECTS OF THE TRANSITION PHENOMENON
DRAG ADJUSTMENT AND OTHER CONSIDERATIONS ...................................... 184
7.4.1 Drag Adjustment ................................................................................ 184
7.4.1.1 Conventional Drag Reduction Techniques ............................ 184
7.4.1.2 Drag Adjustment at the d-k-type Roughness Transition........ 186
7.4.1.3 Supplementary Experiment 1: The Smooth Bed .................... 191
7.4.2 Lateral Effects at the d-k-type Roughness Transition ........................ 194
7.4.2.1 Supplementary Experiment 2: PIV Evidence
from a Transverse (x, y) Plane............................................... 197
7.5 CHAPTER CLOSURE ...................................................................................... 201

8.0 CONCLUSION ............................................................................................... 203


8.1 THESIS CLOSURE .......................................................................................... 207

REFERENCES.......................................................................................................208

APPENDIX A: PROOF OF INTEGRAL THEOREM A.......................................215


APPENDIX B: APPLICATION OF INTEGRAL THEOREMS A AND B ..........220
APPENDIX C: SUPPLEMENTARY PLOTS FOR CHAPTER 5.0 .....................233
APPENDIX D: SUPPLEMENTARY PLOTS FOR CHAPTER 6.0 .....................242
APPENDIX E: SUPPLEMENTARY PLOTS FOR CHAPTER 7.0......................252

PhD Thesis vi Table of Contents


L J Campbell
LIST OF FIGURES
CHAPTER 2
2.1 Definition sketch for coordinate axes (x, y, z) .................................................... 6
2.2 Results from Nikuradses (1933) sand grain roughened pipe experiments ........ 8
2.3 Sub-division of the open-channel flow field (schematic derived from
Nezu & Nakagawa, 1993)................................................................................. 11
2.4 Environmental open-channel flow with small relative submergence............... 14
2.5 Bed roughness geometry notation. ................................................................... 22
2.6 Basic flow types for flows over strip roughness as proposed by Morris (1955)
.......................................................................................................................... 31
2.7 The square bar roughness spacing classification of Knight & MacDonald
(1979a) .............................................................................................................. 33
2.8 The results of Perry et al. (1969) for k-type square bar roughness................... 34
2.9 The results of Perry et al. (1969) for d-type square bar roughness .................. 35
2.10 Flow patterns from the early flow visualisation wind tunnel experiments of
Perry et al. (1969) ............................................................................................. 37
2.11 Photographs of vortex flow patterns in a water channel from Okamoto et al.
(1993)................................................................................................................ 37
2.12 Mean streamlines from the LES simulations of Cui et al. (2003) .................... 38
2.13 The variation in friction factor, f, with longitudinal roughness spacing, L ...... 41
2.14 Subdivision of double-averaged open-channel flow into specific layers
(adapted from Nikora et al., 2001) ................................................................... 48
2.15 The roughness-induced variation in the temporally-averaged streamwise
velocity profiles (from Cui et al., 2003) ........................................................... 51
2.16 The double-averaged streamwise velocity profile from the k-type square bar
flow data shown in Figure 2.15 (from Cui et al., 2003) ................................... 52
2.17 The results for momentum flux of Bohm et al. (2000) from flow over a model
vegetation canopy (the Black Forest model).................................................. 54

CHAPTER 3
3.1 A simplified fluctuating velocity signal showing temporal mean and
fluctuating parts ................................................................................................ 60
3.2 A simplified fluctuating mean velocity signal showing double-averaged
velocity and the spatial disturbance .................................................................. 63
3.3 Schematic pressure field about a series of impermeable fences lying across the
flow (adapted from Raupach and Shaw, 1981)................................................. 66

CHAPTER 4
4.1 Autocorrelation PIV frame depicting fluid ticker-tape trace ......................... 76
4.2 A small area from a pair of images suitable for cross-correlation PIV analysis
.......................................................................................................................... 76
4.3 Example 8 8 pixel AUTOCORRELATION interrogation region showing two
seeding particles passing through ..................................................................... 78
4.4 Example 8 8 pixel CROSS-CORRELATION interrogation region showing
three seeding particles passing through ............................................................ 79
4.5 Autocorrelation function (ACF) plot, clearly illustrating the correlation peaks
for the given example ....................................................................................... 80
4.6 Sketch illustrating the problem of directional ambiguity associated with
autocorrelation PIV........................................................................................... 81

PhD Thesis vii List of Figures


L J Campbell
4.7 The cross-correlation function (CCF) plot........................................................ 83
4.8 Laboratory set-up for the Aberdeen cross-correlation PIV system .................. 88
4.9 The green (532nm wavelength) Nd-Yag laser light sheet illuminating flow in
the longitudinal midline of the flume ............................................................... 90
4.10 Cross-correlation PIV timing diagram.............................................................. 91
4.11 Schematic representation of the double (temporal followed by spatial)
averaging procedure.......................................................................................... 93
4.12 Fixed (6.35mm high) transverse square bar roughness elements within the
flume (L=8 spacing).......................................................................................... 95
4.13 Bed roughness geometry notation (repeated from Figure 2.5) ......................... 96

CHAPTER 5
5.1 Contours of time-averaged streamwise velocity............................................. 106
5.2 Temporally averaged streamwise velocity profiles above roughness spacing
L=5 (wake interference flow) ......................................................................... 108
5.3 The variation in the four velocity profiles shown in Figure 5.2 (roughness
spacing L=5) ................................................................................................... 108
5.4 Streamlines (time-averaged) overlying a contour map of the time-averaged
streamwise velocity for skimming flow over L=3 bar roughness ................ 109
5.5 Streamlines (time-averaged) overlying a contour map of the time-averaged
streamwise velocity for isolated roughness flow over L=12 bar roughness 109
5.6 Temporally averaged Reynolds stress profiles above roughness spacing L=5
........................................................................................................................ 111
5.7 The variation in the four stress profiles shown in Figure 5.6 (roughness spacing
L=5)................................................................................................................. 111
5.8 Comparison of applicability of the log-law for velocity. Experiment
S400_H50_L5 is chosen as the closest match to one of the flow
depth/roughness spacing configurations tested by Grass et al. (1991)........... 113
5.9 The effect of bar spacing, L, on displacement height, d (measured from the
channel bed) for all 3 flow depths .................................................................. 115
5.10 The effect of bar spacing, L, on roughness length, zo, evaluated from a
logarithmic line fit to the measured double-averaged streamwise velocity
profile.............................................................................................................. 116
5.11 The variation in roughness length, zo (normalised with roughness height, k)
with roughness density. Figure adapted from Raupach et al. (1991) ............ 117
5.12 Double-averaged velocity profiles over a range of surface roughness (from
Nikora et al., 2004) ......................................................................................... 119
5.13 Double-averaged streamwise velocity distribution, u ( z ) (normalised with
bulk mean velocity, U=Q/A) for skimming flow cases .................................. 120
5.14 Time-averaged streamlines for roughness configuration L=2, flow depth
H=37mm ......................................................................................................... 121
5.15 Double-averaged streamwise velocity distribution, u ( z ) (normalised with
bulk mean velocity, U=Q/A) for wake interference flow cases ...................... 122
5.16 Double-averaged streamwise velocity distribution, u ( z ) (normalised with
bulk mean velocity, U=Q/A) for isolated roughness flow cases..................... 123
5.17 Sketch illustrating how profile type 1, the S-shaped profile associated with
flow in the cavity vortex, and flow type 2 associated with areas where flow
has reattached to the main bed, combine to yield a linear double-averaged
velocity profile in the interfacial layer between successive square bars ........ 124
5.18 Fluid stress profiles for skimming flow case S400_H50_L3.......................... 129

PhD Thesis viii List of Figures


L J Campbell
5.19 The flow patterns underlying form-induced stress for skimming flow case
S400_H50_L3. ................................................................................................ 130
5.20 Fluid stress profiles for wake interference flow case S400_H50_L5 ............. 132
5.21 The flow patterns underlying form-induced stress for wake interference flow
case S400_H50_L5 ......................................................................................... 132
5.22 Fluid stress profiles for isolated roughness flow case S400_H50_L15 .......... 133
5.23 The flow patterns underlying form-induced stress for isolated roughness flow
case S400_H50_L15 ....................................................................................... 134
5.24 Normalised z-level for the dominant peak in the form induced stress profile,
uw
% % ( z ) , versus bar spacing, L for all three flow depths........................... 136
5.25 Normalised z-level at which form induced stress, uw % % ( z ) , becomes
negligible, versus bar spacing, L, for all three flow depths ............................ 136

CHAPTER 6
6.1 Local form-induced momentum flux, u% w% / u*2 , in planes 1.6-2.4 k above the
channel bed (where k=6.35mm) ..................................................................... 141
6.2 Streamwise variation (largely in the interfacial sub-layer) of the streamwise
form-induced velocity component, u% ............................................................. 143
6.3 Streamwise variation (largely in the interfacial sub-layer) of the bed-normal
form-induced velocity component, w% ............................................................ 143
6.4 Streamwise variation (in the form-induced sub-layer) of the streamwise form-
induced velocity component, u% ...................................................................... 144
6.5 Streamwise variation (in the form-induced sub-layer) of the bed-normal form-
induced velocity component, w% ..................................................................... 144
6.6 Time-averaged streamlines for bar spacings (a) L=7 and (b) L=15 (H=50mm)
showing the diminished extent of the corner vortex (in L=15) in the angle
between the solid bed and the upstream face of the bar roughness ................ 145
6.7 Streamwise variation of u% and w% (m/s) for case S400_H50_L8, showing that
each component undergoes one full cycle over one whole roughness pitch, p.
Profiles for u% and w% are phase shifted by ................................................. 147
6.8 Cross-correlation plots for 4 planes above L=8 rough bed (H=50mm).......... 148
6.9 Form-induced velocities and momentum flux: comparison of suggested
simulation as given in Eq. (6.1) with real data from case S400_H50_L8....... 150
6.10 Quadrant diagrams. (a) Quadrant numbers and associated turbulent events,
(b) Quadrant diagram from experiment S400_H50_L5 generated from velocity
fluctuations a point located above the inter-bar gap at level z=2k .................. 151
6.11 Interpretation of each Quadrant Mapping sector in terms of local (time-
averaged) fluid motion relative to average motion......................................... 153
6.12 Form-induced velocity behaviour in the zone of the rings for experiment
S400_H50_L8 ................................................................................................. 154
6.13 Contours of time-averaged velocity showing the recirculation zones upstream
and downstream of the square bar in experiment S400_H50_L8 ................... 155
6.14 Quadrant diagram showing rings from throughout the form-induced sub-layer
(for experiment S400_H50_L8) ...................................................................... 157
6.15 The meaning of the position vector, r , for data points plotted on quadrant
diagrams.......................................................................................................... 158
6.16 Behaviour of the form-induced velocity components in three planes located
below the roughness crests for experiment S400_H50_L8............................. 160

PhD Thesis ix List of Figures


L J Campbell
6.17 Behaviour of the form-induced velocity components in four planes located
above the roughness crests for experiment S400_H50_L8............................. 161
6.18 Contributions from each of the four spatial quadrant analysis quadrants (Q1 to
Q4) to the total form-induced stress ............................................................... 163
6.19 Evolution of the quadrant rings (experiment S400_H50_L8)......................... 164
6.20 Quadrant maps for wake interference cases ................................................... 166
6.21 Quadrant rings and maps for bar spacing L=10.............................................. 168
6.22 Roughness layer thickness versus square bar spacing. The extent of the
roughness layer is evaluated from quadrant maps of form-induced velocities
........................................................................................................................ 170

CHAPTER 7
7.1 Time-averaged streamlines for (a) d-type roughness, L=5, (b) the d-k-type
transition, L=6, and (c) k-type roughness, L=7 ............................................... 176
7.2 Contours of time-averaged (a) streamwise, and (b) bed-normal velocities (m/s)
in the near-bed region for L=5 d-type spacing................................................ 178
7.3 Contours of time-averaged (a) streamwise, and (b) bed-normal velocities (m/s)
in the near-bed region for L=6 d-k-type transitional spacing.......................... 178
7.4 Velocity vector plot of the temporally averaged (form-induced) flow field at
the d-k-type roughness transition (L=6) for experiment S400_H85_L6 ......... 179
7.5 Temporal quadrant analysis diagrams for (a) L=5 (d-type), (b) L=6 (d-k-type
transition), and (c) L=7 (k-type) roughness configurations ............................ 181
7.6 Contour maps of point density for the quadrant diagrams in Figures 7.5(a) and
(b).................................................................................................................... 182
7.7 Contour maps showing the spatial variation of slope coefficient, a, of temporal
quadrant plots, i.e. w = au............................................................................. 182
7.8 Variation in weir-measured channel discharge, Q (l/s), with roughness spacing,
L, for all three flow depths.............................................................................. 188
7.9 Variation in PIV-measured channel discharge, Q (l/s), with roughness spacing,
L, for all three flow depths.............................................................................. 188
7.10 Weir-measured discharge values for a smooth bed. Dashed line illustrates
discharge calculated using the Manning equation with n=0.012.................... 192
7.11 Weir-measured rough-bed discharge values normalised with the smooth bed
discharge, Qo, for all bar spacings and flow depths....................................... 193
7.12 Double-averaged Reynolds stress profiles for roughness spacings L=5, L=6,
L=7, L=8 (H=50mm) ...................................................................................... 194
7.13 Double-averaged bed-normal (vertical) velocity distribution, w ( z ) for bar
spacings L=2 to L=15...................................................................................... 196
7.14 Results from a bed-parallel plane located 9mm above the bar crests. (a)
contours of time-averaged streamwise velocity, u ( x, y ) , (b) the double-
averaged streamwise velocity profile, u ( y ) , and (c) digital image of the
underlying bed topography ............................................................................. 197
7.15 The transverse velocity component. (a) contour map of time-averaged
transverse velocity, v ( x, y ) (colour bar marked in units of mm/s), and (b) the
double-averaged transverse velocity distribution, v ( y ) .............................. 198
7.16 (a) Secondary flow patterns (looking along the main channel axis) for different
rectangular duct aspect ratios as measured by Knight and Patel (1985). (b) The
pattern of secondary circulation cells suggested for the current dataset above
transitional bar spacing L=6............................................................................ 199

PhD Thesis x List of Figures


L J Campbell
APPENDIX A
A.1 Definition sketch of averaging volume (or area) showing different bounding
surfaces between solid and fluid portions....................................................... 215
A.2 Definition sketch of averaging volume (or area) before and after small shift in
position (xi)................................................................................................... 217
A.3 Geometry of a small portion of the surface, as shown in A.2......................... 218

APPENDIX C
C.1 Contours of time-averaged streamwise velocity, H=37mm ........................... 233
C.2 Contours of time-averaged streamwise velocity, H=85mm ........................... 234
C.3 The double-averaged velocity distribution in the roughness layer for H=37mm
........................................................................................................................ 235
C.4 The double-averaged velocity distribution in the roughness layer for H=50mm
........................................................................................................................ 236
C.5 The double-averaged velocity distribution in the roughness layer for H=85mm
........................................................................................................................ 237
C.6 Form-induced stress (N/m2) distributions for H=37mm................................. 238
C.7 Form-induced stress (N/m2) distributions for H=50mm................................. 239
C.8 Form-induced stress (N/m2) distributions for H=85mm................................. 240
C.9 Sequence of ten representative images of instantaneous streamlines from d-
type experiment S400_H50_L2 ...................................................................... 241

APPENDIX D
D.1 S400_H50_L2: time-averaged streamlines ..................................................... 242
D.2 S400_H50_L3: time-averaged streamlines ..................................................... 242
D.3 S400_H50_L4: time-averaged streamlines ..................................................... 243
D.4 S400_H50_L5: time-averaged streamlines ..................................................... 243
D.5 S400_H50_L6: time-averaged streamlines ..................................................... 243
D.6 S400_H50_L7: time-averaged streamlines ..................................................... 244
D.7 S400_H50_L8: time-averaged streamlines ..................................................... 244
D.8 S400_H50_L10: time-averaged streamlines ................................................... 244
D.9 S400_H50_L12: time-averaged streamlines ................................................... 245
D.10 S400_H50_L15: time-averaged streamlines ................................................... 245
D.11 S400_H50_L20: time-averaged streamlines ................................................... 245
D.12 Simulated versus measured form-induced velocity components: S400_H50_L7
........................................................................................................................ 246
D.13 Simulated versus measured form-induced velocity components: S400_H50_L8
........................................................................................................................ 246
D.14 Simulated versus measured form-induced velocity components:
S400_H50_L10 ............................................................................................... 247
D.15 Simulated versus measured form-induced velocity components:
S400_H50_L12 ............................................................................................... 247
D.16 Simulated versus measured form-induced velocity components:
S400_H50_L15 ............................................................................................... 248
D.17 Simulated versus measured form-induced velocity components:
S400_H50_L20 ............................................................................................... 248
D.18 Form-induced quadrant analysis: S400_H50_L7............................................ 249
D.19 Form-induced quadrant analysis: S400_H50_L8............................................ 249
D.20 Form-induced quadrant analysis: S400_H50_L10.......................................... 250
D.21 Form-induced quadrant analysis: S400_H50_L12.......................................... 250

PhD Thesis xi List of Figures


L J Campbell
D.22 Form-induced quadrant analysis: S400_H50_L15.......................................... 251
D.23 Form-induced quadrant analysis: S400_H50_L20.......................................... 251

APPENDIX E
E.1 Ten representative time-steps showing instantaneous streamlines from flow
over d-type roughness spacing L=5 ................................................................ 253
E.2 Ten representative time-steps showing instantaneous streamlines from flow
over the d-k-type roughness transition L=6 .................................................... 254
E.3 Contour map showing the spatial variation of slope coefficient, a, of temporal
quadrant plots, i.e. w = au : L=5 ................................................................... 255
E.4 Contour map showing the spatial variation of slope coefficient, a, of temporal
quadrant plots, i.e. w = au : L=6 ................................................................... 255
E.5 Contour map showing the spatial variation of slope coefficient, a, of temporal
quadrant plots, i.e. w = au : L=7 ................................................................... 255

PhD Thesis xii List of Figures


L J Campbell
LIST OF TABLES
CHAPTER 2
2.1 Part 1: Closed-channel flows ............................................................................ 17
2.1 Part 2: Open-channel flows............................................................................... 23
2.1 Part 3: Numerical simulations........................................................................... 25

CHAPTER 4
4.1 Roughness geometry parameters L, k and p (refer to Figure 4.13), together with
mean bed level and variation in porosity, (z).................................................. 98
4.2 (a) Flow depth H=37mm: Table of experiments showing roughness, flow, PIV
and averaging parameters ............................................................................... 100
4.2 (b) Flow depth H=50mm: Table of experiments showing roughness, flow, PIV
and averaging parameters ............................................................................... 101
4.2 (c) Flow depth H=85mm: Table of experiments showing roughness, flow, PIV
and averaging parameters ............................................................................... 102

APPENDIX E
E.1 Weir-measured discharge (Q) for flow depth H = 37 mm.............................. 256
E.2 Weir-measured discharge (Q) for flow depth H = 50 mm.............................. 256
E.3 Weir-measured discharge (Q) for flow depth H = 85 mm.............................. 256

PhD Thesis xiii List of Tables


L J Campbell
NOMENCLATURE

Symbol Description Units


A Flow cross-sectional area m2
A(z) Roughness geometry function (A= Af /Ao ) -
Af Area occupied by fluid in the fixed averaging region Ao m2
Ao Area of the fixed averaging region m2
a Slope coefficient of temporal quadrant plots, w = au -
au , aw Amplitude of oscillation of form-induced velocity
-
components u% and w%
B Channel width m
CD Drag coefficient -
c Clauser (1954) boundary layer thickness m
D Characteristic flow length for Reynolds number,
m
e.g. pipe diameter/simulation domain height/duct height
D50 Grain size, subscript denoting percentage finer than m
d Displacement height for logarithmic velocity law m
Fr Froude number -
Fi In-plane loss of pairs from PIV interrogation region -
Fo Out-of-plane loss of pairs from PIV interrogation
-
region
f Friction factor (f=2 u*2 /U2) -
G 8-bit pixel greyscale value -
g Gravity acceleration ms-2
H Mean flow depth m
itot Number of pixels in PIV interrogation region -
k Roughness height m
ks Nikuradses equivalent sand grain roughness m
L Square bar normalised roughness spacing (L=P/k) -
La Accurate roughness spacing (Table 4.1) -
Li Set of lengths characterising a rough boundary m
lx,lz Roughness element dimensions in x and z directions m
l Characteristic length scale, inner layer m
l Mixing length m
Ni Particle image pair density within PIV interrogation
-
region
n Mannings n sm-1/3
ni ith component of the unit normal vector -
P Square bar centre-to-centre roughness pitch m
p Fluid pressure Nm-2
Q Channel discharge m3s-1
Qo Smooth bed channel discharge m3s-1
R Channel hydraulic radius m
Re Reynolds number -
Re* Roughness Reynolds number ( Re* = ku* / ) -
Rep Pipe flow Reynolds number -
Re Reynolds number based on momentum thickness -
r Pipe radius m
r Cross-correlation -

PhD Thesis xiv Nomenclature


L J Campbell
Symbol Description Units
r Position vector for points on quadrant diagrams -
S Bed slope (for experiment identifiers) -
Sint Interface between solids and fluid in the averaging
-
region
s PIV image space interval m
t Temporal averaging period, PIV image time interval s
U Depth-averaged velocity ms-1
U Bulk mean velocity ( U =Q/A) ms-1
Uo Maximum streamwise velocity ms-1
u, v, w Instantaneous longitudinal, transverse, and vertical
ms-1
velocity components, respectively
u,v, w Time-averaged velocity components, i.e. u i = u i u i ms-1
u , v , w Flow velocity fluctuation components about temporal
ms-1
mean, i.e. u i = u i u i
u~, v~, w ~ Form-induced velocity components, i.e. u~ = u u
i i i ms-1
u , v , w Double-averaged (in time and spatial domains) velocity
ms-1
components, i.e. u i = u i + u~i
u/u Roughness function -
u* Friction (shear) velocity ms-1
uiu j Local Reynolds (turbulent) stress Nm-2
uiu j Spatially-averaged Reynolds (turbulent) stress Nm-2
u%i u% j Form-induced (or dispersive) stress Nm-2
Vf Volume occupied by fluid in the fixed averaging region
m3
Vo
Vo Volume of the fixed averaging region m3
vi ith component of the velocity vector of the solid-fluid
ms-1
interface
w Inter-bar gap (Cui et al., 2003) m
x, y, z Coordinate axes (longitudinal, transverse, and bed-
m
normal)
x, z PIV analysis spatial shifts m, pixels
Z Corrected bed-normal coordinate for log law (Z=z-d) m
zo Roughness length m
zt Level of roughness troughs m
zc Level of roughness crests m
zf Level of solid floor m
zR Upper bound of roughness layer m
zL Upper bound of logarithmic layer m
zws Elevation of water surface m
Bed slope, phase shift between form-induced velocity
-
components
Boundary layer thickness m
Porosity (=Vf / Vo) -
von Karman constant -
Fluid (dynamic) viscosity Nsm-2

PhD Thesis xv Nomenclature


L J Campbell
Symbol Description Units
Fluid (kinematic) viscosity m2s-1
time- (or ensemble-) averaged flow variable (e.g.
velocity or pressure) defined in the fluid only
Intrinsic spatial average of
Superficial spatial average of (such that
t
t = )
Density of water Kgm-3
x Strip roughness geometry, x = lx / k (Raupach et al.,
-
1991)
Total fluid shear stress Nm-2
o Bed shear stress Nm-2
ij Fluid stress tensor Nm-2

PhD Thesis xvi Nomenclature


L J Campbell
CHAPTER MARKERS

Chapter 1: INTRODUCTION

Laser texture scan of a portion of cobble bed from the


Waimakariri River, South Island, New Zealand. (Courtesy
of G. Smart, the National Institute of Water and
Atmospheric Research Ltd. (NIWA) Christchurch, NZ).

Chapter 2: LITERATURE REVIEW

Sand ripples in Death Valley, Nevada. (Source: adapted


from www.econetwork.net/~jdavis/DeathValley)

Chapters 3 & 4: METHODOLOGIES

Plan view of PVC cube roughness ready for testing in a


hydraulic flume in the Department of Hydraulic
Engineering at the Technical University of Braunschweig.
(Courtesy of K. Koll).

Chapters 5, 6, & 7: RESULTS & DISCUSSION

Drought: the effect of a prolonged period of dry weather


on the bed of a lake in South Island, New Zealand.
(Reproduced with kind permission from the NIWA photo
library, Christchurch, NZ).

Chapter 8: CONCLUSIONS

Mixed size granular sediment. (Source: adapted from


www.co.vermilion.il.us/Wilderness/gravel.jpg)

PhD Thesis xvii Chapter Marker Information


L J Campbell
1.0 INTRODUCTION

The nature of turbulent fluid flow in the vicinity of a rough wall is complex. Indeed,
in the near-wall region,

. . . we must expect that the mean velocity u ( z ) will depend


considerably on the form and mutual spacing of the irregularities of
the wall and will be different above the protrusions and over the
hollows between them. Thus here it is impossible to hope for any
simple general rules.
Monin & Yaglom (1977)

More than a quarter of a century on, this challenging statement remains largely
unaddressed. Although much research effort has concentrated on furthering our
understanding of the mean velocity distribution in open-channel flow layers beyond
the influence of surface roughness (i.e. in the logarithmic and outer flow layers),
relatively little attention has been granted to elucidating the flow characteristics
closer to a rough wall. More specifically, the near-bed flow structure in deep flows
over rough beds as well as the flow structure of shallow flows with small relative
submergence (e.g. gravel-bed rivers, where the roughness elements may protrude
above the flow surface) is still unclear in many respects.

Timely advances in experimental and analytical methodologies over the past 25


years have however brought the possibility of finding simple rules for velocity
distribution in the roughness layer ever closer to reality. For example, this study
makes use of a Particle Image Velocimetry (PIV) system. PIV provides a
fundamentally different perspective on the flow compared to more traditional point
measurement techniques (e.g. Laser Doppler Velocimetry) as it measures how flow

PhD Thesis 1 Chapter 1.0


L J Campbell Introduction
velocities vary in both time and space. This experimental methodology (PIV) is used
in conjunction with double-averaging of the momentum equations (the analytical
methodology) such that the recorded velocity vector fields are averaged in both
temporal and spatial domains. These experimental and analytical methodologies are
entirely complementary, as the double-averaged form of the Navier-Stokes equations
provides a theoretically sound framework for the analysis of PIV data.

Double-averaging of the momentum equations is necessary because the time- (or


ensemble-) averaged flow structure is highly three-dimensional and spatially
heterogeneous near a rough bed, making the time-averaged momentum equations
(the Reynolds Averaged Navier-Stokes Equations) inconvenient and impracticable.
In some studies, mean velocities and turbulence characteristics have been implicitly
considered as somewhat averaged in the spatial domain, though the averaging
operation was not defined explicitly and all considerations were based on the time-
averaged momentum equations. Such an intuitive approach introduces uncertainties
in the interpretation of both modelling results and turbulence measurements taken
near rough surfaces. For example, even at the most basic level it is contradictory to
infer boundary shear stress (clearly a spatially-averaged parameter) from local flow
properties as provided by the Reynolds equations. Put simply, the answer would
depend upon the measurement location relative to the bed topography, as underlined
in the opening quote from Monin and Yaglom.

To avoid this problem, time (or ensemble) averaging of the Navier-Stokes equations
can be supplemented by spatial (area or volume) averaging in the plane parallel to
the mean bed level. As a result, new continuity and momentum conservation
equations may be obtained which are averaged in both temporal and spatial domains.
These relate to the time-averaged Reynolds equations as the Reynolds equations
relate to the Navier-Stokes equations for instantaneous flow variables. A significant
advantage of these new equations is that they explicitly include drag terms and
dispersive momentum fluxes due to the spatial heterogeneity of the time-averaged
flow. Importantly, these new terms appear naturally as a result of the double
averaging procedure, and not as additional ad hoc terms.

PhD Thesis 2 Chapter 1.0


L J Campbell Introduction
1.1 Aims and Objectives

This study sought to exploit the inherent strength of PIV in analysing non-uniformity
in the flow, and to harness this with a complementary theoretical development,
namely the flow equations in their double-averaged form. In doing so, a key aim of
the project was to address, qualify and challenge the opening statement from Monin
and Yaglom for a range of rough wall open-channel flows. Sound knowledge of the
non-uniformity of near-bed flow is not only of interest from the standpoint of
fundamental fluid mechanics, but is important for assessing flow resistance and in
quantifying the influence of hydraulics on habitat for bottom dwelling flora and
fauna. Furthermore, it is relevant to any fluid flow over a rough surface; such a flow
type is pertinent to many practical problems in diverse fields.

The general impetus behind this study was to deepen our understanding of the effect
of bed roughness on the velocity distribution in open-channel flows, with particular
emphasis on the near-bed region. This involved investigating the nature of temporal
and spatial velocity fluctuations and their relation to fixed, rough bed topographies.
In terms of channel roughness, the selected starting point was necessarily simple and
well defined. Specifically, this thesis presents findings from an extensive series of
flows over two-dimensional (transverse) square bar roughness.

The objectives of the project were,

1) to conduct a major experimental program to further our understanding


of the fundamental mechanics underlying the interaction between fluid
flow and a rough surface; this involves detailed measurements of
temporally, and spatially varying fluid velocities;

2) to evaluate the general applicability of the double-averaged equations


for describing flow over a range of relative roughnesses and roughness
configurations;

3) to test specific relationships derived by double-averaging for the near-


bed velocity distribution over a range of relative roughnesses and
roughness types;

PhD Thesis 3 Chapter 1.0


L J Campbell Introduction
4) to quantify the extent of non-uniformity in the near-bed flow for a range
of relative roughnesses and roughness configurations;

5) to establish an extensive dataset of detailed velocity measurements for


steady, uniform open-channel flows over geometrically-simple bed
roughnesses, which may be readily transferred to other researchers
concerned with the development and benchmarking of numerical
models.

1.2 Thesis Outline

Following this introduction, Chapter 2 presents a selection of the relevant technical


literature. The reviewed literature falls into three broad categories, beginning with a
brief discussion of open-channel flow characteristics. The bulk of the chapter is then
dedicated to summarising both the existing square bar and double-averaging
research.

Chapters 3 and 4 give a comprehensive account of the analytical and experimental


methodologies respectively. The former concentrates on explaining the derivation
and key features of the double-averaged momentum and mass conservation
equations. The latter gives details of the measurement technique (PIV) and
experimental procedure, together with roughness and hydraulic parameters.

Chapters 5, 6, and 7 are all combined results and discussion chapters, each one
pertaining to a particular subsection of the overall findings. These are, in order, the
double-averaged flow characteristics in the near-bed region (Chapter 5), the local
flow characteristics in the near-bed region (Chapter 6), and the newly discovered
transition phenomenon (Chapter 7).

In conclusion to this thesis, Chapter 8 summarises the key outcomes from the results
of the experimental programme.

PhD Thesis 4 Chapter 1.0


L J Campbell Introduction
2.0 LITERATURE REVIEW

2.1 Introduction

The outcomes from this project are potentially significant right across the board of
fluid mechanics research. Indeed, elucidation of the precise nature of the interaction
between moving fluid and a rough surface is broadly relevant to a diverse range of
flow types in many engineering sectors. For example open-channel flows in rivers
and canals, fluids flowing in artificially roughened pipes and ducts, coolant
behaviour in heat exchange matrices, vehicle motion through air or water, and the
interaction of the atmospheric boundary layer with urban and vegetative canopies are
all of interest. Consequently, there is a vast amount of literature concerning the
hydraulics of rough wall flows written from the viewpoint of hydraulic engineers,
mechanical engineers, atmospheric physicists and other scientists. However, beyond
a brief prcis of boundary layer theory, the scope of this literature review is strictly
and necessarily limited to presenting existing studies of flows over square bars,
before discussing the development and application of double-averaging
methodology. Hence, the contents of this chapter are organised as follows,

Section 2.2: An overview of the development of our current understanding


of fluid flows over rough boundaries, including open-channel flows:
Boundary Layer Theory.

Section 2.3: A review of flow patterns and resistance over transverse strip
roughness, or Fluid Flow over Transverse Square Bars.

Section 2.4: The development of double-averaging theory, with particular


emphasis on the application of the methodology to open-channel flows,
Double-Averaging Methodology.

PhD Thesis 5 Chapter 2.0


L J Campbell Literature Review
For clarity at the outset, Figure 2.1 defines the coordinate directions and their
associated velocity vector components that are used throughout this thesis.

w
v
u
S

Figure 2.1: Definition sketch for coordinate axes (x, y, z). Streamwise (x) component of velocity
is u; Transverse (y) component is v, directed towards the left channel boundary when looking
downstream; Bed-normal (z) component is denoted by w, and S is bed slope.

2.2 Boundary Layer Theory

The purpose of this section is to provide a general introduction to the field of rough
wall hydraulics, particularly the turbulent flow regimes that are generated at different
Reynolds numbers.

Fluid flow over any surface, whether it is a ships hull or a gravel-bed river, is
characterised by the development of a boundary-layer. The theory of boundary-
layers owes its origins to the work Ludwig Prandtl reported in 1904 his research
demonstrated how it was possible to mathematically analyse flows around solid
bodies involving low viscosity fluids. Prandtls fundamental hypothesis stated that
the effects of friction need only be considered in the narrow boundary layer in the
immediate vicinity of a solid body. It was suggested that elsewhere, in the core of the
flow, the classical hydrodynamics theory describing ideal (frictionless) fluid flow
applied.

The flow of a real fluid can be distinguished from that of an ideal fluid by two
fundamental features. Firstly, in real flows there is no discontinuity of velocity, and

PhD Thesis 6 Chapter 2.0


L J Campbell Literature Review
secondly the condition of no-slip must apply at solid surfaces. Hence, when fluid
flows past a solid boundary there is a thin region over which the fluid must
accelerate from zero to a velocity asymptotically approaching that of the main
stream. In this zone where there is a high rate of increase of fluid velocity with
distance from the boundary, shearing stresses play an important role. The boundary
layer can therefore be thought of as the narrow band of fluid flow over a solid
surface where shear stress (friction) is of prime importance. Typically, the boundary
layer is assigned an upper limit when the mean flow velocity has reached 99% of the
mean mainstream velocity. Beyond this layer, where velocity gradients are small,
low viscosity fluids can be considered to behave as if ideal.

2.2.1 Hydraulically Smooth and Hydraulically Rough Flow

Although a boundary layer over a rough wall may be either laminar or turbulent in
nature, once turbulent flow dominates, it can be further classified as hydraulically
smooth or hydraulically rough. In the region of laminar flow, all rough walls offer
the same resistance as the equivalent smooth wall. Furthermore, in the turbulent flow
regime, there is a range of Reynolds numbers (Re = UD / , U = bulk mean velocity,
D = flow characteristic length, e.g. pipe diameter for pipe-flow, = fluid kinematic
viscosity) over which flows with a given relative roughness (k/D, k = roughness
height) will behave as if smooth-walled. The rough bed can therefore be said to be
hydraulically smooth, and flow resistance is determined by Reynolds number alone.
However, following a transition region, at some higher value of Reynolds number
the resistance curve for a rough wall deviates from the smooth wall line, resulting in
a quadratic resistance law. The flow is then termed hydraulically rough. Empirical
evidence for both states dates back to the work of Nikuradse, first published in 1933.

Nikuradse conducted a set of experiments observing flow in artificially sand-grain-


roughened round pipes; the findings are presented in Figure 2.2. The straight-line
laminar section of the graph shows that below a pipe-flow Reynolds number (Rep) of
around 2000, the flow is independent of pipe roughness. In this region the friction
factor is a function of pipe Reynolds number alone, and follows the linear relation f
= 16/Rep. In essence, any turbulence caused by protrusions on the pipe wall is

PhD Thesis 7 Chapter 2.0


L J Campbell Literature Review
damped out solely by viscosity. In contrast however, in the transition zone beyond
this critical pipe-flow Reynolds number, the flow does begin to depend on pipe
roughness, with the observed range of curves delineating the effect of different
relative roughnesses, k/D (sand grain roughness height normalised with pipe
diameter).

k/D

Rep

Figure 2.2: Results from Nikuradses (1933) sand grain roughened pipe experiments. f =
friction factor, Rep=pipe-flow Reynolds Number ( = (UD ) / ), k = sand grain size, D = pipe
diameter, k/D = relative roughness Adapted from Massey (1990).

Examination of Figure 2.2 reveals that for any given value of relative roughness,
moving fluid will sense the pipe as being smooth-walled up to the value of Rep that
causes departure from the smooth pipe line. With increasing Rep, for a given k/D,
flow then enters a transitional zone whereby a further boost in Rep results in a fully
rough flow regime, where the friction factor becomes independent of Rep. For
rougher walls, the transition to the fully rough zone of flow will occur at lower
values of Rep. The concept of a viscous sub-layer may be used to explain this flow
behaviour.

Even in fully turbulent flow, the random turbulent motions perpendicular to a pipe
wall must die out as the wall is approached. This dictates that there exists a thin layer

PhD Thesis 8 Chapter 2.0


L J Campbell Literature Review
in the immediate vicinity of the boundary in which the random motions are
negligible, namely the viscous sub-layer. However, as Rep increases so does
turbulence intensity, the result being that the turbulent eddies are able to penetrate
the viscous sub-layer more deeply. The viscous sub-layer therefore thins with
increasing Reynolds number.

Recognising that these results are generally applicable not only to pipe flow, but to
flow over any surface, and returning to the title of this section: hydraulically smooth
flow occurs when the viscous sub-layer completely submerges the roughness
elements that comprise the boundary. Conversely, hydraulically rough flow is
encountered when the viscous sub-layer is unable to shield the main flow from
surface irregularities.

In the smooth zone the viscous sub-layer blankets any surface irregularity so that the
size of roughness elements has no effect on the main flow, hence, all the curves for
the smooth zone coincide (Figure 2.2). However, with increasing Rep, surface
roughness begins to protrude through the thinning sub-layer and turbulent flow
around individual elements causes wakes to be shed into the flow. In turn these
wakes of small eddies give rise to a resistance force known as form drag, resulting
from differences of pressure over the surface. The kinetic energy (extracted from the
mean flow) associated with the continual production of these eddies is proportional
to the square of their velocities, which are themselves proportional to the general
velocity. Form drag is thus proportional to the square of the mean velocity of flow,
and in the fully turbulent region represents the main cause of energy loss, as simple
viscous effects are negligible. In the transition zone form drag and viscous effects are
both present to some extent.

From the physical point of view, it is clear that the balance between the roughness
size and the boundary layer thickness plays a central role in determining the
influence the boundary has on the mean flow. More specifically, the flow behaviour
is expected to depend on the ratio of roughness height to thickness of the laminar
(viscous) sub-layer. Since the viscous sub-layer thickness is of the scale / u*
(Schlichting, 1979), this dependence can be encapsulated in a dimensionless

PhD Thesis 9 Chapter 2.0


L J Campbell Literature Review
roughness factor, the roughness Reynolds number, Re* = ku* / , where k =
roughness height (strictly Nikuradses sand grain roughness size, ks, should be used)
and u* = friction velocity. Different flow regimes can therefore be classified on the
basis of this factor, with accepted bands for each flow type (e.g. Nezu and
Nakagawa, 1993) as follows,

All protrusions are contained


Hydraulically smooth regime 0 Re* 5
within the laminar sub-layer
Protrusions partly outside
laminar sub-layer: additional
Transition regime 5 Re* 70 resistance (compared to
smooth regime) is from form
drag
All protrusions reach outside
the laminar sub-layer: main
Hydraulically rough regime Re* > 70 flow resistance due to form
drag, hence resistance
becomes quadratic

In conclusion to this sub-section, note that the primary focus of this project is fully
developed hydraulically rough flow.

2.2.2 Open-Channel Flow

Geophysical open-channel flows are almost exclusively rough-turbulent. They occur


in nature in rivers, lakes, estuaries and oceans, and in our built environment in man-
made hydraulic structures such as spillways, canals, weirs, irrigation networks, and
other water supply systems. Sound knowledge of the velocity and turbulence
structure of such flows is therefore indispensable for building our understanding of
the interaction of a fluid flow and a rough wall in many practical situations. In turn,
deeper knowledge of the fundamental mechanics governing the near-wall flow
structure may enhance and inform our current treatment of related issues, for
example surface friction and sediment transport.

PhD Thesis 10 Chapter 2.0


L J Campbell Literature Review
Length Velocity
Scale Scale z/H

Free-Surface Effect
Flow Maximum
Free-surface Depth, H Velocity, Uo
Region
~0.6

Intermediate
Region

Wall Effect
~0.15
Inner Length, Friction
Wall Region / u* or k Velocity, u*
0.0

Figure 2.3: Sub-division of the open-channel flow field (schematic derived from Nezu &
Nakagawa, 1993). Illustrates the general zones of applicability of the wall, intermediate and
free-surface layers, together with the appropriate length and velocity scales for each. Vertical
axis is not drawn to scale.

A basic difference between open- and closed-channel flows is that whilst the former
is gravity driven, the latter is governed by pressure. Furthermore, open-channel flows
are typified by the presence of a free surface i.e. an air-water interface open to
atmospheric pressure. Velocity and turbulent structure for any given level in the
water column an open-channel flow is therefore influenced (to varying extents) by
both the enclosing boundary and free surface. On this basis, it is proposed that the
flow field in open channels can be sub-divided into three main zones (e.g. Nezu and
Nakagawa, 1993). Figure 2.3 shows this sub-division for open-channel flows, and
highlights the associated length and velocity scales for each region.

With reference to Figure 2.3 the three open-channel flow layers may be described as
follows,

1) The Wall Region


This region corresponds to the zone of flow closest to the channel boundary, and is
typically considered up to the level z = (0.15-0.2)H (Raupach et al., 1991; Nezu and

PhD Thesis 11 Chapter 2.0


L J Campbell Literature Review
Nakagawa, 1993; Jimenez, 2004). Length and velocity scales are / u* (smooth bed)
or k (rough bed), and u* respectively. In this region turbulent energy production
exceeds dissipation. The wall region is often termed the inner layer and turbulent
structure is controlled by the inner variables associated with the boundary
topography. In this region the law of the wall states simply that u / u* = f ( z / l) ,
where, l is a characteristic length scale for the inner layer. In the case of smooth bed
flows, l is the viscous length scale ( l = / u* ), whereas for rough boundaries l is
usually taken to be the roughness height, k.

Probably the most commonly applied function satisfying the law of the wall beyond
the immediate vicinity of the channel boundary is the logarithmic law. The log-law
may be written as,

u 1 zd
= ln (2.1)
u* zo

where u = (temporal) mean streamwise velocity,


= von Karman constant,
d = displacement height, and
zo = roughness length.

Such a mean streamwise velocity distribution u (z) as given in Equation (2.1) is now
well established, and can be derived in two ways. The first is by a classical
asymptotic matching process (Millikan, 1938, Raupach et al., 1991), whereby the
logarithmic distribution is deduced by considering the transition region between the
near-wall and outer layers. For the velocity laws in both regions (as derived by
dimensional analysis) to be mutually valid in the overlap zone, Millikan (1938)
showed that the overlap-layer velocity must vary logarithmically with z. The second
approach to the log-law arises from consideration of mixing length theory (Prandtl,
1925). Deriving the log-law this way requires the assumptions of constant fluid shear
stress in the vicinity of the wall (a proposition now widely accepted for the bottom

PhD Thesis 12 Chapter 2.0


L J Campbell Literature Review
15-20% of the turbulent velocity field), and that the dimension of turbulent eddies
(the mixing length, l) is proportional to the distance from the wall (i.e. l = z,
where is the von Karman constant, as in Equation (2.1).

Whilst it is of course true that u depends on z, the height above datum, a rough
surface has a tendency to displace the entire flow field upwards, such that the actual
elevation of interest is the displaced height, Z = z d [see Equation (2.1)], where d is
called the zero-plane displacement (0 < d < k, and d = 0 for a smooth surface).
Typical displacement heights, together with other log-law parameters for flows over
square bars, are presented in Chapter 5 (Section 5.3.1.1).

2) The Free-Surface Region


This region is typically allocated the range 0.6 < z / H 1.0 , where the turbulent
structure is controlled by the outer variables. As such, length and velocity scales are
specified as flow depth, H, and maximum mainstream velocity, U o , respectively.

According to the experiments of Coles (1956), velocities in this region may deviate
substantially from those predicted by the log-law of the wall. It was therefore
subsequently suggested (Coles, 1956; Coleman, 1981; Zippe and Graf, 1983) that the
observed discrepancy could be accounted for with the addition of a wake function.
Hence a log-wake law is usually applied in the free-surface region. In contrast to
closed-channel flows, turbulence characteristics (most notably vertical turbulence
intensity) are strongly affected by the presence of a free surface. In this region the
turbulent dissipation rate becomes larger than the production rate.

3) The Intermediate Region


As the name suggests, and Figure 2.3 demonstrates, the intermediate region occupies
(0.15 0.2) z / H 0.6 , and is not strongly influenced by either the wall
properties or the free surface. Nezu and Nakagawa (1993) suggest that length and

velocity scales are z and / , respectively (where is the total shear stress, and
is fluid density. Taken together, the intermediate and free-surface regions
comprise the outer layer. The log-law therefore represents an overlap layer between
the inner and outer flow regions. Viscous effects are negligible in the outer layer.

PhD Thesis 13 Chapter 2.0


L J Campbell Literature Review
As an endnote to this description of open-channel flow layers, recall that the primary
focus of this project is the wall region of rough open-channel flows. The
intermediate and free surface regions are not considered.

Figure 2.4: Environmental open-channel flow with small relative submergence (H/k~1 in
many areas). The log-law distribution for the mean streamwise velocity is not valid for such
flows as the roughness height, k, is of the same order of magnitude as the flow depth, H. The
image shows a typical reach on the River Ehen in the Lake District, which provides an
important habitat for the endangered fresh water pearl mussel. Knowledge of the velocity
distribution of such flows is imperative for developing our understanding of the hydraulic
requirements of such bottom dwelling flora and fauna.

As stated in the opening chapter of this thesis, much research effort has concentrated
on furthering our understanding of the mean velocity distribution in open-channel
flow layers beyond the influence of surface roughness (i.e. the log and log-wake laws
for the inner and outer flow layers respectively), but relatively little attention has
been granted to elucidating the flow characteristics closer to a rough wall. At large
ratios of flow depth, H, to roughness height, k, the flow structure for most of the
depth reveals properties similar to those for flows over smooth boundaries, at least at
distances from the bed sufficiently larger than the roughness height. For such flows
the logarithmic velocity profile has been shown to be valid for the inner flow region,

PhD Thesis 14 Chapter 2.0


L J Campbell Literature Review
away from both the rough bed and the water surface. However, Raupach et al. (1991)
recommend that the overlap region for log-law behaviour lies in the range (2-5)k
z 0.2H. This is important as many environmental open-channel flows are classified
as shallow, having low relative submergence, H/k < 5 (Dittrich and Koll, 1997). For
example, the free-surface of the gravel bed river shown in Figure 2.4 is frequently
broken as rocks protrude through the air-water interface (so H/k < 1 in many areas),
hence it is entirely unrealistic to expect a simple logarithmic velocity distribution
according to the estimate of Raupach et al. (1991). Therefore, although the log-law is
rather well understood, there are many practical situations for which it is simply not
applicable (Dittrich and Koll, 1997). Hence, at present there is a gap in knowledge
concerning both the whole velocity distribution of shallow, rough open-channel
flow, and, more generally, the near-bed characteristics of deeper rough-bed flows.

In this respect, it is worth noting that the primary focus of this project is shallow
open-channel flow.

The above layers (i.e. the wall, intermediate and free-surface regions) in open-
channel flow will be revisited and refined within the context of double-averaging
later on in this chapter (Section 2.4.2.1). However, before this methodology is
reviewed, the following section concentrates on presenting past studies of rough bed
flows over a particular roughness type, transverse square bars.

2.3 Fluid Flow over Transverse Square Bars

Simple strip roughness has been extensively studied. For example, channels or pipes
containing transverse bars of varying height, width and spacing are often used in
friction factor and heat- and mass-transfer investigations (Cui et al., 2000) or to
investigate street canyons in the urban roughness layer. Turbulent flows over regular
rough surfaces are often encountered in practice. Within an engineering context,
such rough surfaces are most commonly used to enhance heat transfer. More
generally, they are also of prime importance for furthering the fundamental science
and understanding of flow over any rough boundary.

PhD Thesis 15 Chapter 2.0


L J Campbell Literature Review
2.3.1 General Overview of Square-Bar Roughness Studies

To provide a general overview of the background literature, Table 2.1 presents the
key features of studies from across the spectrum of available square bar roughness
literature. Presented information includes a brief statement of the impetus behind
each study, and, where experimental, the measurement techniques implemented. In
each case the tested bar spacings, L, are given (L = P/k, where P = streamwise bar
pitch, k = bar height - refer to Figure 2.5 for a definition sketch for this notation),
together with any other relevant parameters such as roughness height, Reynolds
numbers, and relative submergence. The contents of Table 2.1 are sub-divided on the
basis of technique, i.e. whether the method was primarily Experimental: Closed-
Channel (non- free-surface flows, table Part 1), Experimental: Open-Channel (Part
2), or Simulation (Part 3). Then, within each sub-table the studies are listed
chronologically (earliest first). A list of abbreviations and notation used in Table 2.1
appears at the end of Part 3.

Note that an exhaustive review of every technical article containing any mention of
square bar roughness would be impractical within the scope of this thesis, but every
effort has been made to include the most relevant literature (some early reports and
non-English language articles are inevitably omitted).

PhD Thesis 16 Chapter 2.0


L J Campbell Literature Review
Table 2.1, Part 1: CLOSED-CHANNEL FLOWS
(Pipes, Ducts and Wind Tunnels)

Data Source Purpose of Method Bar Spacing Reynolds Other Notes


Study (pitch to Number, relevant
height ratio), Re parameters
L
1 Morris (1955) Identifies basic REVIEW Reviews rough pipe-flow data
flow types over Pipe-flow review article 2.18, 3.95, from many sources. Includes
regular rough using available velocity 4.33, 8.69 sine wave corrugations,
surfaces and distribution measurements. square bars and staggered
corresponding (2D square/ --- --- square-base pyramids.
friction factors. rectangular bar Proposed skimming, wake-
cases selected interference and isolated
only) roughness flow
terminologies.
2 Perry & Joubert Turbulent EXPERIMENTAL wind One velocity profile gathered
(1963) boundary layers in tunnel. Velocity from above crest level only.
adverse pressure measurement not 4 --- k=3.12mm Bars do not span whole
gradients. described channel width towards
downstream end.
3 Perry et al. Fundamental EXPERIMENTAL wind Mean velocity profiles
(1969) analysis of rough tunnel. Pressure tappings recorded over each of 12
k=3.18,
wall turbulent around bars, centreline Re*=250 to bars. Dye streaks and smoke
1.83, 3.6, 8.96 25.4mm,
boundary layers. velocities recorded 2500 used for flow visualisation.
/k=3 to 24
(method not given). Proposed d-type and k-type
roughness classifications.

PhD Thesis 17 Chapter 2.0


L J Campbell Literature Review
4 Raju & Garde Resistance to flow EXPERIMENTAL wind Provides comparison of
(1970) over transverse tunnel and hydraulic flume. hydraulic flume and wind
strip roughness. Prandtl tube for all k=20, 40mm, tunnel velocity data. However
velocities, pressure S=1/347, some wind tunnel rough walls
>1000 1/1333,
tappings in wind tunnel 2.5, 5, 10, 15, comprised only 4 elements.
(exact value 100 H 300
roughness elements. 20, 40 Roughness in flume made of
not given)
mm angle-irons (i.e. baffles
rather than square bars).
Centreline velocity profiles
from only 2 or 3 positions.
5 Antonia & Boundary layer EXPERIMENTAL wind Charts development of
Luxton (1971) response to step tunnel. Mean velocities internal boundary layer, in
change in from Pitot-static probe and terms of mean velocity and
roughness. single hot-wire, turbulence turbulent structure, following
from X-configuration hot- change from smooth to rough
wire probe, pressure- 4 96000 (est) k=3.12mm wall. Datum for rough wall
tapped elements for form taken as bar top level,
drag. d=2mm (downwards from bar
crests). Most velocity data
from above crests, few from
above troughs.
6 Webb et al. Heat transfer and EXPERIMENTAL pipe flow. Turbulent pipe flow. Ribs not
k=0.37, 0.74,
(1971) friction over rib- Thermocouples and square (0.38mm in
6000 to and 1.47mm
roughened resistance thermometers 10, 20, 40 downstream direction), air,
100000 D/k=25, 50, 100
surfaces. used for heat transfer. water and alcohol fluids
tested
7 Wood & Antonia Turbulence EXPERIMENTAL wind Groove roughness (i.e. not
(1975) characteristics tunnel. Mean velocity from upstanding). Velocity
k=3.18mm,
over smooth, Pitot-static tube and hot- measurements taken at
/k=16,
g
2 Re*=60
grooved and wire. Shear stress from midpoint of elements, no data
D/k=72.3
sand-grain rotated, inclined hot-wires. from below roughness crests.
roughened walls.

PhD Thesis 18 Chapter 2.0


L J Campbell Literature Review
8 Islam & Logan Response of a EXPERIMENTAL wind Rectangular duct flow.
(1976) turbulent tunnel. Mean longitudinal Experiments designed to
boundary layer to velocity from static 30800 investigate developing flow,
a smooth-to-rough pressure tappings, (over k=0.4mm, rather than equilibrium
38.4
surface change. micromanometer and upstream D/k=25.6 boundary layer flow. Data
impact tube, turbulence smooth wall) gathered from centreline at
intensity by constant- successive points half-way
current anemometer. between bars.
9 Siuru & Logan Turbulent flow EXPERIMENTAL air flow in Results focus on zone of
(1977) response to tubes. Velocities by Pitot- adjustment to rough bed flow,
change in static pressure probe, and k=2.9mm i.e. flow has not reached
2.21, 4.42,
roughness. velocities and turbulence (one case used equilibrium as the boundary
8.85,17.7, 60000
intensities by hot-wire k=2.0mm), layer is still developing.
37.66
anemometer, pressure D/k=22.1
drop by static pressure
probe.
10 Mulhearn Shear stress in EXPERIMENTAL wind Shows streamwise variation
(1978) deep turbulent tunnel. Pitot tube and of primary shear stress and
flow over periodic micromanometer for mean vertical velocity over 2 bar
square bars. velocities, turbulence pitches. Datum at level of bar
k=3.2mm
intensity from X- 4 --- tops. d=2mm (downwards
configuration hot-wire from bar crest) used for log-
probe. Surface shear linear fitting, zo=0.25mm.
measured by drag
balance.
11 Berger et al. Dependence of EXPERIMENTAL pipe flow. Pipe flow (square ribs in
(1979) local mass and Electric potential applied tubes) studied, fluid was
heat transfer and current measured 10000 to k=4mm electrolyte (NaOH). Flow not
3, 5, 7, 10
distribution on (analogous to heat flux). 250000 D/k=27.5 fully developed for spacings
square bar L=3, 5, and 7.
spacing.

PhD Thesis 19 Chapter 2.0


L J Campbell Literature Review
12 Bandyopadhyay Evaluation of EXPERIMENTAL - wind No velocity data from below
(1986) outer-layer tunnel. Electronic bar crests, point
1.61, 3.77 78000 ---
devices (ribbons) manometer (pressure), measurement only.
for drag reduction. Pitot tube (velocity).
13 Bandyopadhyay Rough wall EXPERIMENTAL wind Results pertain to the
(1987) turbulent tunnel. Mean velocities hydraulically smooth to rough
boundary layers from Pitot tubes and transition zone. Analysis
over square bar elctronic manometers, concentrates on the effect of
1.61, 3.77 --- k=0.31, 0.33mm
(and sandgrain) fluctuating velocity from spanwise aspect ratio (bar
roughnesses in single hot-wire probe, and length/height) on roughness
the transition drag force by drag Reynolds number for
regime. balance. transition.
14 Tani (1987) Turbulent REVIEW Square bar flows, together
boundary layer Review Article of mean with those over sand-grain,
g
development over velocity data with zero 1.61, 2 , 9.03, sphere and circular rod
--- k=0.33 to 5mm
rough surfaces pressure gradient over 10 roughness reanalysed to
various surface roughness. evaluate roughness
parameters.
15 Choi & Fujisawa Assessing the EXPERIMENTAL wind Velocities measured over a
(1993) possibility of drag tunnel. single square cavity only.
reduction over a Mean flow and turbulence k=10mm (and Reduction in skin friction
Re=10000
transverse square quantities from hot-wire
---
k=25mm for persisted 40 downstream
(est.),
cavity. probe, pressure tappings flow vis.), from cavity.
Re*>150
to assess form drag. Flow /k=2.5
visualisation (method not
described).
16 Liou et al.
Largely numerical simulation refer to Table 2.1, Part 3, article 1.
(1993a)

PhD Thesis 20 Chapter 2.0


L J Campbell Literature Review
17 Liou et al. Mean velocities EXPERIMENTAL wind Rectangular duct flow with
(1993b) and turbulence tunnel. Velocities via 2 square bars fixed to 2 walls.
characteristics in component LDV. Velocities collected from 23
k=4mm,
a channel with centreline locations spanning
10 33000 D/k=15,
ribbed walls. one bar pitch, 0.5k grid
B/D=2
spacing in vertical direction.
Additional off-centre
velocities obtained.
18 Matsumoto Features of a EXPERIMENTAL wind Groove roughness with
(1993) turbulent tunnel. Horizontal and X- square cross-section cavities
boundary layer in type hot-wire probes used (as for L=2 square bars), but
flows over for mean and fluctuating g g g elements are not square
10 , 20 , 30 , Re=2200 to
grooved rough velocities. g k=5mm (they are 9, 19, 29, and 39
40 6400
walls. times the cavity length). Skin
friction for certain
roughnesses reported as
lower than smooth wall.
19 Okamoto et al. Flow structure and EXPERIMENTAL wind Mean velocity profiles
(1993) turbulence tunnel. Pitot-static tubes recorded at 14 positions
10000,
augmentation (for for mean velocity, along tunnel centreline. No
7500 (for
heat transfer) over yawmeter for vectors, hot- quantitative velocity or
2, 3, 4, 5, 7, 9, heat
square bar wire anemometer for k=10mm turbulence data shown from
13, 17 transfer),
roughness. turbulence, thermocouples below level of bar tops. Bed-
500 (for flow
for heat transfer. Flow normal origin set at bar crest.
vis.).
visualisation done in water Most measurements from
channel. developing flow.
20 Djenidi et al. Compares EXPERIMENTAL closed- Velocities gathered from
(1994) turbulence circuit, constant head crest and groove midpoints,
characteristics vertical water tunnel. including data from below
over smooth and 3-component LDA used to g Re~1000 k=5mm, crests. Rough wall comprised
2
closely-spaced measure 2 component D/k=52 20 elements only (over 0.3m
bar-roughened velocities. Dye injection for length). Developing smooth
beds. flow visualisation. wall boundary layer before
square groove section.

PhD Thesis 21 Chapter 2.0


L J Campbell Literature Review
21 Djenidi et al. Structure of EXPERIMENTAL closed- Mean velocity profiles taken
(1999) turbulent circuit, constant head over successive cavities to
Re=900 to
boundary layer vertical water tunnel. show streamwise variation.
2300 (1100
over transverse 3-component LDV for g k=5mm, Variation of turbulent
2 for flow vis.),
square cavities, velocities. Flow /k=7 to 9 quantities over a whole
(Re*=124, at
and the interaction visualisation via LIF. roughness pitch presented.
of cavity and main Re=2300)
flows.
22 Keirsbulck et al. Surface EXPERIMENTAL wind Hot-wire velocities from one
(2002) roughness effects tunnel. Mean velocity and vertical only. PIV results not
on turbulent turbulence characteristics fully discussed and were
Re=8549 k=3mm,
boundary layer from single hot-wire and X- 3.33 presented as qualitative, not
D/k=200
structures. wire probes respectively. quantitative, velocity
Some supplementary PIV measure.
vector fields recorded.

Table 2.1, Part 1: Summary of closed-channel square bar studies

P
L=
k
P
lx

z
lz = k
x

Figure 2.5: Bed roughness geometry notation. L is the normalised roughness pitch, or
roughness spacing (roughness pitch, P, normalised with the roughness height, k).

PhD Thesis 22 Chapter 2.0


L J Campbell Literature Review
Table 2.1, Part 2: OPEN-CHANNEL FLOWS
(Hydraulic Flumes)

Data Source Purpose of Method Bar Spacing Reynolds Other Notes


Study (pitch to Number, relevant
height ratio), Re parameters
L
1 Liu et al. (1966) Flow structure in EXPERIMENTAL open Early investigation into
turbulent channel flow in hydraulic hydraulically smooth and
boundary layers flume. Streamwise velocity rough flows over limited
15600 to
with square-bar- and turbulence intensity by number of bar spacings.
54000, k=6.35mm,
roughened wall. hot-wire anemometry. 2, 4, 12, 24, 96 Most velocity profiles taken at
(Re*=48 to /k=7 to 13
Flow visualisation by level of downstream face of
125)
hydrogen bubbles and dye bar, for up to 5 bars. Bars are
injection. mounted on side-wall of open
channel, not bed.
2 Townes & Flow patterns EXPERIMENTAL hydraulic Mostly low Reynolds number
Sabersky from flow flume. Hot-film flows to aid visualisation.
(1966) visualisation in anemometry to assess Note that although slots have
square cross-section (k
g
turbulent flow over mean velocities, dye g Re*=12 to k=3.18, 6.35,
1.5 g
square slots. injection for flow 259 12.7, 25.4mm k ), like the L=2 square bar
visualisation. arrangement, bar width is just
half groove length. Rough
walls had 8-20 slots.
3 Raju & Garde
Largely closed-channel experiments refer to Table 2.1, Part 1, article 4.
(1970)

PhD Thesis 23 Chapter 2.0


L J Campbell Literature Review
4 Knight & Hydraulic EXPERIMENTAL open Velocities measured for 5
MacDonald resistance of strip channel flow in hydraulic vertical points at up to 9
3.47, 5.2, 6.93, 15 H / k 104
(1979a & roughness flume. Preston tube horizontal positions per
1979b) (1979a); Influence (stress), propeller meter
10.4, 13.9,
5150 to
1.5 B / H 15 experiment. All velocity data
20.83, 41.67, k=3mm,
of bed roughness (velocity). 92775 appears to be from above
83.33, 166.67, S=1/1043
and aspect ratio roughness crests. All
333.33
(B/H) on boundary reasonably high relative
shear (1979b); submergence (H/k).
5 Grass et al. Coherent motions EXPERIMENTAL open No velocity data from below
H=50mm,
(1991) in rough and channel flow in hydraulic bar crests, LDA for single
8076 k=5mm,
smooth wall flume. Hydrogen bubble 5 vertical profile.
(Re*=80) S(est.)=1/1890
turbulent and LDA (velocity and
ks/k=5.74
boundary layers. turbulence)
6 Lohasz et al.
Largely numerical simulation refer to Table 2.1, Part 3, article 3.
(2003)

Table 2.1, Part 2: Summary of open-channel square bar studies

PhD Thesis 24 Chapter 2.0


L J Campbell Literature Review
Part 3: NUMERICAL SIMULATIONS
(e.g. LES, DNS)

Data Source Purpose of Method Bar Spacing Reynolds Other Notes


Study (pitch to Number, relevant
height ratio), Re parameters
L
1 Liou et al. Turbulent heat SIMULATION/EXPERIMENTAL Air flow in rectangular ducts.
(1993a) transfer in a Pressure and velocity by Numerical results compared
channel with rib numerical simulation, 12600 to k=8mm (wind to previous LDV velocity
7.2, 10, 15, 20
roughness. temperature field from 60000 tunnel) data. 2D simulations also
wind tunnel experiment. tested ducts with ribs on 2
walls.
2 Cui et al. Turbulent flow SIMULATION - LES Closed channel (duct) flow
(2000, 2003) simulations with simulated with smooth top
square bars wall. Periodic conditions in
(2000, 2003) and streamwise and vertical
sinusoidal wave 2, 5, 10 10020 D/k=10 transverse directions (no
roughness (2000). side-wall effects). No slip
condition imposed on top and
bottom boundaries. . Grids
were 666690 (x,y,z).
3 Leonardi et al. Numerical study SIMULATION DNS. As above, closed channel
(2003) of velocity, form 3D velocitiy and pressure (duct) flow simulated with
drag and skin distributions resolved by smooth top wall and periodic
1.33, 1.6, 2, 4200
friction over numerical solution of conditions in streamwise and
3.07, 4, 5, 6.5, (Re*=190 to D/k=10
transverse square momentum and continuity spanwise directions (no side-
8, 9, 10, 11, 20 460)
bars. equations. wall effects). Flow rate kept
constant. Grids were
40097140 (x,y,z).

PhD Thesis 25 Chapter 2.0


L J Campbell Literature Review
4 Lohasz et al. Comparison of SIMULATION/ EXPERIMENTAL Square duct flow. LES code
(2003) LES and PIV data. Velocities from both PIV from commercial CFD
10 40000 (LES) D/k=3.33
and CFD software (Fluent v6). PIV flow
conditions not given.
5 Stoesser & Rodi Comparison of SIMULATION - LES Closed channel (duct) flow
(2004) LES and DNS for simulated. Finest LES grid
flow over 4200 generated was 26060120
3, 7, 10 D/k=10
transverse bars. (Re*=400) (x,y,z). Shows velocity
profiles from above both
crest and trough centres.

Table 2.1, Part 3: Summary of simulated square bar studies

Abbreviations and notation used in Table 2.1:

CFD=Computational Fluid Dynamics; B=channel width; Re=Reynolds number based


DNS=Direct Numerical Simulation; d=log-law displacement height; on momentum thickness;
LDA=Laser Doppler Anemometry; D=pipe diameter / simulation domain height / duct height; Re*=Roughness Reynolds
g
LDV=Laser Doppler Velocimetry; 2 , superscript refers to groove roughness (i.e. elements are not upstanding); number;
LES=Large Eddy Simulation; H=flow depth; S=channel slope;
LIF=Laser Induced Fluorescence; k=roughness height; zo=log-law roughness length;
PIV=Particle Image Velocimetry; ks=Nikuradses equivalent sand grain roughness height; =boundary layer thickness.

PhD Thesis 26 Chapter 2.0


L J Campbell Literature Review
Table 2.1 has been included to give a general idea of the existing square bar
literature. As such, the discussion of its contents will begin with the identification of
some broad trends (Sections 2.3.1.1 to 2.3.1.5).

2.3.1.1 Research Themes

Certain research trends emerge clearly from the contents of Table 2.1. The main
themes centre around, (1) assessment of bed shear, friction and hydrodynamics
(velocities and turbulence characteristics) in both closed- and open-channels (e.g.
Morris, 1955; Perry et al., 1969; Raju and Garde, 1970; Wood and Antonia, 1975;
Mulhearn, 1978; Bandyopadhyay, 1987; Matsumoto, 1993; Djenidi et al., 1999;
Leonardi et al., 2003), (2) quantifying heat and mass transfer in pipes and tubes
roughened by square elements (e.g. Webb et al., 1971; Berger et al., 1979; Okamoto
et al., 1993; Liou et al., 1993a), (3) validation and benchmarking of numerical
simulations (e.g. Leonardi et al., 2003; Lohasz et al., 2003; Stoesser and Rodi,
2004), and to a lesser extent, (4) the development of internal boundary layers
following a step-change in boundary roughness (e.g. Antonia and Luxton, 1971;
Islam and Logan, 1976; Siuru and Logan, 1977; Djenidi et al., 1994).

2.3.1.2 Closed-Channel versus Open-Channel Flows

Perhaps the most striking outcome from classifying experimentally-based square bar
studies as closed-channel or open-channel is the relative lack of experiments
conducted in hydraulic flumes. In fact, wind tunnel, pipe and duct flow articles
outnumber those on open-channel flows by almost 4 to 1. This is perhaps not entirely
surprising, as much of the past experimental effort has concentrated on the
qualitative and quantitative assessment of flows over practical engineering
roughnesses, which are usually found in pipes or ducts (e.g. Morris, 1955; Webb et
al., 1971; Berger et al., 1979). In this respect, Nezu and Nakagawa (1993) noted that
intensive research on the dynamics of wall turbulence in boundary layers, pipes and
closed-channels for air flows has been performed by many investigators over the last
forty years, i.e. since the 1950s, a much longer time than for open-channel
turbulence in flowing water.

PhD Thesis 27 Chapter 2.0


L J Campbell Literature Review
The fact that there are very few water-based open-channel experimental studies of
flows over square bars must not be overlooked, as there are certain key differences
between closed-channel or wind tunnel studies and those conducted in hydraulic
flumes. Aside from the arguably strong influence of different channel geometries (in
pipe flows for example), the lack of free-surface effects (closed-channel and wind
tunnel flows), and different fluids (air versus water), the most pertinent difference in
the context of this thesis is that air flows over rough walls in wind tunnels usually
have high relative submergence (e.g. Webb et al., 1971; Wood and Antonia, 1975;
Djenidi et al., 1994; Keirsbulck et al., 2002). In other words the roughness height is
much smaller than the flow depth. As wind tunnels generally provide deep
turbulent boundary layers, there is potentially a gap in knowledge concerning rough
flows with small relative submergence, i.e. the shallow open-channel flows that are
central to this study.

2.3.1.3 Velocity Measurement Techniques

The number of different experimental methods used to record fluid velocities listed
in Table 2.1 is testament to the steady development in measurement technology and
computing power over the last fifty years. The techniques range from early flow
visualisation methods (e.g. dye injection as used by Townes and Sabersky, 1966),
through many recognised point measurement devices (hot-wire probes, LDA and
LDV), right up to modern-day non-intrusive optical methods such as PIV that
provide a synchronous spatial description of the flow.

From the total of 27 experimental studies listed in Table 2.1, only two mention any
application of PIV to flows over square bars (those of Keirsbulck et al., 2002 and
Lohasz et al., 2003). However, in both cases PIV data is used to illustrate flow
features qualitatively rather than quantitatively. Therefore, despite PIV now being a
well established and reliable velocity measurement technique, it appears to have
been infrequently applied to flows over simple strip roughness. This most probably
reflects the fact that the toolbox of traditional analysis methods for dealing with
turbulent velocity signals (e.g. spectral analysis, quadrant theory) evolved alongside
experimental point measurement techniques. In doing so these analysis methods
were inherently geared towards the interpretation of high sampling frequency

PhD Thesis 28 Chapter 2.0


L J Campbell Literature Review
velocity data. Although PIV has the undoubted benefit of providing both spatial and
temporal descriptions of the flow field, it is as yet unable to compete with some other
velocity measurement systems (e.g. LDV) in terms of temporal recording frequency
(without compromising spatial resolution). Hence, existing PIV data is largely
unsuitable for the application of techniques such as spectral analysis. It is
nevertheless extremely well-suited to double-averaging theory.

2.3.1.4 Velocity Measurement Locations

The issue of where (in relation to the fixed bed topography) in the bed-parallel (x, y)
plane, and at how many points to measure air or water velocities over square bar
roughness is closely related to the above discussion of measurement techniques.
Clearly, gaining an adequate spatial representation of the velocity field using a point
measurement technique is laborious (and a single probe used in multiple positions
can never provide a synchronous description). Synchronous measurements can be
obtained using an array of probes, however, the use of multiple instruments
introduces practical limitations to spatial resolution. Most studies therefore rely on
gathering data from only a handful of verticals (refer to the notes column of Table
2.1). In essence this means that the nature and extent of the spatial variability in the
flow over strip roughness cannot be fully addressed using existing experimental
datasets.

Besides a lack of experimental data for describing variation in the bed-parallel plane
above bar roughness, there are few studies with measures of fluid velocity below the
roughness crests. The author believes that this may be related not only to technical
difficulties in obtaining reliable measurements close to a rough boundary, but
additionally to a lack of suitable theoretical framework within which to analyse
results gathered from different streamwise locations. Consequently, many studies
tend to neglect the flow region below the bar crests (e.g. Knight and MacDonald,
1979a,b; Grass et al., 1991), treat it entirely qualitatively (e.g. via flow visualisation,
Townes and Sabersky, 1966), or record near-bed velocities from the same position
within each bar pitch (e.g. at successive groove midpoints, Djenidi et al., 1994).

PhD Thesis 29 Chapter 2.0


L J Campbell Literature Review
There is however now a growing number of numerically-based studies (i.e. those
listed in Table 2.1, Part 3) which offer a full spatial description of local flow
velocities and pressure. The obvious disadvantage is that these are simulations which
must be validated with results from physical experiments. Whereas many features of
flow over square bars appear to have been successfully modelled (for example mean
velocities and reattachment lengths), due care must be taken in the interpretation of
simulation results. This is especially true when gaps exist in the underlying
experimental literature that is used for validation and benchmarking purposes.

Finally, it should be noted that, of the studies in Table 2.1 with access to adequate
spatial descriptions of the flow (Liou et al., 1993a; Cui et al., 2000, 2003;
Keirsbulck, 2002; Leonardi et al., 2003; Lohasz et al., 2003; Stoesser and Rodi,
2004), only one (Cui et al., 2000) contains any reference to spatial averaging (in the
longitudinal direction) but even this study neglects to explicitly express any
averaging method or operator. Hence, the recorded spatial variation over and around
square bars has not been properly quantified.

2.3.1.5 Roughness Spacing

Early researchers typically studied just one or a limited number of bar roughness
configurations (e.g. Morris, 1955; Perry et al., 1969). Even the more recent research
has often focussed on a just a few bar spacings (e.g. Grass et al., 1991; Djenidi et al.,
1999; Lohasz et al., 2003). There are, however, some exceptions which merit
particular mention. These are, for wind tunnel flows, Okamoto et al. (1993), for
open-channel flow, Knight and MacDonald (1979a,b), and for numerical simulation,
Leonardi et al. (2003). Nevertheless, it may be argued that there are still potentially
important gaps in the available range of bar spacings (L), most especially towards the
lower end of the range (denser arrangements) where one may expect the flow pattern
to change markedly with just a small change in bar pitch.

The above observations from Table 2.1 have therefore highlighted some general
trends from the spectrum of existing square bar roughness literature. For example,
the dominance in terms of the number of wind tunnel studies and closed-channel
flows over those conducted under open-channel conditions in hydraulic flumes is

PhD Thesis 30 Chapter 2.0


L J Campbell Literature Review
evident. A sub-group of the most pertinent studies (in the context of this project) is
now selected for more focussed discussion. The following sections therefore present,
(2.3.2) strip roughness classification, (2.3.3) near-bed flow patterns over square bars,
and (2.3.4) the resistance associated with square bar roughness.

2.3.2 Transverse Strip Roughness Classifications

There are two early (1950s and 60s) classifications of transverse strip roughness
that are still very much in use today, specifically the terminology suggested by
Morris (1955) and that of Perry et al. (1969). Both terminologies apply equally to
square bars and other cross-sectional shapes of transverse strip roughness.

2.3.2.1 The Strip Roughness Terminology of Morris (1955)

Morris (1955) reviewed studies of pipe flow over a variety of simple surface
roughnesses, including square bars. With the emphasis placed on clarifying the
friction factor for a range of rough wall flows, he proposed the following
terminology, based on the longitudinal strip roughness spacing, for three basic flow
types.

(a) Skimming flow

(b) Wake interference flow

(c) Isolated roughness flow

Figure 2.6: Basic flow types for flows over strip roughness as proposed by Morris (1955). (a)
skimming (or quasi-smooth) flow, (b) wake-interference flow, and (c) isolated roughness flow. All
figures adapted from ESDU 79014 (1979).

PhD Thesis 31 Chapter 2.0


L J Campbell Literature Review
Firstly, if the roughness elements are far apart, Morris observed that they would
behave as isolated bodies with the wake zone of each bar being completed developed
(and dissipated) before the next bar is reached. This was termed isolated roughness
flow and is illustrated in Figure 2.6(a). The second basic flow type occurs when the
bars are sufficiently close together such that the separation zone from each element
is not completely developed before the subsequent bar is encountered. Morris
labelled this flow type as wake interference flow [Figure 2.6(b)]. Finally, the third
flow type is called skimming flow [Figure 2.6(c)], and prevails when the bars are so
close together that flow essentially skims the crests of the elements, leaving an area
of slow, stable recirculation in each groove. This flow pattern was also called quasi-
smooth, as the main bulk of the flow moves over a pseudo-wall composed alternately
of the roughness crests and the upper limbs of the groove vortices.

The three-fold classification (or four-fold, including smooth-wall flow) of Morris


(1955) was subsequently extended by Knight and MacDonald (1979a) for the
purpose of clarifying the nature of flow close to their square-bar-roughened beds. On
the basis of dye injection visualisations, they suggested six basic flow types as
shown in Figure 2.7. According to this classification, isolated roughness flow is
renamed semi-smooth turbulent flow [Figure 2.7(b)], skimming flow is referred to
as quasi smooth flow [Figure 2.7(f)], and three new flow types replace what Morris
termed wake interference flows. These are non-uniform and uniform hyperturbulent
flows (Figures 2.7(c) and (d), respectively), and semi-quasi-smooth flow [Figure
2.7(e)]. The distinction between non-uniform and uniform hyperturbulent flows is
that in the non-uniform case the wake from one roughness element only just reaches
the upstream edge of the next element, whereas in the uniform regime the wakes
from successive elements intermingle. Knight and MacDonald (1979a) proposed that
the latter therefore produced a uniform layer of turbulence above the bed. The semi-
quasi-smooth flow class was included in recognition of the transition between quasi-
smooth and hyperturbulent flows. In this flow type there is a trapped vortex
sheltering in the lee of a roughness element, but it does not quite fill the groove as in
the quasi-smooth pattern.

PhD Thesis 32 Chapter 2.0


L J Campbell Literature Review
(a)

(b)

(c)

(d)

(e)

(f)

Figure 2.7: The square bar roughness spacing classification of Knight & MacDonald (1979a).
Including the relevant bar spacings studied by Knight & MacDonald (1979a), these are, (a)
smooth turbulent flow (L=), (b) semi-smooth turbulent flow ( 333 L 13.9 ), (c) non-uniform
hyperturbulent flow (L=10.4), (d) uniform hyperturbulent flow (L=6.95), (e) semi-quasi-smooth
flow (L=5.21), and (f) quasi-smooth flow (L=3.47). Figure adapted from Knight & MacDonald
(1979a). L is as defined in Figure 2.5.

Whilst the basic terminology of Morris (1955) is more frequently used today, the
work of Knight and MacDonald (1979a) is particularly relevant to this study in the
sense that it highlights the importance of small changes in roughness pitch across the
wake-interference region of strip roughness spacing.

PhD Thesis 33 Chapter 2.0


L J Campbell Literature Review
2.3.2.2 The Strip Roughness Terminology of Perry, Schofield and Joubert (1969)

In addition to the terminology of Morris (1955), it is widely accepted that static strip
roughness can be classified as either k-type in which the roughness elements are
widely spaced and essentially operate as single units, or as d-type in which the
roughness elements are more tightly grouped so that the flow disturbances around
them interact. This is the result of Perry et al. (1969).

Results of Perry
et al. (1969)

Figure 2.8: The results of Perry et al. (1969) for k-type square bar roughness. Graph shows

u / u versus ku / , where u is friction velocity (= u* ) and u / u is the roughness function.

The sketch below the graph shows typical near-bed flow patterns for a k-type bar configuration.
(Both figures have been adapted from Perry et al., 1969)

Perry et al. (1969) reported that, for wind tunnel flow over widely spaced bars, the
roughness function ( u / u* , where u is the shift in velocity profile from the
equivalent smooth bed flow) scaled with roughness height, k. This roughness class
was hence termed k-type roughness. This behaviour is illustrated by the graph in
Figure 2.8, which also shows typical flow patterns over k-type roughness as
proposed by Perry et al. (1969). Alternatively, when the rough bed comprised

PhD Thesis 34 Chapter 2.0


L J Campbell Literature Review
closely spaced bars (L = 1.83) the significant length scale involved in determining
the roughness function was the boundary layer thickness or pipe diameter, d, and not
the roughness height. This roughness was therefore denoted d-type roughness, and
is illustrated in Figure 2.9

Figure 2.9: The results of Perry et al. (1969) for d-type square bar roughness. Main plot shows

u / u versus c u / , where u is friction velocity (= u* ), u / u is the roughness function, and

c is the boundary layer thickness (after Clauser, 1954). Small inset graph shows u / u versus

ku / - note the lack of simple linear trend as in the main figure. The sketch below the graph

shows typical near-bed flow patterns for a d-type bar configuration. (All figure parts have been
adapted from Perry et al., 1969)

In terms of the near-bed flow patterns, Perry et al. (1969) proposed that when the
square bars are closely spaced (d-type), stable vortices occupy the whole inter-
element gap and effectively isolate the outer flow from the full influence of the wall
roughness. The steady vortex makes the momentum exchange across the opening of
the cavity almost independent of the roughness scale (Tani, 1987). Consequently
eddy shedding into the overlying flow is negligible and the outer flow behaviour is
largely independent of roughness height, k. Conversely, when the roughness bars are
less densely spaced (k-type) the flow has a sufficient length to become reattached to

PhD Thesis 35 Chapter 2.0


L J Campbell Literature Review
the wall ahead of the next rib, thus fully exposing the outer flow to each roughness
element, and vice versa. In the case of k-type roughnesses eddies with length scale
comparable to the roughness height are shed into the main core of the flow.

There is however an apparent confusion in the square bar literature as to the precise
delineation between d- and k-type rough surfaces. For example, Bandyopadhyay
(1986, 1987) only considers d-type behaviour for groove width to height ratios less
than unity (i.e. for L 2 with square bars) and therefore his wider bar spacing (L =
3.77) is reported as a k-type roughness configuration. Alternatively, other researchers
(e.g. Tani, 1987; Cui et al. 2000) state that the transition from d-type to k-type
behaviour lies around L = 5. It may simply be that Bandyopadhyay (1986, 1987)
strictly adhered to the d- and k-type roughness classification on the basis of altered
flow hydrodynamics (i.e. the roughness function behaviour reported by Perry et al.,
1969), whereas a looser interpretation founded more on the near-bed flow patterns
(and hence closer to the terminology of Morris, 1955) was adopted by the likes of
Tani (1987) and Stoesser and Rodi (2004). As will be seen in Chapter 7, the latter
interpretation is adopted for this thesis with the line of demarcation between d- and
k-type roughness being drawn at the point where square bars begin to behave as
isolated elements. For clarity, isolated roughness shall be defined as roughness
configurations that allow mean flow reattachment to the solid bed before
encountering the next bar.

2.3.3 Streamlines and Reattachment Length for Flows over Square Bars

This thesis is primarily concerned with near-bed flow, specifically in quantifying the
spatial variation in relation to a range of fixed square bar topographies. Therefore,
Figures 2.10 to 2.12 review some existing measures of mean flow patterns in this
region. Figure 2.10 shows streamline patterns evaluated by smoke and dye flow
visualisation techniques (Perry et al., 1969), and reveals that although the stable
cavity vortex over the densely arranged bars (L = 1.83) is well resolved, when the
bars are further apart (L = 3.60) there is a lack of flow information towards the centre
of the trough because the exchange between the cavity and the main flow is greatly
increased. The results from the improved experimental technique illustrated in
Figure 2.11 (that used by Okamoto et al., 1993) clearly show large vortices filling

PhD Thesis 36 Chapter 2.0


L J Campbell Literature Review
the entire gap between bars for both spacings L = 2 and 5, and ample length for
reattachment to the channel bed when the bar spacing is extended to L = 9. These
patterns have been subsequently confirmed, for example by the numerical
simulations (LES) of Cui et al., 2003 (shown in Figure 2.12).

(a)

(b)

Figure 2.10: Flow patterns from the early flow visualisation wind tunnel experiments of Perry et
al. (1969). The two cases shown are for bar spacings, (a) L=1.83, and (b) L=3.60.

(a) (b) (c)

Figure 2.11: Photographs of vortex flow patterns in a water channel (Reynolds number
approximately 500). Pictures come from Okamoto et al. (1993) and are for spacings (a) L=2, (b)
L=5, and (c) L=9. Note how flow over L=5 does not have sufficient groove space to reattach to
the solid bed, but clearly does so when the bars are further apart in spacing L=9.

Reattachment lengths for flow over an obstacle of height k reported in the literature
are surprisingly varied. Most studies record a reattachment length of around 4 to 5
roughness heights downstream from the downstream edge of the obstacle. For
example, the simulations (DNS) of Leonardi et al. (2003) show reattachment at 4.8k
for L = 8, which is in good agreement with LES modelling results of Stoesser and
Rodi (2004) and those of Cui et al. (2000) who report 4.3k for L = 10. These
numerical results are also supported by earlier experimental findings, for example 5k
for L = 8 and L = 12 (Liu et al., 1966), 4.32k for L = 7.2 (Liou et al., 1993a), and 4k

PhD Thesis 37 Chapter 2.0


L J Campbell Literature Review
for L = 9 (Okamoto, 1993). These results are from studies of flows over repeated
bars, where the adverse pressure gradient arising as a consequence of the backwards-
facing edge of the next roughness element will influence the length of separated
flow. Therefore, in the absence of subsequent downstream obstacles the reattachment
length may be longer than the above estimates. This is obviously also applicable if
the roughness elements are sufficiently far apart. Hence, Liu et al. (1966) recorded a
reattachment length of 6k for L = 24, a figure supported by recent numerical
simulations (Leonardi et al. (2003) found flow reattaches at 5.8k for L = 20).

(a) (b)

Figure 2.12: Mean streamlines (c)


from the LES simulations of Cui et
al. (2003), for square bar
spacings (a) L=2, (b) L=5, and
(c) L=10.

Notwithstanding the variable influence of repeated obstacles, a few studies report


appreciably shorter or longer reattachment lengths than 4k to 5k. Amongst these,
Liou et al. (1993b) state reattachment occurs at 3.5k for L = 10 (this is for closed
channel flow with ribs on two walls), and the CFD modelling of Lohasz (2003)
suggests lengths ranging between 3.2k and 4.2k, again for L = 10. Berger et al.
(1979) assume that reattachment coincides with the local maximum mass transfer
coefficient, which is recorded at just 3.2k for spacing L = 10, whilst longer than
average reattachment lengths of 6k to 8k are suggested by Webb et al. (1971). The

PhD Thesis 38 Chapter 2.0


L J Campbell Literature Review
scatter between 3.2k to 8k suggests that defining reattachment length on the basis of
geometry alone (i.e. the roughness height, k) may be incomplete. In reality, the
impinging flow velocity may also play a role in governing the reattachment length,
as given that it is a wake phenomenon it is likely to be Reynolds number dependent
to some degree.

Neglecting the extreme values, it is reasonable to take the standard reattachment


length to be approximately 5k. As discussed above however, this generally applies to
flows over single elements. When considering the likelihood of flow reattachment to
the bed in cases where the roughness comprises repeated obstacles of height k, it is
necessary to account for areas of leading edge separation [i.e. the small recirculation
zones on the upstream face of each element, clearly evident in Figures 2.11(c) and
2.12(c)]. Therefore, assuming that flow will reattach over a distance of 5k in the
absence of downstream obstacles, and factoring in a further 1k for the leading edge
separation zone as a result of subsequent elements [observing that this region tends
to scale with roughness height, being at most one bar height in length as in Figure
2.12(c)] means that the flow requires a total gap length of around 6k to potentially
reattach to the channel bed, albeit briefly. In essence, L = 7 is therefore the closest
bar spacing which may genuinely be considered to represent isolated roughness flow
(confirmed by the simulations of Leonardi et al., 2003). As will be seen when
discussing the results of the present study, these lengths of flow separation and
reattachment play a crucial role in determining the behaviour of flows over
transverse square bars.

2.3.4 The Resistance of Transverse Square Bar Roughness

The resistance offered by transverse square ribs has been primarily studied from two
distinct angles. Firstly, there is a considerable body of research devoted to
elucidating friction factors over simple roughness geometries (e.g. Morris (1955) for
pipe flow, and Knight and MacDonald (1979a) for open-channel flow). Secondly,
much research effort has focussed on the evaluation of the optimal longitudinal
spacing for maximum heat and mass transfer (e.g. Okamoto et al., 1993; Liou et al.,
1993a).

PhD Thesis 39 Chapter 2.0


L J Campbell Literature Review
Morris (1955) was one of the early investigators to realise that the effects of surface
roughness (in terms of friction factor) could not be successfully addressed solely
with Nikuradses sand grain roughness scale. Consequently, he suggested that for
transverse strip roughness the longitudinal roughness spacing was an important
parameter. Indeed, the threefold classification of bar roughness discussed above (i.e.
isolated roughness, wake interference, and skimming flow types, Section 2.3.2.1)
was founded on the underlying mechanisms that Morris believed to be contributing
the majority of friction in pipe flows.

In the case of isolated roughness flow, Morris (1955) proposed that the apparent
friction factor would result from form drag around individual elements together with
a viscous drag contribution from the areas of smooth surface between the ribs
outwith the wake zones. The magnitude of form drag depends on the roughness
height, k, and the roughness pitch, P. Morris (1955) called the ratio P/k the
roughness index (which is equivalent to L in the current notation) and used this,
together with the drag coefficient (CD = 1.9 was used for rectangular strip roughness)
as a multiplier for the equivalent smooth conduit friction factor to obtain the rough
wall friction factor for isolated roughness flows. Subsequently, Sayre and Albertson
(1961, 1963) studied 3-dimensional arrangements of sheet metal baffles (i.e. the
roughness elements were not continuous in the transverse direction) and concluded
that between 86% and 98% of the total resistance was attributable to form drag,
depending on roughness density. It is probable that the upper end of this estimate is
more applicable when the roughness elements are laterally continuous.

As the roughness spacing decreases and overlapping wakes fill the gap between bars
(wake interference flows), Morris (1955) suggested that roughness height becomes
relatively unimportant and the effect of viscous drag on the channel wall is
eradicated. Under these conditions the pipe radius, r (as it determines the proportion
of the pipe flow area available for normal turbulence, as opposed to what Morris
termed abnormal (i.e. wake) turbulence), and the roughness pitch, P, were
proposed as the key variables correlating with friction factor. The ratio r/P was
named the relative-roughness spacing.

PhD Thesis 40 Chapter 2.0


L J Campbell Literature Review
Ultimately, decreasing roughness pitch means the transverse bars become
sufficiently close to generate skimming flow and the main bulk of the flow moves
over a pseudo-wall. Morris hypothesised that most of the energy expenditure of
this type of flow could be attributed to the maintenance of the steady groove
vortices. As the size of the vortices scales with k, and the number (per unit length of
channel) depends on the pitch, P, the roughness index, L = P/k, was again promoted
as the key correlating parameter. For all three flow types, Morris concluded excellent
agreement between existing friction factor data and those predicted using his newly
developed resistance formulae.

(a)
K & McD (1979a)

(b)

K & McD (1979a)

Figure 2.13: The variation in friction factor, f, with longitudinal roughness spacing, L. (a)
compares the relationships proposed by Morris (1955) and Knight & MacDonald (1979a) (K &
McD (1979a)), whilst (b) includes the results of Adachi (1964), Raju & Garde (1970), and
Sayre & Albertson (1963). (Both plots have been adapted from Knight & MacDonald, 1979a).

In addition to presenting results from their extensive study of square bars in

PhD Thesis 41 Chapter 2.0


L J Campbell Literature Review
hydraulic flumes, Knight and MacDonald (1979a) revisited the friction factor
equations of Morris (1955) and others (Sayre and Albertson, 1963; Adachi, 1964;
Raju and Garde, 1970). Indeed, it was Adachis observation (1964) that bed
resistance peaks at L = 8 which prompted Knight and MacDonald (1979a) to divide
the wake interference flow class into non-uniform and uniform hyperturbulent flow
types (as illustrated by Figure 2.7). Figure 2.13 shows two of Knight and
MacDonalds (1979a) plots demonstrating how the proposed resistance equations
(from each of the reviewed articles) predict friction factor.

From Figure 2.13(a) it is clear that the friction factor values of Morris (1955) peak at
a wider bar spacing, L = 13.9, compared to the value of L = 8.2 predicted by Knight
and MacDonald (1979a). There is also considerable variation in the predicted
roughness spacing for maximum bed resistance from Figure 2.13(b). These values
are L = 10.2 (Adachi, 1964), L = 8.2 (Sayre and Albertson, 1963), and L < 10 or L =
13.3 (Raju and Garde, 1970). Following this analysis, Knight and MacDonald
(1979a) adopted L approximately equal to 8 as the roughness spacing value for peak
bed resistance.

More recently, numerical simulations have proved useful in the investigation of bed
resistance in that they provide both pressure and velocity information that can be
used to assess contributions from viscous and form drag. Leonardi et al. (2003,
DNS) report that for L 3 the total drag is closely approximated by the viscous drag
on the crests of the elements, whilst over a range of isolated roughness flows (they
modelled spacings up to L = 20) the total drag is almost exclusively generated by
form drag. Maximum form drag was found to occur at roughness spacing L = 8,
which coincided with the configuration for minimum viscous drag. Cui et al. (2000,
2003, LES) also partition the total drag and state that for their skimming flow or d-
type configuration, L = 2, form drag is double the viscous drag. For bar spacings L =
5 and L = 10 form drag dominates further to contribute over 90% of the total
resistance.

As mentioned at the start of this section, another approach to the resistance of


transverse strip roughness has been via heat- and mass-transfer considerations. Put
simply, the impetus for such work is that increasing surface resistance augments

PhD Thesis 42 Chapter 2.0


L J Campbell Literature Review
turbulent transfer, thus enabling efficient heat exchange either to or from the fluid as
it moves over a rough wall. Amongst these studies, Okamoto et al. (1993) examined
wind tunnel turbulence and heat transfer over bar spacings 2 L 17 (specific
arrangements are listed in Table 2.1). They found L = 9 to be the optimum pitch ratio
to augment the longitudinal turbulence intensity (and hence heat transfer) with the
position of maximum local heat transfer coefficient closely coinciding with the flow
reattachment point. This link between optimal local heat transfer coefficient and the
reattachment point was also reported by Berger et al. (1979), whilst Liou et al.
(1993a) observed that arrangements L = 7.2 and L = 10 enhance heat transfer by
similar amounts, with lesser heat transfer over L = 15.

There is therefore reasonable variation in estimates for the optimum bar pitch to give
peak bed resistance, with reported values from L = 8.2 (Knight and MacDonald,
1979a) to L = 13.9 (Morris, 1955). This may reflect the different origins of these
estimates, with data coming from wind tunnels, pipes and hydraulic flumes and
flows with different aspect ratios and relative submergence. It is however clear that
an isolated roughness element will exert more drag on the flow compared to one in a
denser skimming or wake interference arrangement.

The above review of strip roughness resistance concludes the discussion of flow over
square bars. The remainder of this chapter considers the use of double-averaging
theory, with particular emphasis on the application of double averaging theory to
open-channel flows. As the entire following chapter is dedicated to the derivation
and discussion of the double-averaged momentum and mass conservation equations,
Section 2.4 is relatively short. Furthermore, when one considers that the use of
double-averaging for open-channel flows has only been developing for a few years,
it becomes clear that there is nowhere near the extensive body of research as exists
for square bar strip roughness.

2.4 Double-Averaging Methodology

Double-averaging refers to the process by which the governing equations of fluid


flow, or more correctly the fluid velocity and pressure terms contained within, are
averaged in both temporal and spatial domains. This double-averaging procedure is

PhD Thesis 43 Chapter 2.0


L J Campbell Literature Review
far more than a convenient statistical data-smoothing exercise. Nikora et al. (2001)
noted that, in some cases, when these equations are just temporally averaged in the
conventional manner (i.e. to yield the Reynolds averaged Navier-Stokes or RANS
equations) they were subsequently artificially modified by introducing form drag as
an extra body force term in order to account for the effects of boundary roughness on
the near-bed flow. In contrast, and as will be seen in Chapter 3, supplementing
temporal averaging with spatial averaging naturally produces explicit, physically
significant terms describing form drag, viscous drag, and form-induced momentum
fluxes.

Although both form drag and viscous drag nomenclature is well known, form-
induced momentum flux may not be as it is specific to double-averaging
methodology. The physical meaning and origin of the form-induced flux term will
become clearer (in Section 2.4.4, and when the double-averaged equations are
derived in Chapter 3), but it is helpful to include a simple introduction here. Just as
the process of temporal averaging produces the turbulent (Reynolds) stress term in
the RANS equations, subsequent application of the spatial averaging operator
generates the form-induced stress term. It stems from the correlation between local
disturbances in the time-averaged flow (just as Reynolds stress is the correlation
between velocity fluctuations about a temporal mean), which occur as a direct result
of surface roughness. In other words the local disturbances are form-induced.

2.4.1 The Development of Double-Averaging Theory

Double-averaging methodology owes its origins to multi-phase and ground-water


flow hydrodynamics, fields in which it has been developing since the sixties (e.g.
Whitaker, 1967; Gray and Lee, 1977). However, in the field of hydraulics the idea of
double averaging was first introduced by Smith and McLean (1977) for analysing
flow over large (up to 100m long and 3m high) dunes in the Colombia River, USA.
They implemented spatial averaging of mean velocities along lines of constant
distance from the wavy bed. In the same year, Wilson and Shaw published details of
newly developed mean momentum and stress equations describing airflow through
vegetative canopies, thus initiating the use of double-averaging methodology for
atmospheric flows. Raupach and Shaw (1982) subsequently formalized the spatial

PhD Thesis 44 Chapter 2.0


L J Campbell Literature Review
averaging procedure and described the key properties of the averaging operator. Both
Wilson and Shaw (1977), and Raupach and Shaw (1982) considered area averages
over a horizontal plane intersecting numerous plants. Later, Finnigan (1983, 1985)
generalized the averaging technique to consider volume averages within a flexible
plant canopy. Both methods are equivalent; if spatial-volume averaging is applied to
an extensive, infinitesimally thin horizontal slab it collapses to the spatial-area
averaged result (specific benefits of volume as opposed to area averaging are
discussed at the start of Chapter 3).

Since this early development, the methodology has been adopted and further
enhanced by contributions from atmospheric physicists largely concerned with the
analysis and/or modelling of turbulent flow through a range of vegetation types,
from wheat fields to forests (e.g. Raupach and Thom, 1981; Raupach et al., 1986,
1991; Brunet et al., 1994; Ayotte et al., 1999; Finnigan, 2000; Bohm et al., 2000;
Poggi et al., 2004a,b,c). Over recent years there have also been many studies
dedicated to the clarification of flow close to more urban-like surfaces,
predominantly comprising different arrays of cubes (e.g. Rotach, 1993; MacDonald,
2000; Cheng and Castro, 2002; Kanda et al., 2004; Xie et al., 2004; Lien and Yee,
2004, 2005; Lien et al. 2005). Most of these urban roughness studies automatically
include double-averaging methodology as it is now well accepted in the field of
atmospheric physics. Furthermore, it follows that because double-averaging is
readily implemented for such boundary layer investigations, the velocity
measurements tend to be tailored accordingly. Hence, although there are few square
bar studies with adequate spatial descriptions of velocity, there are a growing
number of cube roughness data sets, many simulated, more suitable for evaluating
the heterogeneity of near-surface flow.

Following the early work of Smith and McLean (1977), there have been relatively
few attempts to apply double-averaging methodology to water flows. Lopez and
Garcia (2001) used insights from double-averaging to model open-channel flow
through submerged vegetation, whereas McLean et al. (1999) and Maddux et al.
(2003) average the flow field over 2-dimensional and 3-dimensional dunes
respectively. Gimenez-Curto and Corniero Lera (1996) considered oscillating
turbulent flow over rough surfaces and introduced the term form-induced stress to

PhD Thesis 45 Chapter 2.0


L J Campbell Literature Review
describe the spatial disturbances in the time averaged flow. Atmospheric physicists
generally use the term dispersive stress.

With reference to this thesis, the two principal double-averaging studies have been
by Nikora et al. (2001, 2004) as both focus on the application of the methodology to
open-channel flows. The pertinent results of these, and other papers cited in this
section are now discussed.

2.4.2 Double-Averaged Open-Channel Flow

The shift to using double-averaging for the analysis of open-channel flows has been
a recent one, since Nikora et al. (2001) suggested that the double-averaged
momentum equations provided a natural basis for the hydraulics of rough open-
channel flows, particularly those with small relative submergence. In addition to the
inclusion of form-induced stress and surface drag terms, the suggested benefit of this
approach was that the double-averaged flow parameters (velocities and pressure, and
their moments) could be related to roughness parameters obtained by averaging in
the same spatial domain. In this respect, the roughness geometry appears explicitly in
the double-averaged equations via what Nikora et al. (2001) term the roughness
geometry function, A = Af /Ao (where Af is the area occupied by fluid within the total
averaging area, Ao). This is clearly a measure of (areal) porosity, and may equally
well be denoted by (as for the derivation of the volume-averaged equations in
Chapter 3). Previous experimenters (Wilson and Shaw, 1977; Raupach and Shaw,
1982; Finnigan, 1985) had made the assumption that the fluid averaging region, Af,
does not depend on elevation, as the roughness associated with vegetation canopies
can be approximated as vertically homogeneous. For complex roughnesses such as
those found in gravel-bed rivers this assumption is not valid. Instead the ratio Af /Ao
varies with the bed-normal coordinate, z, and typically decreases (from A = 1, see
Figure 2.14) from the roughness crests downward, hence A = A(z).

2.4.2.1 Layers in Double-Averaged Open-Channel Flow

The open-channel flow layers proposed by Nezu and Nakagawa (1993) as discussed

PhD Thesis 46 Chapter 2.0


L J Campbell Literature Review
in Section 2.2.2) are now revised with specific reference to the double-averaging
methodology proposed by Nikora et al. (2001).

It is common to refer to a roughness sub-layer as the hydraulically rough analogy


to the viscous sub-layer in hydraulically smooth flow. The term roughness sub-
layer describes the entire layer of flow dynamically influenced by length scales
associated with roughness elements (Raupach et al., 1991). Nikora et al. (2001)
suggested a further division by splitting this layer into a form-induced sub-layer and
an interfacial sub-layer. Figure 2.14 depicts the division of the entire depth of a free-
surface flow with high relative submergence (i.e. where the mean flow depth, H, is
much larger than roughness height, k) over both permeable and impermeable beds, as
presented by the above authors. Working from the free-surface down to the bed, the
general features of each layer may be described as follows.

1) The Outer Layer


Flow behaviour in the outer layer displays negligible dependence on viscous effects
or form-induced fluxes. Consequently, Nikora et al. (2001) note that double-
averaging the Navier-Stokes equations in this zone provides no additional
information about the flow. This layer covers the near-surface zone of flow and is
similar to the outer layer for flows over hydraulically smooth beds. It corresponds to
the free-surface region in the terminology of Nezu and Nakagawa (1993, Section
2.2.2). The outer layer is a region of relatively weak shear as the velocity gradients
are very much less than when close to the wall; hence in both smooth and rough wall
boundary-layers the turbulence structure in this region is similar. In each case the
outer region is not the site of the dominant instability mechanisms that generate
turbulence, nor is it the site of the strongest production of turbulent kinetic energy.
As discussed, characteristic scales for this outer layer are the maximum velocity, Uo
(assumed to occur at the free surface), and the flow depth, H.

2) The Logarithmic Layer


Like the outer layer, viscous effects and form induced fluxes are negligible in this
layer; the double-averaged equations are therefore identical to temporally-averaged
ones (Nikora et al., 2001). This layer corresponds to the upper part of the wall
region as described in Section 2.2.2. Raupach et al. (1991) conclude that the

PhD Thesis 47 Chapter 2.0


L J Campbell Literature Review
characteristic scales for this layer are friction velocity, u* , distance above the bed, z,
and a set of surface length scales including characteristic roughness height, k, and all
necessary additional lengths Li. (where Li is a set of lengths with numerous
components, for example the roughness element dimensions in the x and z directions,
lx and lz respectively).

For this layer to fully exist the requirement that the flow depth is much greater than
the typical roughness height must be met (H >> k). Given this, flow in this layer is
similar to the log layer over hydraulically smooth beds, and typically occupies the
region (2-5)k < z < 0.2H.

z a
Water surface
zws
Outer layer
zL
Logarithmic layer
zR Flow
Form-induced sublayer Roughness
zc Interfacial sublayer layer Flow
Type I
A=1 R Type III Flow
Flow Type II
A=0 Type IV

zt
A (z)
0 1
z b
Water surface
zws
Outer layer
zL
Logarithmic layer
zR Flow
zc Roughness Form-induced sublayer Type I
layer Interfacial sublayer Flow
zt Flow
Type III Flow
Type II
Subsurface layer Type IV
dA
zf =0
dz
A(z)
0 Amin 1

Figure 2.14: Subdivision of double-averaged open-channel flow into specific layers (adapted
from Nikora et al., 2001), for (a) impermeable bed, and (b) permeable bed. A is the roughness
geometry function (porosity) = Af /Ao. Vertical coordinate subscripts are: zws=elevation of water
surface, zL=upper bound of logarithmic layer, zR=upper bound of roughness layer or lower
bound of logarithmic layer, zc=level of highest roughness crests, zt=level of roughness troughs
where A=0 (impermeable bed) or dA/dz=0 (permeable bed), and zf=level of solid floor. Different
flow types, I to IV, are suggested on the basis of relative submergence.

PhD Thesis 48 Chapter 2.0


L J Campbell Literature Review
3) The Form-Induced Sub-Layer
This region constitutes the upper zone of the roughness layer (previously termed
roughness sub-layer), and lies between the logarithmic layer above, and the
interfacial sub-layer below. It sits just over the crests of the bed roughness, which, as
the name suggests, influence flow by causing separation and vorticity around
individual elements. The result is a region typified by marked local variation in the
temporally-averaged flow field. Hence, the form-induced stresses make an important
contribution to the momentum balance (Nikora et al., 2001). Raupach et al. (1991)
reported that this layer might extend up to 5k above the roughness crests. The
suggested characteristic scales for this region are the friction velocity, u* , and the set
of surface length scales chosen to allow as full a description of the bed topography as
possible.

The time-averaged Navier-Stokes equations (RANS) are adequate for describing


flow in the outer and logarithmic layers, as spatial differences across the fluid field at
this level can be neglected. This is not true in the roughness layer where the flow at
any point is highly influenced by the local spatial variations.

4) The Interfacial Sub-Layer


Further reference to Figure 2.14 shows that the interfacial sub-layer lies in the region
bounded by planes at the level of the crests and troughs of the bed roughness. Like
the form-induced sub-layer, this zone is also heavily influenced by individual
roughness elements. Importantly, at this level the double-averaged equations
explicitly account for form and viscous drag forces (as will be seen in Chapter 3).

Taken together, the interfacial and form-induced sub-layers constitute the roughness
layer. The roughness layer and the logarithmic layer can then be interpreted as the
wall region as identified earlier in Section 2.2.2.

5) The Sub-Surface Layer


If flow over a permeable bed is to be considered, it is necessary to include an
analysis of fluid moving in the sub-surface pore spaces between granular roughness
elements. Such flow is driven by the gravity force and fluid motion in the overlying

PhD Thesis 49 Chapter 2.0


L J Campbell Literature Review
layers. Like the roughness layer, the characteristic scales are the shear velocity, u* ,
and relevant length scales associated with the pore dimensions.

Following this double-averaging approach, Nikora et al. (2004) define four types of
rough-bed flows depending on a combination of flow depth and roughness height
(i.e. the relative submergence, H/k). With reference to Figure 2.14, Type I has high
relative submergence and contains all of the above layers. Type II has intermediate
relative submergence and will exhibit subsurface (where applicable) and roughness
layers, and an upper region that will not manifest logarithmic behaviour as the ratio
H/k is not large enough (e.g. Raupach et al., 1991). Type III has small relative
submergence and is typified by a roughness layer extending right up to the water
surface, whilst Type IV is characteristic of flow over a partially inundated rough bed
such that the interfacial sub-layer is the uppermost flow region.

Now that the flow field has been re-defined within the framework of double-
averaging, attention turns to the potential form of the double-averaged streamwise
velocity profiles in the region affected by surface roughness.

2.4.3 Distribution of Double-Averaged Velocity in the Near-Bed Region

As intimated in the opening quote, the (temporal) mean velocity distribution in the
roughness layer will depend on the measurement location relative to the surface
irregularities. For example, Figure 2.15 shows simulated streamwise velocity data
from Cui et al. (2003) for flow at various positions over an isolated roughness or k-
type (L = 10) square bar arrangement. This serves to illustrate that the mean velocity
distribution is influenced by the roughness throughout the bottom 50% of the duct,
such that it is unclear what the representative profile should be below the level of the
logarithmic layer (Figure 2.14).

Nikora et al. (2004) suggested three potential forms (models) for the double-
averaged streamwise velocity distribution in the lower part of the roughness layer,
the interfacial sub-layer. These were the exponential profile, constant velocity, and a
linear profile, although it may be argued that the constant velocity profile is simply a

PhD Thesis 50 Chapter 2.0


L J Campbell Literature Review
special case of the linear distribution, with zero gradient, d u / dz = 0 . (Note that
the straight overbar signifies temporal averaging, whereas angle brackets imply
spatial averaging).

Figure 2.15: The roughness-induced variation in the temporally-averaged streamwise velocity


profiles (from Cui et al., 2003). Data come from simulated flow over a k-type square bar
roughness. The variation is noticeable throughout the bottom half of the duct. Velocity data are
normalised with bulk velocity (Ub), w is interbar gap length (=P-k), and H is duct height.

Based on simple phenomenological considerations following from double-averaging


methodology, Nikora et al. (2004) proposed that the exponential profile exists for
scenarios that satisfy reasonably homogeneous roughness (in the vertical direction,
i.e. dA / dz 0 ) and inertia, rather than pressure driven flow ( d / dz " gS , where
= fluid stress, = fluid density, g = gravity acceleration, and S = channel bed
slope). Noting that such an exponential profile had been previously used in canopy
flow aerodynamics (e.g. Raupach and Thom, 1981), Nikora et al. (2004) suggest that
it may apply to the core of the interfacial region of low-slope flow over well-
submerged vegetation such as reed beds, where near-bed-surface and free-stream
effects are negligible. Conditions for a constant velocity profile to exist were listed
as vertically homogeneous roughness ( dA / dz = 0 ) and the prevalence of wake
turbulence ( d / dz 0 ). The suggested practical example for this case was partially
submerged vegetation in streams or an inundated floodplain.

PhD Thesis 51 Chapter 2.0


L J Campbell Literature Review
The linear profile is postulated (Nikora et al., 2004) as applicable to a range of
roughness types, primarily sediment beds, and in principal for all flow types I-IV.
Nikora et al. (2001) conclude that initial measurements support the existence of such
a linear double-averaged velocity distribution in the interfacial sub-layer, but state
that unfortunately, direct velocity measurements in the roughness layer are rare and,
moreover, mainly relate to u . However, in addition to the laboratory evidence
presented by Nikora et al. (2001, 2004) (see example shown later in Chapter 5,
Figure 5.12), there is some preliminary support for a linear velocity distribution over
certain square bar and cube roughness arrangements. Although they do not
specifically discuss the result, Cui et al. (2003) present a spatial average of the data
shown in Figure 2.15. From observation, the double-averaged streamwise velocity
profile (Figure 2.16) appears to be reasonably linear up to the level of the bar crests.
Similarly, and again by inspection rather than sound analysis, there is evidence for a
linear profile in the interfacial region of flows over cube arrays in the work of
MacDonald (2000) and Kanda et al. (2004).

Figure 2.16: The double-averaged streamwise velocity profile from the k-type square bar flow
data shown in Figure 2.15 (from Cui et al., 2003). U, V, and W are the double-averaged
streamwise, transverse and bed-normal velocities respectively. Note the apparent linear profile
for U in the near-bed flow.

Despite this initial support from existing laboratory (and simulated) data, a proper
assessment of the validity of the models proposed by Nikora et al. (2004) requires
high-quality experimental measurements to properly quantify the local variation in

PhD Thesis 52 Chapter 2.0


L J Campbell Literature Review
the time-averaged flow in the roughness layer. Hence the application of PIV to flows
over simple rough beds in this study.

2.4.4 Distribution of Double-Averaged Fluid Stress

Following temporal averaging, there are two fluid stress terms in the RANS
equations, namely the viscous, and Reynolds or turbulent stresses (see Chapter 3).
Whilst the former is usually considered to be negligible in rough turbulent flows, the
turbulent stress acts to balance the downstream component of the gravity force in 2-
dimensional uniform open-channel flows. Indeed, a common method of evaluating
friction velocity and hence bed shear stress from experimental data is to extrapolate
the turbulent stress profile to mean bed level (e.g. Nezu and Nakagawa, 1993).

Grass (1971) reported that the Reynolds stress increased linearly from the free
surface to within a very short distance of the roughness tops (closely packed flat beds
of either 2mm sand or 9mm pebbles), from which point it steadily decreased towards
the bottom of the roughness. However, other researchers have been puzzled by an
apparent decrease in the turbulent stress at higher elevations than the results of Grass
(1971) would suggest. For example, Raupach et al. (1980) measured this decrease to
begin at distances up to around 10mm above the tops of their roughness element
arrangements (comprising 6mm high cylinders). Antonia and Luxton (1971) and
Mulhearn (1978) report similar findings over strip roughness, whilst Mulhearn and
Finnigan (1978) concluded that the region of decreased shear stress was also a region
in which there were significant horizontal variations in mean velocity and shear
stress. The last sentence contains a clue as to the probable underlying reason for the
apparent discrepancy between the results of Grass (1971) and e.g. Mulhearn (1978).
Furthermore, the answer lies in the application of double-averaging methodology.

As noted earlier, spatially-averaging the terms in the RANS equations adds an extra
term to the fluid stress balance, the form-induced stress. Form-induced stress
becomes important in the vicinity of rough beds where individual roughness
elements distort the time-averaged flow (relative to average flow at the same level).
This is the very region in which Mulhearn and Finnigan (1978) and others report
decreased turbulent stresses. Put simply, in the near-bed region, more specifically in

PhD Thesis 53 Chapter 2.0


L J Campbell Literature Review
the form-induced sub-layer (Figure 2.14), the form-induced stress may compensate
for the observed drop in Reynolds stress, thus maintaining the required balance
against the gravity force down to the roughness tops (where form and viscous drag
terms begin to contribute) in steady, uniform 2-dimensional open-channel flows.

The nature of the form-induced stress distribution, and whether or not it makes a
significant contribution to the total fluid stress within the roughness layer is still
unclear in many respects. For example, when modelling flow over a 3-dimensional
building array (cubes) Lien and Yee (2004) concluded that form-induced stress is
negligibly small compared to the turbulent stress above the canopy (z/k > 1), but is of
the same order of magnitude within the interfacial layer (Figure 2.14). Subsequently
however, they choose to neglect form-induced stress since reference data do not
exist at this time to guide its modelling (Lien and Yee, 2005). In contrast, Kanda et
al. (2004) modelled flow over simple cube arrays and found that the form-induced
could not be neglected either in or above the canopy, where it still contributed
several percent of the total stress.

Figure 2.17: The results for momentum flux of


Bohm et al. (2000) from flow over a model
vegetation canopy (the Black Forest model).
The solid black line is the turbulent momentum
z/k

flux (spatially averaged), and the black line


with circles shows the form-induced (or
dispersive) momentum flux distribution. Note
that both have similar magnitude throughout
the lower half of the canopy (i.e. in the
interfacial layer for z/k<0.5). momentum fluxes (m2/s2)

Results from experimental investigations permitting evaluation of the form-induced


stress also exhibit disagreement. Raupach et al. (1986) report negligible form-
induced stress at the top of their model plant canopy, but note that the results cannot
be extended down into the canopy due to measurement difficulties. Similarly, the
field measurements of Rotach (1993) in an urban roughness layer show that the
dispersive stress is an order of magnitude smaller than the turbulent stress. They also
note that obtaining velocity data from a sufficiently large number of horizontal

PhD Thesis 54 Chapter 2.0


L J Campbell Literature Review
positions in order to reliably calculate spatial averages is very difficult in practice.
More recently, the results of Bohm et al. (2000) and Cheng and Castro (2002)
support earlier studies by finding negligible form-induced stress for z>k, but Bohm et
al. (2000) show that for z < 0.5k form-induced stress is of the same order of
magnitude as the turbulent stress (see Figure 2.17). However, both allude to
measurement difficulties, even in wind tunnels where constructing a true spatial
average is often compromised by probe and traverse gear arrangements (Bohm et
al., 2000) and proper assessment remains an open question and requires very
extensive and technically difficult measurements (Cheng and Castro, 2002).
Velocity data from flows over the model canopy (cylinders) of Poggi et al. (2004c)
revealed that for sparse canopies the form-induced stress can be up to 35% of the
total stress for z/k < 0.5, whereas for dense canopies it is negligible.

As a closing note to this sub-section, all the results cited above pertaining to form-
induced stress levels come from studies of flow over 3-dimensional roughness
arrangements (largely cubes and cylinders). It is entirely reasonable to expect greater
disturbance over 2-dimensional strip roughness, as the flow is restricted to moving
up and over obstacles, having no option to go around the sides. This may be
especially true when the strip roughness has a sharp-edged cross section with vertical
faces to deflect the oncoming stream, i.e. square bars.

2.5 Chapter Closure

This study is concerned with assessing recent developments in double-averaging


methodology (in terms of its application to flow in open-channels) to flows over a
very well documented roughness type, transverse square bars. Therefore the aim of
this chapter has been to bring together and review the existing literature from two
distinct areas, i.e. that concerned with flows over square bars (Section 2.3) and that
with double-averaging methods (Section 2.4), especially as applied to open-channel
flows.

The most pertinent features of existing square bar roughness studies were recorded in
Table 2.1. This table served to demonstrate a number of shortcomings in the
available literature of square bar flows. For example, the number of closed-channel

PhD Thesis 55 Chapter 2.0


L J Campbell Literature Review
flows and wind tunnel studies greatly outnumber those conducted in open channels.
Closely related to this observation is the need for experimental data from open-
channel flows with low relative submergence. The issue of a representative velocity
distribution in shallow open-channel flows over rough beds (or in the near-bed
region of deeper flows) remains unclear, yet it is a flow type frequently encountered
in practice, e.g. in natural gravel bed rivers and streams. It was also argued that there
are potential gaps in the range of roughness spacings previously tested, especially for
relatively dense arrangements (L < 10) where a small change in roughness pitch may
drastically alter the near-bed flow patterns.

Perhaps most importantly, at present there is a lack of synchronous and non-intrusive


velocity data with adequate temporal and spatial resolution in the roughness layer
(and above) suitable for assessing individual terms in the double-averaged
momentum equations. In this study the matching of PIV data to double-averaging
methodology therefore represents a significant forward step. Furthermore, the
resultant spatial maps of time-averaged velocity are central to discovering which, if
any, of the three models for the double-averaged streamwise velocity distribution as
proposed by Nikora et al. (2004) apply to flows over square bars. They are also
necessary for quantifying the primary form-induced momentum flux. Both
distributions of the double-averaged streamwise velocity and the form-induced stress
are presented in Chapter 5, Double-averaged flow characteristics in the roughness
layer. Before this however, there are two methods chapters. The first one, Chapter 3,
follows immediately and presents the derivation of the double-averaged momentum
and continuity equations, whilst the second one describes experimental techniques.

PhD Thesis 56 Chapter 2.0


L J Campbell Literature Review
3.0 ANALYTICAL
METHODOLOGY: THE
DOUBLE-AVERAGED FLOW
EQUATIONS

3.1 Introduction

As outlined in Section 2.4, double-averaging refers to the process by which the


governing equations of fluid flow are averaged in both temporal and spatial domains.
The spatial-averaging step can be either implemented as spatial-area averaging (e.g.
over an area in the x-y plane) or as spatial-volume averaging. The equations
pertaining to the latter method are described and derived here, with the observation
that the two methods are essentially equivalent (i.e. if spatial-volume averaging is
applied to an extensive, infinitesimally thin horizontal slab it collapses to the spatial-
area averaged result). However, volume averaging carries distinct advantages over
areal averaging. For example, (1) volume averaging better suits real measurements
which are always made within a finite volume, not area, (2) volume averaged
equations are equally applicable to surface and subsurface flows, and permit variable
sizes of averaging region depending on the heterogeneity of the studied flow region,
and (3) the development of the volume averaged equations is more straightforward.

The following sections introduce the essentials of volume averaging, with the use of
simple examples where appropriate. The Navier-Stokes equations provide the logical
starting point from which to proceed, via traditional time-averaging, to the full
volume-averaged momentum equations. The contents of this chapter are therefore,

Section 3.2: Introducing the basic momentum equations, The Navier-Stokes


Equations.

Section 3.3: The first step towards double-averaging, i.e. averaging the
momentum equations in the temporal domain, or The Reynolds-Averaged
Navier-Stokes Equations (RANS).

PhD Thesis 57 Chapter 3.0


L J Campbell Analytical Methodology: The Double-Averaged Equations
Section 3.4: Implementing the second averaging step in the spatial domain,
or The Volume-Averaged RANS Equations.

Section 3.5: Clarifying fluid stress on the basis of double-averaging


methodology: Fluid Stress in Double-Averaged Terms.

Section 3.6: The breakdown of the double-averaged equations subject to


simple flow conditions Simplified Equations for Two-Dimensional Flows.

Recall that the following right-hand coordinate system is used throughout this thesis:

x: streamwise
y: transverse (towards left bank)
z: bed-normal

3.2 The Navier-Stokes Equations

The Navier-Stokes (N-S) equations form the starting point for describing any
Newtonian fluid flow. Derivation of the N-S equations is relatively straightforward,
beginning with Newtons Second Law of Motion, i.e. force = mass acceleration.
When applied to a small control volume this gives the following linear momentum
equation (using tensor notation with i, j = 1,2,3. Einstein summation convention
applies throughout this chapter),

ui u
+uj i = g i + . ij (3.1)
t x j

where: = fluid density,


ui = ith component of velocity vector,
gi = ith component of gravitational acceleration, and
ij = stress tensor

PhD Thesis 58 Chapter 3.0


L J Campbell Analytical Methodology: The Double-Averaged Equations
Taken in isolation, Equation (3.1) is not readily applicable to real flow situations as
there are more unknowns than there are equations to solve. To overcome this
limitation it is necessary to express the fluid stresses ij in terms of the velocity
vector. Both Navier and Stokes achieved this during the first half of the 19th century.
Consequently the fluid stresses and strains were related by assuming some viscous
deformation-rate law for a Newtonian fluid. The result was Equation (3.2), the
Navier-Stokes equations (shown for incompressible flow only),

ui ui 1 p 2ui
+uj = gi + (3.2)
t x j xi x j x j

where p = local fluid pressure, and


= fluid kinematic viscosity (= /)

3.3 The Reynolds-Averaged Navier-Stokes Equations (RANS)

Traditionally, a turbulent velocity signal is interpreted mathematically by


substituting a temporal mean and fluctuating velocity component into the momentum
conservation (N-S) equation governing fluid flow. Subsequent averaging yields the
RANS equations [note that the required Reynolds averaging rules are included at the
beginning of Appendix B, Equation (B.1)].

Temporal averaging of the flow variables is commonly used to solve Equation (3.2)
for turbulent flows characterised by random variations of velocities and pressure. To
do so, instantaneous fluctuations are superimposed on to a mean value as illustrated
in Figure 3.1. For example, for the streamwise velocity, u, at any time, t, the velocity
is the sum of the mean time-averaged velocity, u , and an instantaneous fluctuation,
u . Assigning the average velocity in this manner clearly imposes the fundamental
requirement that the averaging result (e.g. u ) is independent of the length of
averaging time. In other words, the flow must be steady-on-average. If this
requirement is satisfied, Equation (3.3) below is valid and u = 0 over the averaging
time t ,

PhD Thesis 59 Chapter 3.0


L J Campbell Analytical Methodology: The Double-Averaged Equations
1 t + t
u ( x, y, z , t ) = u ( x, y, z ) + u ( x, y, z , t ) ;
t t
u= u dt (3.3)

u(t)
u

u (t )
u = u + u

time, t

Figure 3.1: A simplified fluctuating velocity signal showing temporal mean and fluctuating parts

Substituting the mean and fluctuating parts of velocity and pressure into Equation
(3.2) and then averaging produces the Reynolds-Averaged Navier-Stokes equations
(RANS),

u i u i 1 p 2ui uiu j
+u j = gi + (3.4)
t x j xi x j x j x j

From Equations (3.2) and (3.4) it is clear that most terms in the N-S equations are
simply replaced by their time-averaged counterparts in the RANS equations. The
mean momentum equation is however complicated by the last term in Equation (3.4)
which appears only after temporal averaging; this is the Reynolds (or apparent) stress
term. The Reynolds stresses (equal to u iu j ) arise as a by-product of the averaging

operation due to the non-linearity of the convective acceleration term [ u j (u i / x j )

in Equation (3.2)]. In a boundary layer (and considering the coordinate directions in


Figure 2.1), the dominant turbulent shear stress term is u w .

As will be seen below there is a volume averaging equivalent, namely form-induced


stress, to complement the apparent stress term in the time-averaged equations. The
definition of fluid stress was raised when discussing existing spatial averaging

PhD Thesis 60 Chapter 3.0


L J Campbell Analytical Methodology: The Double-Averaged Equations
literature, but it is an important point to reiterate. Within the rigorous theoretical
framework of double averaging, the conventional view of fluid stress as comprising
viscous and turbulent shearing components is (correctly) supplemented by form-
induced stress to account for spatial heterogeneity in the time-averaged flow near a
rough bed.

3.4 The Spatially-Averaged RANS Equations

For the present study of uniform open-channel flows, temporal averaging of the
momentum equations is supplemented by spatial volume averaging at all elevations,
z, implemented over a thin slab oriented parallel to the mean channel bed (the
thickness of the slab being dictated by the PIV measurement control volumes, often
known as interrogation areas, as described in the following experimental methods
chapter). As PIV gives access to u and w velocity components only (and it is a fair to
assume that these are the dominant components for flows over strip roughness), the
double-averaging process simplifies further to computing the average along the
direction of the x-axis (see Section 4.3.4)

For the same reason that flow must be steady (or steady-on-average) for temporal
averaging, spatial averaging requires that the bed roughness is statistically
homogeneous, and that the flow is globally uniform. The procedure for volume
averaging at level z is defined as,

1
( x, y, z ) = dx dy dz (3.5)
Vf Vo

In Equation (3.5), is a time- (or ensemble-) averaged flow variable (e.g. velocity or
pressure) defined in the fluid but not at points occupied by the roughness elements
(the angle brackets denote volume averaging). Vf denotes the volume occupied by
fluid within the fixed averaging region centred at (x,y,z) with total volume Vo.

PhD Thesis 61 Chapter 3.0


L J Campbell Analytical Methodology: The Double-Averaged Equations
The average, , in Equation (3.5) is calculated by considering only the fluid parts

(Vf) of the total averaging domain (Vo), and is termed the intrinsic average. The
simpler superficial average, , has often been used (e.g. Whitaker, 1973; Wang
t

& Tackle, 1994), but is inherently incorrect as superficial averages are computed
from the whole averaging domain (Vo) even when it contains non-fluid portions, for
example sediment grains. The two averages are related to each other by
t = , where = V f / Vo . This ( ) may be interpreted simply as porosity, and

is the volume averaging analogy to A, the roughness geometry function for areal
averaging (as defined by Nikora et al., 2001). Only the intrinsic average is
considered when averaging flow variables in this study.

Just as the instantaneous velocity ( u i ) was split into a temporal mean ( u i ) and

fluctuating components ( u i , Figure 3.1), the mean velocity ( u i ) can be broken into a

volume average ( u i ) and spatial disturbance ( u~i , Figure 3.2), to develop the
spatially-averaged Navier-Stokes equations.

When the volume-averaging operator defined in Equation (3.5) is applied to the


RANS equations (3.4), the result obtained varies depending on whether the
averaging plane or volume intersects the roughness elements or not. Above the
roughness crests the averaging region is simply connected and completely occupied
by fluid ( =1), but below them it is intersected by roughness elements and may
become multiply connected. Whilst the averaging procedure for the flow region over
roughness does not pose any difficulties [indeed, it simply follows the Reynolds
procedure in the same way as the RANS equations are derived from the N-S
equations (e.g. Finnigan, 2000)], the operation of averaging for the flow region
below roughness tops is not trivial because operator (3.5) does not commute with
spatial differentiation for variables that are not constant at the water-bed interface
(refer to the surface integral terms presented later, and in Appendix B). Section 3.4.2
discusses this averaging difficulty with the use of a simple example, before Section
3.4.3 gives the full mathematical derivation for flow below the roughness crests.
Meanwhile the following Section 3.4.1 introduces the volume-averaged equations
for the simple case, i.e. above the roughness crests.

PhD Thesis 62 Chapter 3.0


L J Campbell Analytical Methodology: The Double-Averaged Equations
3.4.1 Flow above the roughness crests (z > zc)

Figure 3.2 shows the decomposition of the temporally-averaged velocity into a


spatial mean and a spatial disturbance about the mean (e.g. from averaging over a
thin bed-parallel volume),

u ( x, y, z ) = u ( z ) + u~ ( x, y, z ) (3.6)

where the straight overbar and angle brackets denote the time and volume average of
flow variables respectively. The wavy overbar denotes the difference between the
time averaged and double averaged values.

u% ( x )
u (x)

u = u + u~

streamwise distance, x

Figure 3.2: A simplified fluctuating mean velocity signal showing double-averaged velocity
and the spatial disturbance

Substitution of this relationship (3.6) into the RANS equation (3.4) followed by
averaging yields the full double-averaged equations for flow above the roughness
crests where porosity, =1:

u i ui 1 p 2 ui u iu j u~i u~ j
+ uj = gi + (3.7)
t x j xi x j x j x j x j

The relationship between the double-averaged equation (3.7) and the RANS equation
(3.4) is analogous to the relationship between the RANS equation and the Navier-
Stokes equation (3.2). Therefore, just as temporal averaging produced the extra
Reynolds stress term, volume averaging has produced a new term, the form-induced

PhD Thesis 63 Chapter 3.0


L J Campbell Analytical Methodology: The Double-Averaged Equations
flux [the last component of Equation (3.7)]. This flux and the associated form-
induced stress ( u~i u~ j ), which is non-zero in the roughness layer, are linked to

discrepancies between time-averaged and double-averaged velocities. Gimenez-


Curto and Corniero Lera (1996) presented a mathematical analysis to show that a
necessary condition for this stress to be non-zero is the existence of vorticity in the
disturbed flow. PIV data acquisition permits accurate measurement of these stresses;
results are presented in Chapters 5 to 7.

3.4.2 Flow below the roughness crests (z < zc)

Below the crests of individual roughness elements the volume-averaging procedure


is complicated as the volume averaging operator (3.5) does not commute with the
differential and Laplacian operators for pressure and velocity, respectively. In other
words the Reynolds averaging conditions (Monin and Yaglom, 1977; Raupach and
Shaw, 1982) are violated when the averaging volume is interrupted by solids. For
example, this means that for the pressure term, the average of the differential is not
equal to the differential of the average when the ratio Vf /Vo (porosity, ) drops
below 1,

~
p ~
p
for z < zc, or Vf /Vo < 1 (3.8)
x x

As a general rule, if the flow variable in question can be considered constant around
the boundary where the averaging plane or volume intersects the roughness element,
the averaging and differential operators will commute. Thus, expressions involving
first order spatial differentiation of velocities, u, v, w, and their fluctuations, u , v , w
do not produce extra terms when double averaged on a plane located below the
roughness crests. However the viscous and pressure terms in the RANS equations do
not satisfy this general rule, hence one may derive extra components from double-
averaging these terms to produce the double-averaged momentum equation.
Although a full derivation is given for the volume-averaged equation in the next
section (3.4.3), the result is replicated below for convenience,

PhD Thesis 64 Chapter 3.0


L J Campbell Analytical Methodology: The Double-Averaged Equations
ui ui 1 p 2 ui 1 uiu j
+ uj = gi +
t x j xi x j x j x j

1 u%i u% j 1 p% 2u%i
+ (3.9)
x j xi x j x j

This is the volume-averaged expression for flow below the roughness peaks,
including porosity, (z) = Vf /Vo, to account for spatial variation in the roughness
geometry. The penultimate and final terms in Equation (3.9) are the extra terms
arising from the non-commutativity of the averaging and differential/Laplacian
operators, as described above; these terms represent components of form drag and
viscous drag respectively.

To clarify the origins of the form drag (and viscous drag) terms, it is helpful to
consider a simple scenario in conjunction with the general rule for flow variables at
the water-bed surface interface as described above. This example has been adapted
from Raupach and Shaw (1982).

3.4.2.1 Pressure Field about a Series of Fences: Introducing the Form Drag Term

Consider a series of impermeable, infinitely long fences lying at right angles to the
flow (Figure 3.3). It is reasonable to assume the time-averaged 2-dimensional
pressure field is as shown, with a pressure differential existing across each fence
because of the occurrence of form drag. Thus, in the space between fences
p / x >0. Furthermore, the spatial average, p , taken at height z < zc takes a
constant value, thus ~
p / x >0 in the inter-fence (interfacial) region (as
~
p = p p ). However, by definition ~
p / x = 0 . Clearly therefore the spatial

differentiation and averaging operators do not commute for pressure when the
averaging region is intersected by roughness elements (fences in this case);
~
p / x ~
p / x . This explains the appearance of the form drag term in Equation

PhD Thesis 65 Chapter 3.0


L J Campbell Analytical Methodology: The Double-Averaged Equations
(3.9) it is necessary to fully account for pressure variations around individual
roughness elements.

x
Flow
z Height at which p
zc and p are
considered

Figure 3.3: Schematic pressure field about a series of impermeable fences lying across the flow
(adapted from Raupach and Shaw, 1982)

Similar reasoning (albeit with a different example) explains why the viscous drag
term arises in Equation (3.9) when the flow is averaged below the highest roughness
crest, for example,

2ui ui
(3.10)
x j x j x j x j

In other words the Laplacian and horizontal averaging operators do not commute for
velocity.

The above example provides a useful insight as to how the spatial averaging operator
works below the roughness tops. The following section derives the double averaged
equations more formally.

PhD Thesis 66 Chapter 3.0


L J Campbell Analytical Methodology: The Double-Averaged Equations
3.4.3 Derivation of the Volume-Averaged RANS Equations (for all z-levels)

Above the roughness tops the averaging domain neither changes with space nor with
time, so both temporal and spatial averaging operators (Equations 3.3 and 3.5
respectively) satisfy all commutation properties. Below the roughness top, the spatial
averaging operator does not commute with the spatial derivatives, so local averaging
theorems must be invoked (e.g. Whitaker, 1967; Gray, 1977).

There are 2 necessary integral theorems for full derivation of the volume-averaged
equations. These are:

Theorem A: The Spatial Averaging Theorem

1 1
xi
=
xi

Vf n dS
Sint
i (3.11)

Theorem B: The General Transport Equation

1 1
t
=
t
+
Vf v n dS
Sint
i i (3.12)

Where is some tensorial quantity, vector or scalar, defined only in fluid (note also
that dV must be additive over the averaging volume); Sint is the solid-fluid
interface surface; vi is the ith component of the velocity vector of the solid-fluid
interface; ni is the ith component of the unit normal vector (directed from the solid
into the fluid); as before angle brackets denote volume averaging, and i = 1,2,3,
where x1,x2,x3 are the coordinate axes of the Cartesian coordinate system.

The first theorem (A), the spatial averaging theorem (Whitaker, 1967; Slattery,
1967), relates to obtaining the volume average of a quantity that varies in space

PhD Thesis 67 Chapter 3.0


L J Campbell Analytical Methodology: The Double-Averaged Equations
whilst the second theorem (B), the general transport equation, is used to correctly
assess the volume average of a temporal derivative. Theorem B will not be derived
here as it is based on the well-known Reynolds Transport Theorem and is only
applicable to one term (the local acceleration term, u i / t ) when double averaging
the RANS equations. Theorem A however is used to average all other (spatial)
derivative terms in the RANS equations; hence a proof, drawing on Gauss Theorem,
is given in Appendix A.

Using the timeaveraged momentum (RANS) equations as a starting point it is


straightforward to apply, as appropriate, either Theorem A or B term-by-term to
obtain the full double-averaged momentum equations. This working is presented in
Appendix B (Part 3). Collating all double-averaged terms gives the full volume (or
spatially) averaged momentum equation for flow over a rigid, no-slip and non-
porous bed as,

u i ui 1 p
+ u j = gi
t x j xi

fluid shear
{ 1 uiu j 1 u%i u% j 1 u (z > zc)
above
crests

terms + i
x j x j x j x j

{
(z < zc)

surface ui
below
crests

1 1
drag terms +
V f pni dS n j dS
Sint Vf Sint x j

(3.13)

In addition to deriving all relevant momentum balance components, Appendix B


(Part 4) contains a further derivation illustrating how the terms in the equations
above (3.13) are precisely consistent with those given in Equation (3.9). Indeed, the
discrepancy in the appearance of both sets of volume-averaged equations (3.9 and
3.13) only arises because the author feels it is more helpful to see form drag and

PhD Thesis 68 Chapter 3.0


L J Campbell Analytical Methodology: The Double-Averaged Equations
viscous drag terms left as surface integrals. The explicit inclusion of these drag terms
(or more specifically the total time averaged drag exerted by the solid on the fluid
per unit mass of fluid) is after all one of the key benefits gained by double averaging
the momentum equations. For instance, the form drag term may be given as:

1 ~
p
as in Equation (3.9),
xi
or as,
1
V f pni dS in Equation (3.13)
Sint

Although the above expressions both relate to form drag (indeed the integral term is
precisely equal to form drag, whereas the derivative term is one component of form
drag, see Appendix B) one can intuitively assess the physical meaning of the latter
term as the sum of time-averaged fluid pressure over the surface interface between
fluid and solid boundaries (in the ith direction, for example in the x-direction when
considering the streamwise component of velocity). Such direct physical
interpretation of the spatial derivative term as it appears in Equation (3.9) is far less
obvious. A similar argument applies for the viscous drag term, i.e. it is far more
meaningful in surface integral form as presented in Equation (3.13). (note also that
although the surface integral version of viscous drag is legitimate, the term given in
Equation (3.9), ( 2 u~i / x j x j ) , is only one component of skin friction, just as

p% / xi is only one component of form drag refer to Appendix B, Equations


(B.30) and (B.40) for further details).

In simple terms, and recalling the origins of the momentum equation as Newtons
Second Law of Motion, Equation (3.13) consists of fluid acceleration terms on the
left, balanced by force-per-unit-mass terms on the right. Key components of this
double averaged momentum balance are (Nikora, 2004):

u i
LOCAL FLUID ACCELERATION
t

PhD Thesis 69 Chapter 3.0


L J Campbell Analytical Methodology: The Double-Averaged Equations
u i
u j CONVECTIVE FLUID ACCELERATION
x j

gi GRAVITY ACCELERATION

1 p
PRESSURE TERM
xi

1 uiu j
TURBULENT (REYNOLDS) STRESS TERM
x j

1 u%i u% j
FORM-INDUCED STRESS TERM
x j

1 u
i VISCOUS STRESS TERM
x j x j

1
V f pni dS FORM DRAG (per unit mass of fluid)
Sint

1 u i
V f x n j dS

VISCOUS DRAG (per unit mass of fluid)
Sint j

For the flow region above roughness crests the last two terms, form and viscous
drag, necessarily disappear, and porosity, , is constant and equal to 1. In this flow
region therefore Equation (3.13) becomes identical to Equation (3.7) given in
Section 3.4.1.

3.4.4 The Double-Averaged Continuity Equation

In addition to the double-averaging the momentum conservation equation, the mass


conservation, or continuity equation can be treated in a similar manner. For an
incompressible fluid, application of integral Theorem A to the time-averaged
continuity equation yields (Appendix B, Part 2):

PhD Thesis 70 Chapter 3.0


L J Campbell Analytical Methodology: The Double-Averaged Equations
ui
=0 (3.14)
xi

3.5 Fluid Stress in Double-Averaged Terms

As noted in the literature review, a full description of the near-bed momentum


balance in rough open-channel flows requires a measure of the spatial variability in
the time-averaged flow. As this spatial heterogeneity can contribute significantly to
momentum transfer in the roughness layer it is not sufficient to consider turbulent
(Reynolds) and viscous shearing alone; instead these should be complemented by the
form-induced stress. In double-averaging terms, fluid stress therefore comprises
three distinct components,
ui
ij = uiuj u%i u% j (3.15)
x j

The first term in Equation (3.15), viscous fluid stress, is normally considered to be
negligible compared to Reynolds stress (the middle term) in rough turbulent flows,
except in the immediate vicinity of the wall where velocity gradients are large. In the
vast majority of cases, form-induced stress (the last term) has been hitherto ignored
in studies of rough open-channel flow, but form-induced stress is not always
negligible. Under uniform, 2-dimensional flow conditions in open-channels the
primary Reynolds and form-induced stress terms (i.e. those involving the
fluctuations of the streamwise, u, and bed-normal, w, velocity components) act to
balance the downstream component of the gravity force. The interplay between these
primary Reynolds and form-induced stresses over square bars will be presented and
discussed in Chapter 5.

3.6 Simplified Equations for Two-Dimensional Flows

The open-channel flows investigated as part of this project were globally uniform
and steady-on-average, and as such were readily spatially-averaged along the x-
direction at all levels, z, through the flow depth. Furthermore they were assumed to

PhD Thesis 71 Chapter 3.0


L J Campbell Analytical Methodology: The Double-Averaged Equations
be 2-dimensional, high Reynolds number flows over very rough (regular) beds. Thus
the following assumptions are justified:

1. There are no long-term trends in flow properties (/t=0, steady flow),

2. All temporal variations are due to turbulence only (flow is steady-on-


average),

3. Longitudinal and transverse gradients (/x and /y respectively) of spatially


averaged quantities are negligible compared to vertical changes (/z), e.g.
there is negligible correlation between velocity components u and v, or v and
w (2-D flow condition),

4. Porosity, =Vf /Vo, depends solely on the vertical coordinate, z (bed


elevations are statistically homogeneous in the x, y plane), and,

5. Viscous shear stress is negligible (high Reynolds number condition).

Furthermore, it follows from continuity that the spatially-averaged transverse and


vertical velocities ( v and w ) are equal to zero.

Upon application of the above assumptions (and with reference to the coordinate
definition sketch, Figure 2.1), the double averaged momentum equations [using the
version given in Equation (3.9)] reduce to:

u w u~w
~
Above crests: z > zc

x-direction g sin =0 (3.16)


z z

~2
1 p w 2 w
z-direction g cos + + + =0 (3.17)
z z z

PhD Thesis 72 Chapter 3.0


L J Campbell Analytical Methodology: The Double-Averaged Equations
1 p% 1 uw 1 uw
% %

Below crests: z < zc


x-direction g sin =0 (3.18)
x z z

1 p 1 p% 1 w2 1 w% 2
z-direction g cos + + + + =0 (3.19)
z z z z

3.7 Chapter Closure

Beginning with the Navier-Stokes equations, this chapter has presented the
derivation of the double-averaged momentum and continuity equations. This has
included the source of the form drag and viscous drag terms that both appear
explicitly upon application of the averaging operator below the level of the
uppermost roughness crests.

A core aim of this investigation was to demonstrate the applicability of the double-
averaged equations [e.g. the primary stress terms in Equation (3.16)] and the
characteristics of the form-induced velocity components with particular reference to
the roughness layer (comprising the interfacial and form-induced sub-layers). The
experimental methodology, described in the following Chapter, permitted excellent
visualisation and quantification of both turbulent and form-induced velocity
components and stresses, in addition to double-averaged velocity profiles in the near-
bed region.

PhD Thesis 73 Chapter 3.0


L J Campbell Analytical Methodology: The Double-Averaged Equations
4.0 EXPERIMENTAL
METHODOLOGY

4.1 Introduction

In order to successfully address the objectives listed in Section 1.1 of this thesis, it
was necessary to conduct a series of laboratory-based flume experiments
investigating a range of rough bed open-channel flows. The basic geometry of
transverse square bars was deliberately chosen for the fixed bed roughness elements.
This decision reflected the need to restrict the initial assessment of the spatially-
averaged momentum equations to flows over simple bed topographies. Fluid motion
over such an uncomplicated channel roughness was not confounded by mobile or
porous bed conditions, or indeed by the considerably more complex bed
arrangements offered by natural gravel sediments.

Assessment of the spatial variability in the time averaged flow over the square bar
roughness configurations required a detailed and synchronous account of how the
flow velocities behaved at a number of different locations above and along the
channel bed. Particle Image Velocimetry (PIV) provides the ideal measurement
technique for recording the necessary spatial description of the flow for double
averaging analysis. Therefore, the primary data-gathering tool used to gather flow
velocities (from a planar area located along the main axis of the flume, i.e. in the
flume midline) was a PIV system. PIV is a non-intrusive technique used to measure
the velocity of micron-sized seeding particles following the flow, and can provide a
near instantaneous spatial description of a cross-section of the flow. It is this spatial
information that is central to examining the interaction of bed roughness and flow in
the near-bed region. For example, PIV vector map data allows visualisation of
certain double-averaged variables introduced in the last chapter, such as the primary
~ , and the form-induced velocity components, u~ and
form-induced stress, u~w
~ . These form-induced variables are key to furthering our understanding of how
w
fluid motion relates to the boundary over which it flows.

PhD Thesis 74 Chapter 4.0


L J Campbell Experimental Methodology
The remainder of this chapter is dedicated to a description of the laboratory
equipment, experimental techniques and analysis methods used to investigate the
nature of temporally and double averaged flow velocities in the vicinity of a rough
bed. Validation of, and further research into the PIV technique was not the purpose
of this project; hence the reader is directed elsewhere for an exhaustive description
of PIV methodology and related issues (e.g. Raffel, 1998). The first of the following
sections does however present a brief account of the underlying principles of PIV via
a discussion of the most popular PIV correlation algorithms. Section 4.3 then
introduces the laboratory hardware utilized, including the hydraulic flume. The
square bar roughness configurations which comprised the channel bed are described
in Section 4.4, and the hydraulic conditions for all experimental runs are given in
Section 4.5. Hence, the contents of this chapter are organised as follows:

Section 4.2: An overview of the velocity measurement technique,


Principles of PIV Measurement.

Section 4.3: Details of the hydraulic flume and other hardware, or


Experimental Apparatus.

Section 4.4: The square bar roughness configurations: Bed Roughness.

Section 4.5: Ranges of key hydraulic parameters, Hydraulic Conditions.

Section 4.6: Bringing together a range of experimental, roughness and


hydraulic parameters in the form of a Table of Experiments.

4.2 Principles of PIV Measurement

As mentioned above, PIV is a non-intrusive technique which measures the velocity


of micron-sized seeding particles following the flow. It relies on recording digital
images of stroboscopically illuminated seeded flow for subsequent correlation
analysis. The output is a series of velocity vector maps produced by calculating
seeding particle separation in a known time interval. The time between illuminations

PhD Thesis 75 Chapter 4.0


L J Campbell Experimental Methodology
is governed by the time separation of the illumination pulses (t), whilst the particle
separation (s) is determined by applying a correlation function to small regions of
the imaged flow. These small areas are called interrogation areas, and are usually 32
32 or 64 64 pixels in size. This study used 32 32 pixel interrogation regions
which typically corresponded to a physical flow area of 1.6 1.6mm.

sx

x
Figure 4.1: Autocorrelation PIV frame depicting fluid ticker-tape trace (each seeding particle is
illuminated up to 15 times within the image exposure time). Flow is from left to right and the full

image represents a flow region of approximately 60 30 mm. White rectangle shows location of

enlarged area illustrating the particle displacement, s travelled during illumination pulse
separation, t.

z Image 1 Image 2

x
Figure 4.2: A small area from a pair of images suitable for cross-correlation PIV analysis. Flow is
from left to right; the square roughness element shown is approximately 6mm high. Each particle is
illuminated only once in each frame, with time separation t between illumination pulses.

PhD Thesis 76 Chapter 4.0


L J Campbell Experimental Methodology
There are two commonly applied correlation techniques; autocorrelation and cross-
correlation. The former calculates the most likely average displacement within an
interrogation region by correlating pixel greyscale values within one frame (ideally
over a large number of seeding particles), whilst the latter correlates particle
illuminations from two successive frames (Figures 4.1 and 4.2 show sections of
typical frames suitable for PIV autocorrelation and cross-correlation algorithms,
respectively). The relative merits of each method are outlined below by means of a
simple example.

4.2.1 PIV Correlation Algorithms by Example

To highlight how both the autocorrelation function (ACF) and the cross-correlation
function (CCF) work, consider a small planar area of fluid flow which on the
recorded digital image occupies a region of 8 8 pixels. For the purposes of this
example, this area is interpreted as a PIV interrogation region. Within this area the
flow velocity is such that seeding particles move 2 pixels in both the x and z
directions before subsequent illumination. Clearly in the case of autocorrelation
subsequent illuminations are contained within the same image, however with cross-
correlation the second illumination is captured in a second digital image. Figures 4.3
and 4.4 depict this idealised case.

In reality the greyscale value of each pixel can vary between 0 and 255 if recorded
using an 8-bit digital camera, however here the illuminated particles (shown as grey
squares in Figures 4.3 and 4.4 and idealised as covering only one pixel) are
arbitrarily assigned a value of 150; white squares where no particles have been
detected are given a greyscale value of 0. Once the digital images have been split
into suitable interrogation regions, PIV software is used to implement either the
auto- or cross-correlation function. The general form of these functions is,

itot (G ( x, z ) G ( x + x, z + z ))
ACF = 2
(4.1)
i =1 G av

PhD Thesis 77 Chapter 4.0


L J Campbell Experimental Methodology
and,

itot [G ( x, z )] [G ( x + x, z + z )]2
CCF = 1
2
(4.2)
i =1 G av

where G(x, z) refers to the 8-bit greyscale rating at position (x, z) of the ith pixel and
x and z are horizontal and vertical pixel shifts respectively. Gav is the average
greyscale value over the entire interrogation area. In the case of the 8 8 pixel
interrogation region, itot is equal to 64 pixels. The Cross-Correlation Function [CCF,
Equation (4.2)] is identical to the Autocorrelation Function [ACF, Equation (4.1)]
with the exception that it compares the greyscale value of pixel i(x, z) in image 1
with pixel i(x + x, z + z) in image 2 (the subscripts 1 and 2 in the numerator of
Equation (4.2) denote pixels in image 1 and 2 respectively).

i=1 2 3 4 5 6 7 8

9 10 11 12 13 14 15 16

17 18 19 20 21 22 23 24 Flow
direction?
25 26 27 28 29 30 31 32

33 34 35 36 37 38 39 40
= 150
41 42 43 44 45 46 47 48
=0
49 50 51 52 53 54 55 56

57 58 59 60 61 62 63 64
z

Figure 4.3: Example 8 8 pixel AUTOCORRELATION interrogation region showing two


seeding particles passing through. Shaded pixels represent particle positions and are assigned
a greyscale value of 150. All other pixels are given a value of zero. The time between
successive illuminations is t, and by inspection it is evident that the most probable particle
displacement in this time interval is (x, z) = (2 , 2) or(-2 , -2) Note that the autocorrelation
technique therefore suffers from directional ambiguity.

PhD Thesis 78 Chapter 4.0


L J Campbell Experimental Methodology
i=1 2 3 4 5 6 7 8 i=1 2 3 4 5 6 7 8

9 10 11 12 13 14 15 16 9 10 11 12 13 14 15 16
Flow
17 18 19 20 21 22 23 24 direction 17 18 19 20 21 22 23 24

25 26 27 28 29 30 31 32
t 25 26 27 28 29 30 31 32

33 34 35 36 37 38 39 40 33 34 35 36 37 38 39 40

41 42 43 44 45 46 47 48 41 42 43 44 45 46 47 48

49 50 51 52 53 54 55 56 49 50 51 52 53 54 55 56
= 150
z 57 58 59 60 61 62 63 64 57 58 59 60 61 62 63 64
=0

x
Figure 4.4: Example 8 8 pixel CROSS-CORRELATION interrogation region showing
three seeding particles passing through (illumination pulses in images 1 and 2 are separated
by time interval t). Shaded pixels represent particle positions and are assigned a greyscale
value of 150. All other pixels are given a value of zero. As in Figure 4.3, the time between
successive illuminations is t, and by inspection it is evident that the most probable particle
displacement in this time interval is (x, z) = (2 , 2). Note that there is no directional
ambiguity.

The next two subsections calculate the ACF and CCF for the sample case depicted in
Figures 4.3 and 4.4, respectively. This exercise not only aids explanation of the
principles underlying PIV velocity measurement, but it allows an evaluation of the
strengths and weaknesses of each correlation algorithm.

4.2.1.1 Applying the Autocorrelation Function (ACF)

Applying the ACF [Equation (4.1)] to the test case depicted in Figure 4.3 yields the
following:

The denominator of the ACF (the square of the average greyscale value over the
whole interrogation region) is:

Gav2 = (
(5 150) + (59 0)
64
)
2
= 137.33

PhD Thesis 79 Chapter 4.0


L J Campbell Experimental Methodology
The numerator of the ACF is computed for all combinations of pixel shifts, x and
z, for example (x, z) = (2 , 2):

64
(G ( x, z ) G ( x + 2, z + 2)) = . . .
i =1

... {(0 0) + . . . + (150 0) + . . .+ (0 0) + . . .+ (150 0) + . . .


. . . + (0 0) + . . . + (150 150) + . . . + (0 0) + . . . + (150 150) + . . .
. . . + (0 0) + . . .+ (150 150) + . . . + (0 0)} = 67500

Hence, for (x, z) = (2 , 2), the ACF carries the value:

67500
ACF 2, 2 = = 491.5
137.33

Likewise for (x, z) = (0 , 0), or zero displacement:

5(150 150) + 59(0 0)


ACF 0, 0 = = 819.2
137.33

Figure 4.5: Autocorrelation function (ACF) plot,


clearly illustrating the correlation peaks for the
given example. The highest (self-correlation)
ACF value

peak [for (x, z) = (0 , 0)] must be masked to


enable detection of the real displacement
correlation peak. Note also the symmetry of the
ACF map, with displacements detected at both
(x, z) = (2 , 2) and (-2, -2). This gives rise to
directional ambiguity, one of the core difficulties

z attributable with the ACF algorithm.

Repeating this process for (x, z) = (-2 , -2) will also give ACF-2, -2 = 491.5. There
are also a number of secondary correlation peaks associated with, for example, the

PhD Thesis 80 Chapter 4.0


L J Campbell Experimental Methodology
displacements (x, z) = (2 , -4), (-2 , 4), (-4 , 1), (4 , -1), (5 , -1) and (-5 , 1).
However all of these combinations will produce ACF values lower than those
already calculated above as they involve contributions from fewer particle
illuminations. Indeed for this simple case the majority all other combinations of x
and z produce ACF = 0. The results for all x and z can be plotted as a series of
ACF correlation peaks; this correlation map is shown in Figure 4.5 (please note that
this figure shows only the larger correlation peaks for simplicity).

Inspection of Figure 4.5 reveals an undesirable by-product of the autocorrelation


function the highest peak is centred on (x, z) = (0 , 0) where the particle
illuminations are self-correlated. In order to obtain meaningful velocities using this
method it is therefore necessary to mask this central peak (before the PIV software
determines the most likely displacement by searching for the next highest correlation
peak). A related issue with autocorrelation is the inability to resolve zero or near-
zero particle velocities. This arises partly out of a necessity to have sufficient
separation between subsequent particle illuminations in order that they appear as
discrete dots on the digital image, but also because correlation peaks associated with
small displacements tend to become engulfed by the large central correlation peak
and mask.

FLOW FLOW

real fluid motion incorrect autocorrelation representation

Figure 4.6: Sketch illustrating the problem of directional ambiguity associated with autocorrelation
PIV. The eddy located downstream from the square roughness element is incorrectly represented in
the figure on the right, as the software cannot correctly resolve upstream velocities.

Furthermore, the direction of fluid motion cannot be resolved using an ACF


algorithm, as the software cannot determine the temporal order of particle
illuminations. Instead the user must impose a flow direction on the software, with the

PhD Thesis 81 Chapter 4.0


L J Campbell Experimental Methodology
result that areas of recirculation (negative, or upstream fluid movement) are not
correctly detected. The sketch shown in Figure 4.6 highlights how such an area of
flow in the wake of a roughness element would be wrongly interpreted by
autocorrelation PIV.

4.2.1.2 Applying the Cross-Correlation Function (CCF)

Applying the CCF [Equation (4.2)] to the test case depicted in Figure 4.4 yields the
following:

The denominator of the CCF (the square of the average greyscale value over the
whole interrogation region) is:

2
(3 150) + (61 0)
Gav2 = = 49.44
64

As for the ACF, the numerator of the CCF is computed for all combinations of pixel
shifts, x and z. The key difference is that the pixel at position x in the first image
is compared with the pixel at position x + x in the second image. For example (x,
z) = (2 , 2):
64
([G ( x, z )]1 [G ( x + 2, z + 2)]2 ) = . . .
i =1

3 (150 150) + (61 0) = 67500

Hence, for (x, z) = (2 , 2), the CCF carries the value:

67500
CCF 2, 2 = = 1365.3
49.44

Note however that the reverse displacement (x, z) = (-2 , -2) does not generate a
large correlation peak as was produced by the ACF algorithm. Likewise for (x, z)

PhD Thesis 82 Chapter 4.0


L J Campbell Experimental Methodology
= (0 , 0), or zero movement, the CCF method will not give a high correlation peak,
unless of course the fluid is at rest and it represents the true displacement. For the
simple example case the highest correlation peak arises for the pixel shift
combination of (x, z) = (2 , 2); this is shown in Figure 4.7 (please note that only
the highest peak is shown for simplicity) alongside a cross-correlation map obtained
from real data.

Therefore using a CCF as opposed to an ACF PIV algorithm avoids complicating the
search for the true displacement peak, as the search routine is hampered neither by
large central [(x, z) = (0 , 0)], nor opposite direction peaks. Once the correlation
peak is located, PIV software typically uses a three-point estimator1 to compute both
the magnitude and direction of the particle displacement. Equipped with knowledge
of the illumination pulse time separation, t, calculation of the resultant velocity
vector becomes a straightforward application of speed=distance/time.

CCF
z
CCF value

z
x

Figure 4.7: The cross-correlation function (CCF) plot on the left clearly illustrates the correlation peak
for the given example centred at (x, z) = (2 , 2). Note how this peak is not dwarfed by a central peak
at (x, z) = (0 , 0) as with autocorrelation (i.e. compared to Figure 4.5). The image on the right shows
a cross-correlation plot for real data. Although this plot contains many more peaks than for the simple
example, the highest peak associated with the actual particle displacement is easily discernable.

1
If the highest point in the correlation plane is Ri,j, its nearest neighbours are Ri-1,j, Ri,j-1, Ri+1,j, and
Ri,j+1. A curve is fitted to the 3 points in the x-direction and the maximum value of this curve is
assumed to be located at the x-coordinate of the correlation peak. This process is repeated for the z-
coordinate. The function fitted to the correlation values is generally a Gaussian curve (for further
details refer to e.g. Raffel et al., 1998).

PhD Thesis 83 Chapter 4.0


L J Campbell Experimental Methodology
In summary, cross-correlation represents a more flexible alternative to
autocorrelation as there is no ambiguity regarding flow direction and it deals
successfully with issues such as zero or near zero velocity detection by avoiding the
need for a central peak mask. However, autocorrelation does carry one distinct
advantage in that the flow patterns can be clearly visualised. The fluid ticker tape
traces recorded on autocorrelation PIV images (refer back to Figure 4.1) provide a
useful insight into fluid motions in the vicinity of rough beds. Whilst the detailed
quantitative data generated by cross-correlation PIV is vital to understanding the
spatial distribution of mean and turbulent properties of the flow, much can be gained
by simply observing the fluid flow patterns captured on autocorrelation images.

This study made use of both auto- and cross-correlation PIV methods. The former
was put to use for preliminary investigations, whereas the latter method was
implemented to generate the bulk of the quantitative flow velocity data. All of the
results presented in Chapters 5, 6 and 7 were gathered using cross-correlation, as
autocorrelation was unable to adequately resolve the flow structure in the near-bed
region.

4.2.2 PIV Correlation Peak Location: Sub-Pixel Accuracy

Raffel et al. (1998) propose that estimation accuracies for the position of the
correlation peak lie in the range 1/10th to 1/20th of a pixel. Such subpixel accuracy
does however depend upon there being sufficient seeding (particle image density)
recorded in each interrogation region. Put simply, the probability of a valid
displacement detection increases when more particle image pairs enter the
correlation calculation. Keane and Adrian (1992) demonstrated that for cross-
correlation PIV the valid detection probability exceeds 95% when,

Ni Fi Fo > 5.0 (4.3)

where Ni = particle image pair density within the interrogation


region;

PhD Thesis 84 Chapter 4.0


L J Campbell Experimental Methodology
Fi = a factor expressing the in-plane loss of pairs (due to
velocities in the plane of the PIV cross-section
moving the particle out of the interrogation region
before subsequent illumination), and,

Fo = a factor expressing out-of-plane loss of pairs (due to


velocities carrying particles out of the PIV
measurement plane).

From observation of a representative set of images, the PIV images generated as part
of this project easily satisfy the above criterion. This suggests two things, firstly that
the time interval (t) between cross-correlation images was kept sufficiently short to
restrict the average particle displacement to less than the physical size of a PIV
interrogation region (low Fi), and secondly that the effect of cross-channel fluid
motion was negligible (low Fo). Despite this, a small number (up to 1%) of
erroneous vectors were generated in each image. Due to the large number of frames
being analyzed it was necessary to automate a validation procedure. Two methods
were used to search for and replace invalid vectors. The first was based on the work
of Westerweel (1994), in which vectors are classified as erroneous if they varied
significantly from their nearest (spatial) neighbours. In this investigation, eight
neighbouring vectors were examined and vectors varying by more than two standard
deviations from the mean value were defined as erroneous and deleted. The second
was a simple time-series search which detected velocities varying by more than
seven standard deviations from the mean signal (such a wide range was necessary to
account for the highly turbulent flow in the vicinity of the rough bed). In both cases
invalid vectors were replaced by the mean value of a group of neighbouring vectors.

Now that the basics of PIV measurement have been presented, the following section
details the equipment utilized for obtaining and analysing cross-correlation PIV
images, including the hydraulic flume.

PhD Thesis 85 Chapter 4.0


L J Campbell Experimental Methodology
4.3 Experimental Apparatus

During each experimental run around 300 seconds of rough bed open-channel flow
was recorded from the longitudinal midline of the flume. As the chosen bed
roughness comprised 2-dimensional bars and the aspect ratio of the flow (the ratio of
flume width to flow depth) was kept above 4.5 (e.g. Nezu and Nakagawa, 1993), it
was not considered necessary to collect data from other locations across the flume
width (aspect ratios for all experiments are tabulated at the end of this chapter in
Table 4.2). The PIV images (8192 per experiment giving 4096 cross-correlation
time-steps) were acquired straight to PC SCSI hard drives via a Kodak ES1.0 digital
camera, and a dedicated frame-grabbing board. Illumination was provided by a
double-pulsed Nd-Yag laser.

4.3.1 Hydraulic Flume

All experiments were conducted using a recirculating hydraulic flume in the


University of Aberdeen Fluids Laboratory. The channel has a straight, rectangular
section with dimensions 11m long, and 400mm wide. It is possible to vary the bed
slope from level to 1 in 50, although this was fixed at 1/400 for the experiments
reported in this thesis. Downstream control is provided by means of a series of 6
vertical vanes, which can be simultaneously pivoted about their vertical axes to
constrict the flow (Yalin, 1977). Beyond this control the water spills freely into a
large holding tank, where discharge is evaluated by means of a right-angled v-notch
weir. From the downstream holding tank the water is pumped through a 150mm
diameter pipe to the header tank at the upstream end of the flume. The main control
valve operates on this return pipe just beyond the pump. Further details of this
installation can be found in Cunningham (2000).

Uniform flow was obtained in all experiments by careful adjustment of the control
valve and the downstream vertical vanes. For each experiment the procedure was as
follows. Firstly, the bed slope was checked to be 1/400 (this was set prior to the first
experiment and was not subsequently altered between experiments). Secondly, a
small glass sheet (see Section 4.3.2.2) was solidly clamped at the measurement
location at the chosen level of the free surface and parallel to the channel bed. This

PhD Thesis 86 Chapter 4.0


L J Campbell Experimental Methodology
was done dry, i.e. before the pump was switched on, and provided the target for
setting flow depth. The downstream vanes were then checked to be fully open, and
the control valve largely closed before the pump was switched on. Once a small
amount of water was running through the flume the control valve was opened slowly
until the water just reached the underside of the glass sheet. This generated the
required flow depth at the measurement location. However, with the downstream
vanes left completely open the flow was rarely satisfactorily uniform at this stage of
the procedure (in most cases there was a drawdown curve ahead of the overfall into
the holding tank). To achieve uniform flow the downstream vanes and the control
valve were alternately adjusted whilst measuring flow depth along the channel length
(see below). In general, the vertical vanes were closed slightly from the fully open
position to counter the natural drawdown curve at the downstream end of the flume.
Once the position of the vanes had been altered the control valve was appropriately
adjusted to maintain the required flow depth such that the water surface just touched
the underside of the glass sheet. If necessary this iterative process of adjusting the
vanes and control valve was repeated until uniform flow depth was reached.

Verification of uniform flow conditions was initially achieved by the use of a steel
rule marked with 0.5mm increments, and latterly by a Vernier scale clamped to a
wheeled trolley that ran in aluminium channels the full length of the flume. The
Vernier scale was used to record both bed height and water surface height every 0.5
m along the channel mid-line; the difference between the two readings gave flow
depth. Flow was deemed to be uniform when the variation in flow depth along the
flume was 1 mm (noting that most of this variation occurred a short distance
within the upstream and downstream ends of the flume where the flow adjusted to
inlet and outlet conditions respectively).

4.3.2 PIV Equipment

Component parts of the cross-correlation PIV set-up included a digital camera, a


computer equipped with a frame-grabbing board, image acquisition and PIV analysis
software, and an illumination source (the general set-up is shown in Figure 4.8).
These components are now introduced in turn.

PhD Thesis 87 Chapter 4.0


L J Campbell Experimental Methodology
Laser light
sheet coming Signal
from laser generators to
heads situated control system
above flume timing

Cross- High-spec. PC
correlation with frame
digital camera grabbing board,
image
Laser power acquisition and
supply, PIV analysis
cooling and software
control unit

Figure 4.8: Laboratory set-up for the Aberdeen cross-correlation PIV system

4.3.2.1 Digital Camera and PC Components

A dedicated cross-correlation camera suitable for recording two digital frames in


rapid succession was used. The Kodak Megaplus ES1.0 is capable of capturing two
digital images at full 1008 1008 pixel resolution (i.e. approximately one
megapixel) separated by only a few microseconds; this pattern is can be repeated up
to 15 times per second (refer to the system timing diagram presented in Section
4.3.2.5). In real terms the system recorded 26-28 frames per second, which translated
into a 13-14Hz experimental sampling rate (recall each time step requires 2 images
for cross-correlation). Due to the image data download characteristics of the CCD
chip, the exposure times for the first and second images in each image pair varied
significantly. The first exposure (or integration) time could be varied, but the second
was fixed at 32 msec. (again, refer to the system timing diagram in Section 4.3.2.5).
It was therefore vital to eliminate any extraneous light from the recording area to
maintain comparable greyscale intensities across both images in spite of the differing
camera exposure windows. This ensured correct functioning of the cross-correlation
algorithm.

PhD Thesis 88 Chapter 4.0


L J Campbell Experimental Methodology
The computer used to acquire, store and analyse the PIV bitmap images was
necessarily highly equipped. The specification included dual P4 Xeon 1.7GHz
processors, 1GB Rambus RAM, three SCSI hard drives for the direct-to-disk
acquisition of images, and a DVD writer. The inclusion of a DVD writer was
especially valuable as it enabled efficient data storage. In the initial stages of the
experimental program, each experiment generated 8192 8-bit 1008 1008 pixel
images (approximately 8GB of data); there were 52 experiments conducted in total.
Core software applications used were VideoSavant for image acquisition and
VidPIV for PIV correlation analysis.

4.3.2.2 The Illumination Source

Cross-correlation PIV requires a well-controlled and intense light source to


successfully illuminate the seeding particles. Nd-Yag laser light pulses fulfil both
these criteria. A New Wave Research Solo III double-pulsed, frequency doubled Nd-
Yag laser was used for PIV illumination. This is essentially two laser heads each
capable of firing a high energy (50mJ) pulse of coherent green laser light (532nm
wavelength) up to 15 times per second. One laser head was therefore used to fire
during the odd PIV frames, whilst the other discharged in the even frames. After
leaving the laser housing, the 2.5mm diameter beam was spread and focussed into a
light sheet (see Figure 4.9) by a series of spherical and cylindrical lenses.

It was necessary to position a small glass sheet (250 400mm in the x, y plane) on
the water surface during all experiments. This was essential for maintaining the
integrity of the laser light sheet as it passed through the air-water interface.
Otherwise the light sheet was non-uniformly diffracted as free-surface undulations
(small waves) passed through the measurement area, resulting in a vertical banding
effect which is partly evident in Figure 4.9. Whilst the use of a glass lid was
undesirable, it was preferable to losing vertical strips of data in areas where the lack
of illumination meant the camera could not detect sufficient seeding. Furthermore,
the likely hydrodynamic influence of the experimental square bar surfaces, which are
very rough and irregular when compared to smooth glass, dominated any effects
from a short length of free-surface sheet. Preliminary testing confirmed that the

PhD Thesis 89 Chapter 4.0


L J Campbell Experimental Methodology
effect of the glass sheet typically penetrated around 10mm into the flow (from the
water surface). As this study was primarily focussed on the near-bed region (i.e. the
roughness layer), this was deemed to be an acceptable concession.

Figure 4.9: The green


(532nm wavelength) Nd-
Yag laser light sheet
illuminating flow in the
longitudinal midline of the
flume. The digital PIV
camera is especially
sensitive to green light.

4.3.2.3 PIV Seeding

The ideal PIV seeding is neutrally buoyant (i.e. has the density of water), and
suitably small. The latter characteristic will depend on the scale of the flow to be
recorded. Glass microspheres, with a mean diameter of 10m and specific gravity
marginally above 1, was the seeding chosen for this study. Alternative particles
include polystyrene balls or conifer pollen.

4.3.2.4 PIV System Timing

The timing for cross-correlation PIV is critical. The laser light sheet must illuminate
the flow field seeding only once within each camera exposure window. The activity
of the camera, laser and frame grabbing boards must therefore be accurately
controlled throughout the duration of each 300 second experiment. This control was
effected by means of a series of pulses from signal generators (Thurlby Thandar and
BNC digital pulse generators). Figure 4.10 shows a sample timing diagram for
producing 14 image pairs per second, with particle illuminations (i.e. laser light
pulses) separated by 1msec. Implementation of this timing regime enabled a
temporal sampling rate of around 14Hz, with a pulse delay (t) of 1msec.

PhD Thesis 90 Chapter 4.0


L J Campbell Experimental Methodology
4.3.2.5 PIV Processing Parameters

As detailed above, the time interval between PIV illuminations was 1msec. This was
suitable for all experiments as the flow velocities encountered were never fast
enough to carry the seeding particles out of the interrogation regions in this time
period. Interrogation regions for all experiments were 32 32 pixels and were
overlapped by 75% to give a velocity vector every 8 pixels in each direction the
physical flow area corresponding to this size of interrogation region (which is
essentially the PIV sampling area) can be seen in Table 4.2, Section 4.6 for all trials.

~ 72msec (14Hz)
SG1 master
trigger

This master trigger is sent to the frame grabber board 14 times per second,
which in turn controls the digital camera (each trigger signal causes the
camera to grab a pair of images in rapid succession). The SG1Master
trigger also provides the input to SG2 which controls the activity of the two
laser heads.
Camera exposure
windows

1 2 1 2

The camera becomes light-sensitive 20sec after it is triggered. The first


exposure window is kept deliberately short (0.5msec) to avoid problems
with extraneous light entering the image. The second exposure window
begins 0.5sec after the end of the first, and lasts for a fixed 32msec.

t = 1msec
SG2 laser triggers
(driven from SG1
signal)

1 2 1 2

Each laser head fires once every 72msec (driven by the master trigger
from SG1). The pulse width is 9ns. This ensures one near-instantaneous
illumination pulse per camera exposure window. The time delay between
pulses is controlled via SG2, for example here it is shown as 1msec. This
delay represents the time interval, t, which is passed on to the PIV cross-
correlation software to calculate particle velocities.

Figure 4.10: Cross-correlation PIV timing diagram. Signal generator 1 (SG1) = Thurlby
Thandar; Signal generator 2 (SG2) = BNC digital. Timing lines are not drawn to scale.
PhD Thesis 91 Chapter 4.0
L J Campbell Experimental Methodology
The storage and PIV processing time involved for each sequence of bitmaps was
considerable. Fifty two experiments were conducted in total (although a subset of 36
is presented in this thesis); each one took around 20 hours of computing time to
arrive at a series of velocity vector text files suitable for further analysis with Matlab.

4.3.3 Other Parameters

Water temperature in the flume, which has a direct bearing upon fluid viscosity, was
recorded using a standard mercury thermometer at the beginning and end of each
experiment. In order to verify uniform flow conditions were reached and maintained,
the flow depth was verified as constant ( 1mm) along the length of the flume prior
to, and during, every data acquisition run. This depth can also be measured from the
PIV bitmaps.

4.3.4 Averaging Procedure

For each of the 4096 time-steps per experiment, PIV analysis produced an
instantaneous velocity vector map containing 122 velocity vectors in both the
streamwise (x) and bed-normal (z) directions. The first step towards double-
averaging the data was to temporally average the whole time series, before spatially
averaging along lines parallel to the mean bed (i.e. in thin slabs, of height equal to
the dimension of a PIV interrogation area). The cross-correlation camera was always
carefully aligned so that the base of the image was parallel to the slope of the flume
bed, hence the grid of vectors was naturally organised in lines parallel to the solid
bed. This made averaging along lines of constant distance from the mean bed level
relatively straightforward. Figure 4.11 shows this averaging routine schematically.

Double averaging was achieved by implementing a combination of Matlab .m files


and C programme files. Particular care was taken to select suitable image regions
from each experiment, for example to avoid areas of shadow behind square bars at
the periphery of the camera viewing area, and also to accurately define the correct
averaging length in space.

PhD Thesis 92 Chapter 4.0


L J Campbell Experimental Methodology
t =4096
RECALL: ui = ui + ui and ui = ui + u~i

t =2
t =1

Temporal Spatial
averaging averaging

u ( m / s )

ui ( x , z , t ) u i ( x, z ) ui (z )
Instantaneous velocities in Time-averaged velocity in The double-averaged
the (x,z) plane, recorded for the (x,z) plane, averaged velocity profile is calculated
4096 time steps. The over the full 4096 time by averaging the time-
streamwise velocity steps at each vector grid averaged velocity data
component (u) is illustrated. point. along lines parallel to the
mean bed.

Figure 4.11: Schematic representation of the double (temporal followed by spatial) averaging procedure.

PhD Thesis 93 Chapter 4.0


L J Campbell Experimental Methodology
In all cases the spatial-averaging window covered a whole number of roughness
pitches, usually between 1 and 3, and was specified between the centrelines of the
roughness bars (the number of elements averaged over for each experiment is listed
in Table 4.2 at the end of this chapter). Whenever the edges of the averaging area or
the bed topography did not exactly coincide with a velocity vector grid point (which
occurred in most cases), the averaging routine interpolated linearly to the correct
window size and also to zero (no-slip) velocity at all solid surfaces. As the vector
grid spacing was always 8 pixels in each direction, the maximum interpolation
distance was 4 pixels (the corresponding physical distances varied depending on the
size of the viewed area and can be calculated using the information in Table 4.2).

4.4 Bed Roughness

The rough beds under study for this project comprised exclusively square bars (6.35
0.2mm high). As stated in the introduction to this chapter such a basic 2-
dimensional bed geometry was deliberately chosen as it reflected the need to restrict
the initial assessment of the spatially-averaged momentum equations to flows over
simple roughness topographies. The range of literature reviewed in Chapter 2 also
demonstrated that square bars are a well-studied roughness type (e.g. Knight and
MacDonald, 1979a,b) meaning the current work may be readily compared to
previous research.

Consideration was restricted to constant, periodic roughness spacing along the length
of the flume (i.e. square bars placed transversely at a constant centre-to-centre pitch
Figure 4.12 shows the bars in situ for one of these periodic arrangements).
Although this choice of bed topography may seem overly constrained, it enabled a
dedicated and practically exhaustive investigation into the effect that roughness
spacing exerts on the mean and turbulent flow properties in the vicinity of such
rough beds. Of particular interest was the anticipated change in flow structure as the
gaps between roughness elements grew. At close spacings one would expect a
regime traditionally characterised as skimming flow, whilst at less dense roughness
configurations wake interference and ultimately isolated roughness flow would in
theory dominate.

PhD Thesis 94 Chapter 4.0


L J Campbell Experimental Methodology
Figure 4.12: Fixed (6.35mm high) transverse square bar roughness elements within the flume
(L=8 spacing). The wooden square bars, which were first coated with marine varnish to avoid
swelling and distortion under water, were pinned to the marine ply flume bed. The roughness
pattern ran the whole length of the laboratory channel in all cases.

Initial autocorrelation PIV results had indicated that roughness elements spaced at, or
close to, the natural reattachment length of the near-bed flow could markedly alter
the overlying fluid flow; full investigation of this preliminary result required a
comprehensive programme of experiments with systematically varying roughness
element spacing. Indeed, undertaking such a series of focussed tests over a suitable
range of bed configurations provided the ideal conditions for examining the flow
behaviour over both d-type and k-type roughnesses, and perhaps more importantly
the zone of transition between the two.

Table 4.1 lists the key lengths used to define the roughness arrangements for all 12
spacings tested, together with the mean bed levels and the variation in porosity, (z),
introduced in the previous chapter. Note that (z) is step function for all roughness
configurations. The notation allocated to the bed roughness geometry was given in
the literature review (Figure 2.5), but is repeated here as Figure 4.13 for
convenience. The normalised roughness pitch or spacing, L, was originally assigned
the Greek letter lambda, , in recognition that the periodic roughness positions were
effectively a bed topography wavelength. However, is already used extensively

PhD Thesis 95 Chapter 4.0


L J Campbell Experimental Methodology
in the field of hydraulics, most notably as a friction factor in the Colebrook-White
equation. The letter L was therefore substituted for . The roughness pitch is referred
to as P, whilst the roughness height is denoted by k in line with much existing
notation.

p
L=
k
P
lx

z
lz = k
x

Figure 4.13: Bed roughness geometry notation (repeated from Figure 2.5). The bars have constant
square cross section, hence lx = lz. L is the normalised roughness pitch, or roughness spacing
(roughness pitch, P, normalised with the roughness height, k). The roughness height, k, was 6.35
0.2mm; this small variation was due to manufacturing and material tolerances. Bar sizes which fell
outside this range, or significantly warped bars were rejected.

The roughness height, k, was 6.35mm in all cases, therefore a bed assigned
roughness spacing L = 8 had a centre-to-centre roughness pitch of approximately 8k
(P49mm) and an inter-bar gap of a nominal 7k (~43mm). The reason that these
values are approximate and not exactly equal to 7k or 8k is because many of the
aluminium spacers (used to ensure equal spacing of the bars as they were fixed to the
bed of the hydraulic flume) were machined prior to treating the wooden bars with
marine varnish. Although the bed roughness material was purchased as 6.1mm
square bars, after waterproofing this mean value was increased to 6.35mm. For
completeness, both the nominal (e.g. L = 7) and accurate values for L are given in
Table 4.1, however all further reference to the roughness spacing will be made on the
basis of the nominal integer values. The positional error in manually fixing the bars
to the base of the flume was estimated to be less than 0.1mm.

PhD Thesis 96 Chapter 4.0


L J Campbell Experimental Methodology
4.5 Hydraulic Conditions

All 12 bed configurations (11 different roughness spacings and a supplementary


smooth bed case) were tested with three open-channel flow depths: H=37, 50, and
85mm, giving a range of relative depths, H/k, between 5.8 and 13.4, and ensuring the
aspect ratio, B/H (where B=flume width), remained above 4.5. The water depth was
measured from the solid flume bed, which implies that the mean flow depth varied
by up to 3.18mm across trials [i.e. the difference between the mean bed levels for the
smooth and L = 2 cases as given in Table 4.1; accurate mean flow depths are
presented in Tables 4.2(a) to (c)]. Whilst this difference is negligible in terms of the
measured flow velocities, it must be acknowledged when calculating terms such as
bed shear stress or friction velocity from the slope-depth product (boundary shear
stress, o = gRS gHS, where is fluid density, R is the hydraulic radius, H and S
are flow depth and bed slope as before, respectively).

Flow conditions were always steady-on-average and were kept globally uniform,
meaning that changes in instantaneous velocities were due to turbulent fluctuations
alone, and that there was no change in mean flow depth along the length of the
channel. Bed slope was fixed at 1/400. With fixed bed slope, discharge was adjusted
to reach uniform flow conditions at the chosen flow depth. Within each subset of
constant flow depth, discharge was therefore free to vary as a function of the bed
topography.

Ranges of key hydraulic and roughness parameters were as follows. Global Reynolds
number, 4260<Re=UR/<26700, signifying that turbulent open-channel flow
prevailed (Re>3000); roughness Reynolds number, 175<Re*=u*k/<241, implying
that the experiments were in the rough-turbulent regime (Re*>70, Section 2.2.1);
Froude number, 0.28<Fr=U/(gR)0.5<0.55, i.e. sub-critical flow (Fr<1.0); and shear
velocity (m/s), 0.0276<u*= (o/ )0.5<0.0380; where U=bulk mean streamwise
velocity (=Q/A), R=hydraulic radius, =kinematic viscosity, o=boundary shear
stress, g=acceleration due to gravity.

PhD Thesis 97 Chapter 4.0


L J Campbell Experimental Methodology
L k [mm] P [mm] La Mean (z)
(nominal (roughness (centre-to- (accurate Bed (porosity as a function of

Level
roughness height) centre roughness z-coordinate, axes not to
spacing) pitch) spacing) scale)
[mm]
Smooth --- --- --- 0.00 z/k 1

bed! 0
0 1

2 6.35 12.4 1.95 3.18 z/k 1

0
0 1/2 1

3 6.35 18.4 2.94 2.12 z/k 1

0
0 2/3 1

4 6.35 24.4 3.85 1.59


z/k 1

0
0 3/4 1
5 6.35 30.6 4.84 1.27 z/k 1

0
0 4/5 1
6 6.35 36.9 5.87 1.06 z/k 1

0
0 5/6 1

7 6.35 43.4 6.80 0.91 z/k 1

0
0 6/7 1

8 6.35 49.8 7.80 0.79


z/k 1

0
0 7/8 1
10 6.35 62.2 9.72 0.64 z/k 1

0
0 9/10 1
12 6.35 74.8 11.75 0.53 z/k 1

0
0 11/12 1

15 6.35 93.7 14.53 0.42 z/k 1

0
0 14/15 1

20 6.35 124.4 19.45 0.32


z/k 1

0
0 19/20 1

Table 4.1: Roughness geometry parameters L, k and P (refer to Figure 4.13), together with mean bed
level and variation in porosity, (z). Mean bed level is taken as the mean elevation of the roughness
topography above the solid flume bed. !Smooth bed discharge measured only (no PIV measurements).
PhD Thesis 98 Chapter 4.0
L J Campbell Experimental Methodology
4.6 Chapter Closure: Table of Experiments

In conclusion to this chapter, the following table of experiments, Table 4.2, draws
together many of the roughness and hydraulic conditions (mean flow depth, relative
submergence and flume aspect ratio) set out in Sections 4.4 and 4.5, in addition to
showing some key PIV analysis and double-averaging parameters (image resolution,
the physical size of interrogation regions, and some idea of the spatial averaging
window). Furthermore, each roughness configuration is classified according to the
terminology of Morris (1955) and Perry et al. (1969) as presented in Sections 2.3.2.1
and 2.3.2.2, respectively. Owing to the number of experiments, Table 4.2 is laid out
in the form of three sub-tables; this subdivision is on the basis of flow depth (i.e.
Tables 4.2 (a), (b), and (c) are for 37, 50, and 85mm flow depths, respectively).

Parameters readable from experiment tag


[as used in Tables 4.2 (a) to (c)]:

S 400 _ H 37 _ L6

Bed slope, Nominal flow Roughness


S = 1/400 depth, H = 37mm spacing, L = 6

PhD Thesis 99 Chapter 4.0


L J Campbell Experimental Methodology
ROUGHNESS and FLOW PARAMETERS PIV and VOLUME AVERAGING PARAMETERS
EXPERIMENT ROUGHNESS CLASSIFICATION MEAN RELATIVE FLUME PIV IMAGE PHYSICAL SIZE OF SPATIAL
FLOW SUBMERGENCE ASPECT RESOLUTION PIV AVERAGING
TAG
(after Morris, (after Perry et al., DEPTH RATIO (m/pixel) INTERROGATION WINDOW
1955) 1969) (mm) REGION
(giving slope, (mm) (number of
skimming,
flow depth, and Hm Hm/k B/Hm roughness pitches)
wake
bar spacing) d-type or k-type (32 32 pixels)
interference, or
(Hm=H- (k=6.35mm) (channel
isolated
mean bed width/flow
roughness flow
level) depth)
S400_H37_L2 skimming d-type 33.82 5.33 11.83 49.5 1.58 1.58

S400_H37_L3 skimming d-type 34.88 5.49 11.47 41.9 1.34 1.34


wake
S400_H37_L4 d-type 35.41 5.58 11.30 49.5 1.58 1.58
interference
wake
S400_H37_L5 d-type 35.73 5.63 11.20 36.9 1.18 1.18
interference
wake
S400_H37_L6 d-k transitional 35.94 5.66 11.13 50.7 1.62 1.62
interference
isolated
S400_H37_L7 k-type 36.09 5.68 11.08 44.4 1.42 1.42
roughness
isolated
S400_H37_L8 k-type 36.21 5.70 11.05 51.0 1.63 1.63
roughness
isolated
S400_H37_L10 k-type 36.36 5.73 11.00 79.5 2.54 2.54
roughness
isolated
S400_H37_L12 k-type 36.47 5.74 10.97 84.3 2.70 2.70
roughness
isolated
S400_H37_L15 k-type 36.58 5.76 10.93 91.1 2.91 2.91
roughness
isolated
S400_H37_L20 k-type 36.68 5.78 10.91 128.0 4.10 4.10
roughness

Table 4.2 (a): FLOW DEPTH H=37mm: Table of experiments showing roughness, flow, PIV and averaging parameters.

PhD Thesis 100 Chapter 4.0


L J Campbell Experimental Methodology
ROUGHNESS and FLOW PARAMETERS PIV and VOLUME AVERAGING PARAMETERS
EXPERIMENT ROUGHNESS CLASSIFICATION MEAN RELATIVE FLUME PIV IMAGE PHYSICAL SIZE OF SPATIAL
FLOW SUBMERGENCE ASPECT RESOLUTION PIV AVERAGING
TAG
(after Morris (after Perry et al. DEPTH RATIO (m/pixel) INTERROGATION WINDOW
1955) 1969) (mm) REGION
(giving slope, (mm) (number of
flow depth, and skimming, Hm Hm/k B/Hm roughness pitches)
bar spacing) wake d-type or k-type (32 32 pixels)
interference, or (Hm=H- (k=6.35mm) (channel
isolated mean bed width/flow
roughness flow level) depth)
S400_H50_L2 skimming d-type 46.82 7.37 8.54 48.8 1.56 1.56

S400_H50_L3 skimming d-type 47.88 7.54 8.35 49.6 1.59 1.59


wake
S400_H50_L4 d-type 48.41 7.62 8.26 52.0 1.66 1.66
interference
wake
S400_H50_L5 d-type 48.73 7.67 8.21 49.6 1.59 1.59
interference
wake
S400_H50_L6 d-k transitional 48.94 7.71 8.17 50.5 1.62 1.62
interference
isolated
S400_H50_L7 k-type 49.09 7.73 8.15 53.3 1.71 1.71
roughness
isolated
S400_H50_L8 k-type 49.21 7.75 8.13 51.3 1.64 1.64
roughness
isolated
S400_H50_L10 k-type 49.36 7.77 8.10 79.6 2.55 2.55
roughness
isolated
S400_H50_L12 k-type 49.47 7.79 8.09 83.6 2.67 2.67
roughness
isolated
S400_H50_L15 k-type 49.58 7.81 8.07 91.9 2.94 2.94
roughness
isolated
S400_H50_L20 k-type 49.68 7.82 8.05 125.0 4.00 4.00
roughness

Table 4.2 (b): FLOW DEPTH H=50mm: Table of experiments showing roughness, flow, PIV and averaging parameters.
.
PhD Thesis 101 Chapter 4.0
L J Campbell Experimental Methodology
ROUGHNESS and FLOW PARAMETERS PIV and VOLUME AVERAGING PARAMETERS
EXPERIMENT ROUGHNESS CLASSIFICATION MEAN RELATIVE FLUME PIV IMAGE PHYSICAL SIZE OF SPATIAL
FLOW SUBMERGENCE ASPECT RESOLUTION PIV AVERAGING
TAG
(after Morris (after Perry et al. DEPTH RATIO (m/pixel) INTERROGATION WINDOW
1955) 1969) (mm) REGION
(mm) (number of
skimming, Hm Hm/k B/Hm roughness pitches)
(giving slope, wake d-type or k-type (32 32 pixels)
flow depth, and interference, or (Hm=H- (k=6.35mm) (channel
bar spacing) isolated mean bed width/flow
roughness flow level) depth)
S400_H85_L2 skimming d-type 81.82 12.88 4.89 101.2 3.24 3.24

S400_H85_L3 skimming d-type 82.88 13.05 4.83 88.0 2.82 2.82


wake
S400_H85_L4 d-type 83.41 13.13 4.80 97.0 3.10 3.10
interference
wake
S400_H85_L5 d-type 83.73 13.19 4.78 82.9 2.65 2.65
interference
wake
S400_H85_L6 d-k transitional 83.94 13.22 4.76 78.9 2.52 2.52
interference
isolated
S400_H85_L7 k-type 84.09 13.24 4.76 87.7 2.81 2.81
roughness
isolated
S400_H85_L8 k-type 84.21 13.26 4.75 100.2 3.21 3.21
roughness
isolated
S400_H85_L10 k-type 84.36 13.28 4.74 79.7 2.55 2.55
roughness
isolated
S400_H85_L12 k-type 84.47 13.30 4.73 83.5 2.67 2.67
roughness
isolated
S400_H85_L15 k-type 84.58 13.32 4.73 90.8 2.91 2.91
roughness
isolated
S400_H85_L20 k-type 84.68 13.33 4.72 134.7 4.31 4.31
roughness

Table 4.2 (c): FLOW DEPTH H=85mm: Table of experiments showing roughness, flow, PIV and averaging parameters.

PhD Thesis 102 Chapter 4.0


L J Campbell Experimental Methodology
4.7 Overview of Results and Discussion Chapters

The following three chapters (5 to 7) are devoted to presenting and discussing the
central findings from the programme of experiments of open-channel flows over
square bars (as detailed in Section 4.6 above).

Chapter 5 begins with an initial brief overview of results with the aim of
demonstrating why spatial averaging is necessary for flows over rough beds.
Attention then moves to a handful of cases to examine results such as velocity and
fluid stress profiles in more detail, with the emphasis firmly placed on double-
averaged flow characteristics in the roughness layer. By doing so, results from this
study can be placed in the wider context of existing square bar and double averaging
research.

Following on from the results presented in Chapter 5, Chapter 6 is dedicated to


illustrating the detailed interaction between fluid flow and a rough bed. This is
achieved by shifting focus from mean flow variables (for example, temporally and
double averaged velocity profiles) to the spatial variation in the flow in the vicinity
of the fixed bar topography, i.e. the spatial variation in the form-induced velocity
components, u~ and w ~ . The pairing of cross-correlation PIV measurement with

double averaging methodology has facilitated an exceptionally thorough


investigation of how fluid velocities vary close to a rough boundary. This work
culminates in an entirely new application for quadrant analysis (a method
traditionally used to visualise and quantify the turbulent structure of a flow);
quadrant maps of the form-induced velocity components ( u~ and w
~ ) are reported for

the first time.

In Chapter 7, the final results and discussion chapter, the roughness geometry is
classified according to the work of Perry et al. (1969) as d-type or k-type (as
discussed in Section 2.3.2.2, see Table 4.2 for experiment classification). A key
finding of this study is that channel discharge is extremely sensitive to roughness
spacing, with a marked peak in conveyance occurring for lateral square bars
positioned periodically at the transition point between d-type and k-type wall
roughness. This exciting new finding is hence termed the transition phenomenon.

PhD Thesis 103 Chapter 4.0


L J Campbell Experimental Methodology
Chapter 7 therefore examines the effect of roughness spacing on the global flow,
with particular emphasis placed upon the transition from d-type to k-type roughness
configurations.

PhD Thesis 104 Chapter 4.0


L J Campbell Experimental Methodology
5.0 DOUBLE-AVERAGED FLOW
CHARACTERISTICS IN THE
ROUGHNESS LAYER

5.1 Introduction

This chapter, the first of three results and discussion chapters, begins by presenting a
general introduction to findings from the extensive experimental programme of
flows over square bars before moving on to take a closer look at selected cases. This
order is chosen to reflect two specific aims. Firstly, by presenting real data, the
author aims to demonstrate and underline the need for spatial averaging of the flow
field over rough beds. Secondly, selecting a small number of experiments for fuller
analysis permits comparison and discussion of the current work with existing square
bar and double averaging research. The contents of Chapter 5 can broadly be broken
down into the following three themes,

Section 5.2: Overview of Results A traditional (time-averaged) look at


the global flow, or, Why is Double-Averaging Necessary?

Section 5.3: The Roughness Layer in Open-Channel Flow The move


from time averaged to double-averaged velocity profiles, with particular
emphasis on the near-bed region, i.e. The Double-Averaged Velocity
Distribution over Transverse Square Bars.

Section 5.4: How turbulent and form-induced fluid stresses compare and
interact: The Nature of Fluid Shear Stress.

PhD Thesis 105 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
5.2 Why is Double Averaging Necessary?

By way of introduction to the dataset, Figure 5.1 shows contours of time averaged
streamwise velocity for the middle flow depth (H = 50mm) over all bar spacings
(equivalent figures for the other two flow depths can be found in Appendix C,
Figures C.1 and C.2).

Flow depth, H=50mm


Velocities are
normalised by bulk
mean velocity (U =
Q/A)
Slope, depth,
roughness parameters
in experiment tag as
described in Section
4.6

Figure 5.1: Contours of time-averaged streamwise velocity. Note: subplots are not drawn to
the same horizontal scale.

PhD Thesis 106 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
The darkest blue regions in the immediate vicinity of the bed between successive
square bars correspond to upstream flow ( u <0) associated with flow recirculation.
In other words, these areas are dominated by large eddies that form as the flow
separates in the lee of each roughness obstacle. As the bar spacing increases these
recirculation regions occupy progressively less of the inter-bar gap, until the mean
flow is able to reattach to the main bed between bars. Mean flow reattachment to the
main bed is generally a characteristic of the flow pattern overlying roughness
configurations L 7. It is therefore evident from Figure 5.1 that the programme of
experiments successfully captured a wide range of rough bed flows, from the tightly
spaced bar arrangements (beginning with L = 2) which generate skimming, or quasi-
smooth flows, through wake interference patterns, to the widely spaced bars
(towards the L = 20 end of the spectrum) that behave as isolated roughness
obstacles.

As we move away from the bed one would expect the effect of the roughness on the
time-averaged flow field to diminish. In terms of Figure 5.1 this decreasing
influence is illustrated by a gradual smoothing of the coloured contours such that
there is little variation in the streamwise (x) direction. When the contour lines
become parallel to the bed across one whole roughness pitch we can conclude that
time-averaged and double-averaged profiles will necessarily coincide. However, the
same cannot be said for regions exhibiting velocity variation in the streamwise
direction, and it should be evident from Figure 5.1 that there is significant spatial
variation in the x direction throughout much of the flow depth, most particularly for
those cases classed as isolated roughness flows. In these regions the double-
averaged and time-averaged flow variables will differ, with the exact relation
between the two depending on location relative to the fixed bed. It is therefore
imperative to introduce double averaging methodology to gain insight into spatial,
form-induced fluctuations of time-averaged flow variables in the vicinity of a rough
bed.

5.2.1 Spatial Variability in the Roughness Layer

To further illustrate the spatial variation in the flow field, and to provide a rough

PhD Thesis 107 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
working estimate of the extent of influence of the square bar bed roughness, Figure
5.2 shows temporally-averaged streamwise velocity profiles ( u ( x, z ) ) from a
number of different locations along the bed. The form of these velocity profiles in
the vicinity of bed roughness varies considerably, and without double averaging it is
unclear which is most representative of the flow field; Section 5.3 will reveal the
answer to this question.

" ! ! !

~2.5k

!
z=0
" ! !

Figure 5.2: Temporally averaged streamwise Figure 5.3: The variation in the four
velocity profiles above roughness spacing L=5 velocity profiles shown in Figure 5.2
(wake interference flow). Symbols indicate the (roughness spacing L=5). The dashed line
plotting locations in relation to the fixed bed. highlights the tangible zone of influence of
the bed roughness.

With the four profiles shown in Figure 5.2 collected and displayed on the same
velocity axis (Figure 5.3) it is evident that the bed roughness influences the bottom
2.5k of the velocity field in this experiment. This is however subject to some
variation, most notably with roughness spacing, L. The example chosen for
illustration in Figures 5.2 and 5.3 is bed configuration L = 5 (i.e. the gap between
square bars is 4k), which induces what may be considered as wake interference flow.
As noted, wider roughness spacings that allow the overlying flow to reattach to the
solid bed between bars cause spatial disturbances that penetrate much higher into the
flow. Streamline plots of the time-averaged flow patterns recorded for L = 3 (a
skimming or d-type flow) and L = 12 (isolated roughness or k-type flow), Figures
5.4 and 5.5 respectively, help us understand why this should be the case.

PhD Thesis 108 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
Figure 5.4: Streamlines (time-averaged) Figure 5.5: Streamlines (time-averaged)
overlying a contour map of the time- overlying a contour map of the time-
averaged streamwise velocity for averaged streamwise velocity for isolated
skimming flow over L=3 bar roughness roughness flow over L=12 bar roughness
(H=85mm, bottom half of the flow shown (H=85mm, bottom half of the flow shown
only). Note that there is little streamline only). Note the marked streamline
disturbance above the roughness tops. curvature throughout much of the plotted
area. White diamond denotes the area of
mean flow reattachment (see text).

The streamlines over the skimming flow example (L = 3, Figure 5.4) serve to show
that the space between bars is completely occupied by a large persistent vortex of
similar vertical dimension to the roughness height, and with horizontal dimension
necessarily equal to the gap length. There is also a second, smaller, counter-rotating
corner eddy immediately downstream of the bar. Both theses vortices are almost
wholly contained in the cavity between the roughness elements. The effect of eddies
entirely filling the interfacial sub-layer up to the level of the roughness crests is to
smooth the interface between the overlying flow and the sharp, angular geometry of
the rough bed. Above the cavity velocity profiles exhibit negligible dependence on
streamwise location, and for this reason the streamlines in the L = 3 case are
reasonably straight and parallel above the tops of the square bar roughness. As the
roughness gap grows however, the stable vortex cannot fill the entire gap length and
the square bars become exposed to the main flow (somewhat reminiscent of the
transition between hydraulically smooth and rough flow regimes where, with
increasing Reynolds number, the viscous sub-layer is progressively reduced to

PhD Thesis 109 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
reveal the influence of bed roughness). Hence, in contrast to the uniform streamlines
for L = 3, there is marked streamline curvature persisting some distance into the flow
(~6k from the solid bed) in the isolated roughness example (L = 12, Figure 5.5). This
wavy pattern is generated as a direct result of the repeating sequence of 1) flow
separates behind a square bar, 2) flow reattaches to the main bed, 3) in the absence
of cavity-filling vortices, flow encounters the angular geometry of the next bar and
is forced upwards over the crest. Clearly this flow pattern will only occur for the
isolated roughness flows and explains the increasing zone of marked spatial velocity
variation evident in Figure 5.1 for roughness spacings L7 to L20. Note that the
reattachment area is marked by the white diamond in Figure 5.5 (this is a qualitative
assessment achieved by observing the location at which the streamlines of the time-
averaged flow above the bar crests rejoin, and become parallel to, the channel bed),
and lies some 30-32mm from the downstream face of the upstream square bar. This
gives a reattachment length of 4.7-5k, which compares well with previous estimates
(4.8-5k, e.g. Cui et al., 2000; Leonardi et al., 2003). As the inter-bar gap is 5k in the
L = 6 case, L = 7 is the narrowest bar spacing configuration which allows full
reattachment to the main bed, hence the increasing spatial variation in velocity for
spacings L7 (Figure 5.1). As this behaviour is so strongly dependent on the
reattachment length, which is largely unaffected over the range of measured
Reynolds number, the same pattern appears for all flow depths, H = 37, 50, and
85mm (see Appendix C), with the most marked variation for H = 37mm where the
wavy flow pattern, i.e. the roughness layer, occupies the whole flow depth over the
wider bar spacings.

Considering the effect of roughness pitch on the zone of spatial variability in flow
variables, the approximation of 2.5k above the solid bed, an estimate following from
the wake interference example, L = 5 (Figures 5.2 and 5.3), should be treated as a
conservative estimate for wider bar spacings. Furthermore, one would expect
turbulent quantities to exhibit greater sensitivity (i.e. the effect of the square bars
would penetrate further into the flow depth) as compared to the mean flow velocities
examined thus far. For example, the primary Reynolds stress ( u w ) is subject to
spatial variation in the streamwise direction up to z 5k. Figures 5.6 and 5.7 depict

PhD Thesis 110 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
this variation in Reynolds stress profiles (with data from the same experiment as
shown in Figures 5.2 and 5.3).

" ! ! !
~5k

z=0
!
" ! !

u w

Figure 5.6: Temporally averaged Reynolds Figure 5.7: The variation in the four stress
stress profiles above roughness spacing L=5. profiles shown in Figure 5.6 (roughness
Symbols indicate the plotting locations in spacing L=5). The dashed line highlights
relation to the fixed bed. the tangible zone of influence of the bed
roughness.

Using Reynolds stress profiles as an indicator would therefore suggest the presence
of bed roughness can be felt up to 5 roughness heights from the main bed. This
estimate of 5k concurs with the extent of spatial variability in velocity profiles for
isolated roughness flows (L = 7 to L = 20). The zone of variability in velocity
profiles for wake interference flows (L<7) is typically restricted to ~2.5k from the
channel bed as the flow cannot reattach to the main bed and establish a fixed
undulating pattern as is present under isolated roughness conditions. Raupach et al.
(1991) stated that the level of the roughness layer varied between 2k and 5k for
atmospheric boundary layers over various surface vegetations. More specifically, the
lower limit of z = (2-3)k pertains to d-type roughness, and the upper 5k limit to k-
type. For example, Thom (1971) concluded that the roughness layer depth was little
more than the roughness height for closely spaced wheat canopies, whereas Garratt
(1980) found roughness effects as high as 5k into the flow over some rough,
scrublike vegetated surfaces. The current results for flows over square bars are
entirely in agreement with both the pattern and magnitude of these estimates, i.e. the
roughness layer lies between 2.5k (closely spaced bars) and 5k (larger bar pitches)
and comprises the flow region dynamically influenced by the bed roughness. As

PhD Thesis 111 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
such, in the roughness layer it is unclear what the representative velocity profile
should look like. This matter is now addressed.

5.3 The Double-Averaged Velocity Distribution over Transverse Square


Bars

Having established that the roughness layer (interfacial plus form-induced sub-
layers) occupies the flow region from z = 0 (solid bed, i.e. = 1) to (2.5-5)k (where
the upper limit depends upon the roughness geometry), attention now turns to the
form of the double-averaged velocity profile in this layer. Firstly however,
consideration is given to the structure of the velocity profile in the flow region
immediately above the roughness layer (refer back to Figure 2.14 for a schematic of
the possible flow layers in double-averaged open-channel flow).

5.3.1 Above the Roughness Layer: The Potential for a Logarithmic Velocity
Profile

The occurrence, or otherwise, of a physically meaningful logarithmic velocity


profile in open-channel flow depends on a combination of the characteristic
roughness height and water depth. In this respect, Raupach et al. (1991) recommend
that the overlap region between the roughness and outer layers lies in the range (2-
5)k z 0.2H1. Therefore, the best case scenario for the existence of a well-
developed logarithmic layer in the current square bar dataset is with the highest flow
depth tested. For these experiments the roughness height was k = 6.35mm, and the
flow depth, H = 85mm. Hence, one would expect the overlap between the roughness
and outer layers to lie in the range (12.7-31.75) z 17 mm. Clearly, there are few
or no possible solutions for z with this combination of roughness and flow
constraints, and as such care must be taken when interpreting parameters, e.g. shear
velocity, roughness lengths, evaluated from the law of the wall (refer also to

1
Note: The words physically meaningful refer to the fact that in the derivation of the law of the
wall, the log-law range of (2-5)k z 0.2H assumes a zone of constant turbulent stress for z 0.2H,
which, for uniform 2-dimensional open-channel flow, introduces a 11% discrepancy between actual
and assumed stress levels. If the upper bound for fitting the log-law is increased to 0.3H, the error
rises to 16%. Thus, there is no strong physical justification for deriving parameters from a
logarithmic fit to data above 0.2-0.3H. See e.g. Schlichting (1979) for further details of the derivation
of the log-law.

PhD Thesis 112 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
footnote 1). Nevertheless, it is certainly worth investigating the effect of bar spacing
(L) on typical log-law parameters (displacement height, d, roughness length, zo, and
the von Karman constant, - refer to Equation 2.1, Section 2.2.2), with the
understanding that the use of the log-law in such shallow open-channel flows is an
empirical approximation (i.e. the fundamental physical meaning of the derived
parameters remains unclear).

(a) (b)
Grass et al. (1991)
bar height, k=5mm
bar spacing, L=5
bed slope unknown

3
2
1

3 (c)

Present study
bar height, k=6.35mm
bar spacing, L=5
bed slope, S=1/400

Figure 5.8: Comparison of applicability of the log-law for velocity. Experiment S400_H50_L5 is
chosen as the closest match to one of the flow depth/roughness spacing configurations tested by
Grass et al. (1991). Grass et al. velocity profile (where y is the bed-normal coordinate) is given
in subplot (a), logarithmic fitting in subplot (b), whilst subplot (c) shows the velocity profile and
log-law fit to a similar experiment from this project. Zone 1 is the roughness layer, zone 2 the
log-law region, and zone 3 is the outer layer. Differences in the pattern of deviation from the log-
law fit between plot (b) and (c) are explained in the text.
PhD Thesis 113 Chapter 5.0
L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
To begin, Figure 5.8(c) shows a log-law fit to data from experiment S400_H50_L5
(i.e. flow depth 50mm, bar spacing, L = 5). This is presented to compare the current
square bar data to that in the existing literature [e.g. Figures 5.8(a) and (b)]. The data
illustrated in Figure 5.8 are from similar experiments (albeit, with slightly different
roughness heights and bed slopes, which explains the disparity in velocity
magnitudes). In general, streamwise velocity profiles show good agreement with the
log-law over the flow region 1.6k z 0.27H [Grass et al. (1991), subplot 5.8(b)]
and 1.5k z 0.3H [current study, subplot 5.8(c)]. There are however some
differences in experimental technique that account for the inconsistency in the
pattern of deviation from the log-linear trend [noting that the axes are reversed
between Figures 5.8 (b) and (c)]. Firstly, for the Grass et al. (1991) case velocity
data was gathered from only one streamwise location, directly above the crest of a
bar, whereas the velocity profile in subplot (c) is the double averaged distribution.
Thus, the former profile lacks the S-shaped form in the absence of data from the
gap between bars. When velocities in the interfacial sub-layer are fully represented
in the profile, this S-shaped distribution reveals that velocities are indeed slower
throughout the roughness layer than the profile of Grass et al. (1991) would suggest.
Secondly, the current study made use of a glass sheet positioned on the water surface
for measurement reasons (refer to Section 4.3.2.2). This most probably explains why
velocities tend to slow slightly in proximity to the free surface as compared to
those from Grass et al. (1991). In general however, the extent of logarithmic
velocity behaviour from the current study compares favourably, and the decision
was therefore made to evaluate the displacement height, d, roughness length, zo, and
the von Karman constant, , for all square bar experiments.

5.3.1.1 Log-Law Parameters

The displacement height, d, may be interpreted as the boundary datum plane, or


effective bed level (it is the height above the channel bed that the logarithmic
velocity profile sees as the bed). For a range of square bar flows (see Table 2.1)
Knight and MacDonald (1979a) used d = 0 (i.e. the smooth bed value) for L>6.94, d
= k/2 for L = 5.21, and d = k for L = 3.47, which they classed as a skimming, or
quasi-smooth, flow type. The simulated results of Leonardi et al. (2003) follow a

PhD Thesis 114 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
similar pattern, although d remains in the range 0.45k to 0.5k for 5<L<20. For the
current data set, displacement height for each roughness configuration and flow
depth was evaluated from the (double-averaged) velocity profiles according to
Nikora et al. (2002)2. The results (Figure 5.9) would suggest that Knight and
MacDonalds (1979a) d values for L = 5.21 and L = 3.47 were a very good choice.
Furthermore, their deduction that d = 0 for L>6.94 is intuitively easier to accept than
the trend from the current experiments, which is for d to increase for increasingly
isolated roughness flows (L>6, Figure 5.9). Whilst it is unclear why the
displacement height grows towards the L = 20 end of the roughness pitch range, it is
possible that the log-law becomes progressively more violated as the vertical extent
of the flow disturbance increases (i.e. the roughness layer gets thicker as bar spacing
increases, see Sections 5.4.4 and 6.3.3). Indeed, the logarithmic line fits for well-
isolated arrangements were observed to pass through a few data points only.

Figure 5.9: The effect of bar spacing, L, on displacement height, d (measured from the channel
bed) for all 3 flow depths. Recall the square bar roughness height, k=6.35mm. The erratic values
for the lower flow depth, H=37mm, including one negative value for L=4, may indicate that
there is little physical justification for the existence of a logarithmic velocity profile in such
shallow open channel flows.

2
This method is based upon slightly modified Prandtl mixing length theory, where the displacement
height is interpreted as the level that large-scale turbulent eddies feel as the bed origin, and thus
their sizes scale linearly with distance from this level [i.e. mixing length, l = ( z d ) ]. It assumes
that fluctuations of the streamwise velocity, (u u ) l ( d u / dz ) , and of the bed-normal velocity,
( w w ) u* . After substituting the above expression for the mixing length into
/ = u*l ( d u / dz ) , the displacement height can be evaluated simply by linear regression.

PhD Thesis 115 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
The variation in roughness length, zo, with bar spacing, L, is shown in Figure 5.10.
The roughness length is closely related to the roughness topography, most notably
the height, k, and roughness element distribution, which in this case shall be
represented by the bar spacing, L. The pattern of roughness length variation shown
in Figure 5.10 is supported in the existing literature, at least for the higher two flow
depths, H = 50 and 85mm. For example, Figure 5.11 shows roughness length plotted
as a function of roughness density (the inverse of L) as reported by Raupach et al.
(1991). Furthermore, the magnitude of zo from the experiments illustrated in Figure
5.8 are in reasonable agreement, given the different hydraulic conditions Grass et
al. (1991) found zo =0.96mm, whilst the value for experiment S400_H50_L5 is zo
=1.6mm.

Figure 5.10: The effect of bar spacing, L, on roughness length, zo, evaluated from a logarithmic
line fit to the measured double-averaged streamwise velocity profile

Considering values of displacement height, d, and roughness length, zo, in


conjunction helps understand the meaning of each term in relation to the logarithmic
velocity distribution. This is perhaps most evident for spacing L = 2, which is
characterised by a skimming flow pattern. For all three flow depths, the
displacement height for L = 2 is close to the bar height, k, whilst the roughness
length stays close to zero. This means that although the origin of the logarithmic
profile is displaced upwards, almost to the bar crests, the flow senses the effective
datum plane (i.e. at z k) as practically smooth. Both of these observations make

PhD Thesis 116 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
physical sense given the skimming flow pattern, hence the alternative terminology
of quasi-smooth flow.

zo/k

Roughness density

Figure 5.11: The variation in roughness length, zo (normalised with roughness height, k) with

roughness density (note that L=1/roughness density, and x = l x / k , i.e. x =1 for square

bars). Figure adapted from Raupach et al. (1991).

Interestingly, the lower flow depth (H = 37mm; relative submergence, H/k = 5.8)
square bar experiments somewhat surprisingly displayed a reasonable logarithmic
region despite the whole flow depth arguably being dynamically influenced by the
square bar roughness. However, the large variation in both displacement heights and
roughness lengths at this flow depth may indicate that the underlying physics of the
log-law are invalidated for such shallow open-channel flows. Therefore, the physical
meaning of these parameters, as well as the capability to compare their magnitude to
other published values, is most probably lost. For example, McLean et al. (1999)
found double-averaged streamwise velocity profiles measured over wavy bed forms
to be highly logarithmic even though the effect of the bottom topography is
observed throughout the water column. However, logarithmic fits of these averaged
profiles do not yield accurate estimates of the measured total boundary shear stress.

The von Karman constant (usually taken as = 0.4) was also estimated from the
slope of the logarithmic velocity profile in all cases. Values varied between 0.24 and

PhD Thesis 117 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
0.72. This variation is not unprecedented (e.g. Sayre and Albertson, 1961; Dittrich
and Koll, 1997), but again this indicates that the validity of the log-law, together
with the physical meaning of parameters derived from it, requires clarification for
shallow open-channel flows.

In summary therefore, the current square bar data exhibits a log-linear portion in
good agreement with the literature, but parameters based on the log layer must be
treated with caution. The primary objective of this study was to investigate open-
channel flow in the roughness layer. Hence, the overlying flow regions (i.e. the log
layer and the outer layer, which in this case is modified by the glass sheet and shall
not be analysed) are not discussed beyond this section. Indeed, it has already been
stated that in layers above the roughness layer temporally- and double-averaged
profiles coincide. As a primary aim of this project was to examine the characteristics
of the double-averaged velocity distribution, the scope of the following sections is
strictly limited to the roughness layer. Even at the stage of designing the
experimental programme, combinations of roughness height and flow depths were
chosen to allow good spatial resolution in the interfacial and form-induced sub-
layers, at the potential expense of a well-developed logarithmic layer. The velocity
profile in the roughness layer is now presented.

5.3.2 Velocity Distribution in the Roughness Layer

This thesis opened with a quote from Monin and Yaglom (1977) stating that it was
impossible to hope for any simple general rules for the velocity distribution in the
roughness layer. However, as described in the literature review a small number of
studies have indicated that, under certain conditions, either a linear or exponential
double-averaged velocity profile may prevail in the lower portion of the roughness
layer, the interfacial sub-layer (e.g. Nikora et al., 2001, 2004; Barrantes and Madsen,
2000). Nikora et al. (2004) suggest that an exponential profile would appear when
roughness elements are well submerged (i.e. H>>k) and dA / dz 0 . Forest canopies
in the atmospheric boundary layer generally fulfill these conditions, but the current
square bar data do not as they are from shallow open-channel flows with low
relative submergence ( H / k 13.3, Table 4.2). However, there is potential for the
square bar flow conditions to yield a linear profile. Data illustrating the double-

PhD Thesis 118 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
averaged linear distribution of velocity from flows over various surface roughnesses,
ranging from three-dimensional arrangements of beads, cubes and spheres to
triangular bars, are shown in Figure 5.12 (reproduced from Nikora et al., 2004). This
poses a key question: do double-averaged profiles for flows over certain square bar
arrangements follow a simple linear trend in the interfacial sub-layer, and if so,
why? To answer this, we will consider the influence of bar spacing on the double-
averaged velocity profile by examining each flow type in turn, i.e. skimming, wake
interference, and isolated roughness flows.

Figure 5.12: Double-averaged velocity profiles over a range of surface roughness (from Nikora
et al., 2004), showing good agreement with a linear fit in the interfacial sub-layer. zc is the
vertical coordinate at the level of the roughness tops, whilst lc is described as a scale
characterising flow dynamics below the roughness crests.

5.3.2.1 Skimming Flow (L = 2 and L = 3)

As seen in earlier in this chapter (e.g. Figure 5.4) the interfacial sub-layer in
skimming flow types is dominated by a large fixed eddy filling the entire gap
between successive square bars. This pattern allows the overlying flow to effectively
skim over the tops of the square bars. One may therefore expect the double
averaged velocity profile above the roughness tops ( z > zc ) to appear similar in form

to that over a smooth bed, whilst at lower levels ( z zc ) the double averaged profile

will be strongly influenced by both the shape and strength of the inter-bar vortex.
Figure 5.13 shows the double-averaged streamwise velocity profiles for cases L = 2

PhD Thesis 119 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
and L = 3 (both are for flow depth H = 37mm, plots for all flow depths may be found
in Appendix C, Figures C.3 to C.5).

The cavity between bars in the L = 2 case shown in Figure 5.13(a) is square (as its
length is equal to the height of the bars that enclose it). Hence the time-averaged
cavity eddy is sensibly symmetric about its centre of rotation, which is
approximately located in the middle of the gap at level z = k / 2 (see Figure 5.14).
Such a steady near-circular rotational pattern produces roughly equal and constant
velocity gradients, u / z , in the lower and upper portions of this eddy [but
clearly with opposite sign, see Figure 5.13(a)]. Although the eddy is almost
symmetric, the absolute velocity magnitude at the upper and lower bounds of the
vortex must satisfy the boundary conditions of no-slip, i.e. u = 0 at z = 0 , and a

slip velocity of u = u ( zc ) at z = zc , hence the switch in the sign of the double-

averaged velocity occurs around z = k / 3 . The skewing of the rotational path


towards the top right in Figure 5.14 reflects the eddy being driven by the left-to-right
momentum of the overlying fluid layers.

Figure 5.13: Double-averaged


(a) streamwise velocity distribution,
u ( z ) (normalised with bulk mean
velocity, U=Q/A) for skimming flow
cases (a) L=2, and (b) L=3 (flow
depth, H=37mm). In both cases
momentum is, on average, carried

(b) upstream (negative x-direction) for the


bottom third of the interfacial sub-
layer, i.e. u < 0 for z < k / 3 . The

dashed line is at the level of the

roughness tops, z = zc .

Increasing bar spacing by one element height from L = 2 to L = 3 causes the counter-
clockwise cavity vortex, and hence the double-averaged velocity profile to lose its
symmetry and angularity [Figure 5.13(b); recall streamlines for a L = 3 configuration
were shown in Figure 5.4)]. However, the gap remains fully occupied by a zone of

PhD Thesis 120 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
recirculating fluid, there is a linear velocity distribution over the range
0.25k z k , and the flow pattern can still be termed skimming. In both cases (L =
2 and L = 3) velocity profiles above the cavity resemble those in a flat channel,
albeit with a reduced flow depth of H-k (Cui et al., 2000). Such profiles above the
interfacial sub-layer of skimming flows are often therefore termed quasi smooth
(after Morris, 1955).

Figure 5.14: Time-averaged


streamlines for roughness
configuration L=2, flow depth
H=37mm. The eddy is driven by
the left-to-right momentum of
the overlying layers, and has a
centre of (clockwise) rotation
located roughly in the middle of
the gap at z k / 2 .

As a final note to this subsection, it can be tempting to assume that in skimming


conditions there is limited interaction between the interfacial (cavity) sub-layer and
the main flow, aside from the steady driving force from the mainstream to the
underlying eddy at the level of roughness crests. However, Djenidi et al. (1994) used
flow visualisation to demonstrate relatively strong momentum exchange associated
with strong ejections of fluid from the cavity, telling that the cavity vortex may not
be as perfectly formed at instants through time as the time-averaged picture (Figure
5.14) would suggest. Animations of instantaneous velocity contours and streamlines
for the square bar flows considered in this study corroborate this finding (see
Appendix C, Part 4 for a sequence of instantaneous streamline plots for bar spacing
L = 2).

5.3.2.2 Wake Interference Flow (L = 4, 5, 6)

Upon first glance the double-averaged velocity distributions for wake interference
cases, L = 4, 5, and 6 (Figure 5.15) look very similar to the L = 3 skimming flow
regime. Indeed, in many ways they are. For example, both flow types show a

PhD Thesis 121 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
characteristic S-shaped profile, governed by the largest cavity eddies moving fluid
in the negative x-direction close to the bed, and in the positive streamwise direction
for z 0.15k . The main difference between the profiles shown in Figure 5.15 and
5.13(b) is that the interface linking the interfacial and form-induced sub-layers
becomes progressively smoother (the velocity gradient is continuous) with
increasing bar pitch. In other words there is more dynamic (momentum) exchange
across the bed-parallel plane at the level of the roughness tops, hence the flow in the
cavity becomes less isolated from the fluid motion in the free stream, and vice versa.
There is also a systematic reduction in the height of the inflection point in the
double-averaged flow (i.e. where u / z = 0 ). For example, this occurs at z 2mm
for L = 2, at z 1.3mm for L = 3, at z 1.1mm for L = 4, and at z 0.7mm for L =
5.

(a)

(b)

(c)

Figure 5.15: Double-averaged streamwise velocity distribution, u ( z ) (normalised with bulk


mean velocity, U=Q/A) for wake interference flow cases (a) L=4, (b) L=5, and (c) L=6 (flow

depth, H=37mm). The dashed line is at the level of the roughness tops, z = zc .

PhD Thesis 122 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
5.3.2.3 Isolated Roughness Flow ( L 7)
.
This set of bar spacings is interesting as it encompasses a large change in roughness
pitch. Behaviour at the L = 7 end of the spectrum is similar to that in the above
Section 5.3.2.2, but as we move to the sparse element arrangements at the L = 20
end of range the influence of the main cavity eddy diminishes. Eventually, for
L 10 , the negative part of the double-averaged velocity distribution disappears
(Figure 5.16).

(a) (d)

(b) (e)

(c) (f)

Figure 5.16: Double-averaged streamwise velocity distribution, u ( z ) (normalised with


bulk mean velocity, U=Q/A) for isolated roughness flow cases (a) L=7, (b) L=8, (c) L=10, (d)
L=12, (e) L=15, and (f) L=20 (flow depth, H=37mm). The dashed line is at the level of the

roughness tops, z = zc . Note the evolution of the linear trend in the interfacial sub-layer for
the widest bar spacings, L=15 and L=20.

PhD Thesis 123 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
The loss of the S-shaped velocity profile with increasing bar spacing is
accompanied by a systematic trend towards a linear profile over the range 0 z k .
This evolution of this straight-line distribution can be traced from bar spacing L =
10, until it is fully evident throughout the interfacial sub-layer for both L = 15 and L
= 20 [Figures 5.16(e) and 5.16(f)]. The underlying reason for the observed linear
velocity distribution is as follows.

(a) z z z

Profile Profile Full


type 1 + type 2 = Double-
averaged
(Double- (Double- profile
averaged averaged
over over
recirculation reattached
zone) zone)

u ( x, z ) u ( x, z ) u ( z )

(b)

Profile Profile Full


type 1 type 2 Double-
averaged
profile

Figure 5.17: (a) Sketch (not drawn to scale) illustrating how profile type 1, the S-shaped
profile associated with flow in the cavity vortex, and flow type 2 associated with areas where
flow has reattached to the main bed, combine (by weighted sum) to yield a linear double-
averaged velocity profile in the interfacial layer between successive square bars. Such a linear
profile was realised for L=15 and L=20, as plotted in Figures 5.16(e) and (f), respectively. (b)
Double-averaged profiles type 1 and 2 and the full double-averaged profile (i.e. the weighted
sum of Profile type 1 + Profile type 2) for real data (experiment S400_H50_L15)

The precise balance between the recirculating part [typified by profile type 1 see
Figure 5.17(a)] and the reattached part [with profile type 2, Figure 5.17(a)] of the

PhD Thesis 124 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
fluid flow in the inter-bar gap is an important control on the shape of the double-
averaged velocity profile. When the cavity vortex occupies the whole gap, or a
sufficient percentage of the gap, the net fluid displacement close to the bed will be
upstream. In other words there is enough gap length characterised by profile type 1
to ensure that the double-averaged profile is negative near the channel bed. In
contrast, profile type 2 is not affected by recirculation zones, and double-averaged
velocities in such regions are always positive. Summing the negative (profile type
1) and positive (profile type 2) velocity distributions in the interfacial sub-layer
results in a linear trend seen schematically in Figure 5.17(a), and with real data from
case S400_H50_L15 in Figure 5.17(b) (see also profiles for L = 15 and L = 20 in
Figures 5.16(e) and 5.16(f), respectively). It should be noted that Figure 5.17
presents a simplified view of the full double-averaged velocity distribution because
profiles type 1 and 2 are in fact also double-averaged. Profile type 1 is the result of
double-averaging velocity over a streamwise length corresponding to the
recirculating portion of the near-bed flow, whilst profile type 2 is the double-
averaged distribution across the reattached area of the flow. Thus, the full double-
averaged profile is then simply the weighted average of profiles type 1 and 2.

As discussed (Section 2.3.3), the streamwise reattachment length downstream of a


square bar of height k is around 5k. Hence, one may conclude that a double-averaged
S-shaped profile will prevail when the length of separated flow, SF (=5k), occupies
at least 70% of the gap (as in the L = 8 configuration). Additionally a linear velocity
profile appears in the interfacial sub-layer when SF drops to around 35% of the gap
length (as in L = 15). As this linear trend is present for the widest bar spacing
considered (L = 20, where SF = 26% of gap), the lower bound of SF to maintain a
linear profile is unclear. Presumably, when bar pitch is sufficiently long, profile
type 2 (Figure 5.17) begins to dominate and the double-averaged velocity
distribution ceases to be linear and becomes increasingly like a smooth bed profile.

The influence of relative submergence on the near-bed double-averaged velocity


profiles appears to be minimal, at least across the range tested. However, for
completeness, the velocity profiles for all flow depths (H = 37, 50 and 85mm) are
given in Appendix C, Part 2. The effect of channel bed slope does though require
clarification (as it has not been measured). Although the reattachment length is

PhD Thesis 125 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
primarily scaled with roughness height, the strength of the recirculation, and hence
the associated velocity magnitudes in the cavity vortex (the velocity extremes of
profile type 1) will most likely reduce with flatter bed slopes. One may therefore
expect a linear profile to develop at smaller bar pitches when the bed slope is
reduced.

In summary, the double averaged streamwise velocity profile is indeed linear over
the range of bar spacings 15 L Lmax , where Lmax is the upper limit for generating

a linear trend in the interfacial sub-layer (at least L = 20). This velocity behaviour
corresponds closely to one of the models proposed by Nikora et al. (2004 data are
shown in Figure 5.12, as discussed in Section 2.4.3), although the explanation given
here is more mechanistic than the analytical reasoning presented by Nikora et al.
(2004). Unfortunately the roughness and flow conditions suggested by Nikora et al.
(2004) for a linear double-averaged velocity distribution (monotonically decreasing
roughness geometry function, A(z), together with monotonically increasing total
drag as the channel surface is approached) do not quite match those of the isolated
roughness square bar flows considered here. Nevertheless, the current results are in
agreement with the double-averaged velocity distributions in the interfacial sub-
layer shown (but not discussed) by Cui et al. (2003), MacDonald (2000) and Kanda
et al. (2004).

The velocity profiles in this thesis pertain to flow over a simple rough bed,
compared to three-dimensional arrangements or natural sediment beds. However, a
similar argument to that given above may be employed to explain why a linear
profile would evolve over more complex surface geometries. For example, in flow
close to a gravel bed, there will be a mixture of troughs and peaks, and more open
pathways through the bed between the surface grains. A trough-peak type flow path
will produce areas of flow recirculation (and hence something akin to profile type
1, Figure 5.17), whilst a less cluttered pathway will yield only positive
(downstream) velocities (profile type 2, Figure 5.17). The key difference between
three-dimensional roughnesses (e.g. spheres, cubes, gravel) and the two-dimensional
square bars is that strong recirculation regions downstream of each square bar persist
across the whole width of the channel. Therefore, a greater area (streamwise length)

PhD Thesis 126 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
of obstacle-free bed is required to generate sufficient quasi-smooth velocity
behaviour to counteract and balance the S-shaped profile. These patterns, and the
associated linear distribution of double-averaged velocity, have been reproduced in a
separate experimental programme conducted in the same flume as the experiments
described herein for flows over spheres arranged in a cubic pattern (C Manes,
personal communication). Such a simple trend, which is somewhat contrary to the
opening quote from Monin and Yaglom (1977), only appears when the velocity field
is averaged in both temporal and spatial domains.

Interestingly, all the double-averaged velocity profiles (with the exception of


skimming flow case L = 2) shown above for flow depth H = 50mm (Figures 5.13,
5.15, and 5.16) are characterised by u ( zc ) 0.5U , where zc is the level of the bar

tops, and U is the bulk mean velocity. Reference to Appendix C shows that this
approximation is equally applicable to the other two flow depths, H = 37 and H =
85mm. The reason why this should be the case, especially over such a wide range of
bar spacings, and hence flow types, remains unclear.

5.4 The Nature of Fluid Shear Stress

Within the scope of the conventional Reynolds-Averaged Navier-Stokes equations


(RANS) fluid stress comprises viscous and turbulent components. As introduced in
Section 3.5, volume-averaging the RANS equations results in a third stress term, the
form-induced stress. Assuming viscous fluid stress to negligible, the primary
stresses to consider with the current square bar two-dimensional flows are turbulent
(Reynolds) and form-induced stresses,

(
xz = u w + u% w% ) (5.1)

(1) (2)

The primary Reynolds stress, term (1) in Equation (5.1), arises from turbulent
mixing, such that fluid mixing from slower moving lower layers tends to impede
momentum transfer in the faster moving overlying layers. Hence a positive w
temporal fluctuation is associated with a negative u fluctuation. The reverse is also

PhD Thesis 127 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
true, i.e. a downward transfer of faster moving fluid ( w < 0 ) tends to speed up the
velocity in slower moving layers ( u > 0 ). When averaged in time, the correlation

between orthogonal velocity fluctuations is the root of Reynolds stress ( u w ).


The primary form-induced stress also arises from the correlation between two
orthogonal velocity components, the form-induced velocities, u% and w% . In simple
terms, when flow approaches an obstacle, the streamwise velocity may slow down
compared to fluid in the same layer (i.e. u% < 0 ) whilst the bed-normal velocity
increases to allow flow up and over the obstruction. ( w% > 0 ). When averaged in
space the correlation between u% and w% is the reason for the primary form-induced
stress ( u% w% ). Clearly in the absence of roughness forms (a smooth bed), form-
induced stress, u% w% , will equal zero.

For steady, two-dimensional, globally uniform open channel flow, the force
associated with the turbulent and form-induced stresses must balance the force from
the downstream component of gravity acting on the fluid. The fluid shear stress
distribution should therefore vary linearly from zero at the free surface, to the bed
shear stress at z = 0 (for a smooth bed). For a rough bed, the shear stress distribution
will deviate from linear at the level of the roughness crests as porosity drops below
1, and as form and viscous drag forces enter into the momentum balance (refer to the
terms in the double-averaged momentum equations for two-dimensional flow, as
given in Section 3.6).

This stress distribution is now examined for a small number of cases, one each for
skimming, wake interference, and isolated roughness flows (form-induced stress
profiles for all cases are however presented in Appendix C, Part 3).

5.4.1 Fluid Shear Stress for Skimming Flow (case S400_H50_L3)

Discussion of the turbulent stress profile forms a logical starting point for examining
fluid shear, as the linear distribution discussed above is sensitive to any deviation
from two-dimensional or uniform flow conditions. Additionally, the distribution of
primary shear stress is the most appropriate tool for evaluating the friction velocity,
u* , and hence bed shear, o (Nezu and Nakagawa, 1993). Bed shear stress

PhD Thesis 128 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
calculated this way3 is used for normalising all double-averaged fluid shear profiles
(Figures 5.18, 5.20, and 5.22).

(a) (b)

u w / u*2 u% w% / u*2

Figure 5.18: Fluid stress profiles for skimming flow case S400_H50_L3. (a) shows spatially-

averaged Reynolds stress, u w (b) shows form-induced stress, uw


% % . Both are

normalised with bed shear stress, o = u* . The dashed line indicates the level of the roughness
2

crests.

Figure 5.18 shows the primary spatially-averaged Reynolds (turbulent, or


apparent) stress distribution, u w (z), alongside the form-induced stress
profile, uw
% % (z) from the same skimming flow experiment, S400_H50_L3. The
fact that the turbulent stress profile is linear above the roughness layer [Figure
5.18(a)] demonstrates that the assumption of two-dimensional flow conditions is a
valid one for the central part of the flume, where - u w is the dominant momentum
flux and other contributions may be considered negligible throughout the depth (as
would be expected by maintaining flume aspect ratios above 4.5).

For this quasi-smooth case, the Reynolds stress profile is linear throughout the
flow depth to within a very short distance from the roughness crests (~1mm).
Consequently the form-induced stress is negligible beyond the tops of the bars, but
instead peaks immediately below z = k [Figure 5.18(b)]. This marked peak, which
constitutes 20% of the total fluid shear at its maximum, compensates for the rapid
decrease in Reynolds stress within the interfacial sub-layer. Turbulent and form-

3
Friction velocity, u* was evaluated by extrapolation of the linear Reynolds stress profile, u w to
mean bed level ( u* " u w o . Bed shear stress was then evaluated from o = u*
2 2

PhD Thesis 129 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
induced momentum fluxes therefore complement each other well right through the
flow depth.

The form-induced flow fields underlying this positive form-induced stress are shown
in Figure 5.19. Subplots (a) and (b) show the spatial variation in the streamwise, u% ,
and bed-normal, w% form-induced velocity components, respectively. Subplot (c)
gives their product, u% w% , whilst subplot (d) is effectively the spatially-averaged form
of (c), i.e. the primary form-induced stress, uw
% % (as shown in Figure 5.18(b), but
repeated here for convenience).

10-3
(a) (c)

(b) (d)

uw
% % (N/m2)

Figure 5.19: The flow patterns underlying form-induced stress for skimming flow case
S400_H50_L3. (a) the streamwise form-induced velocity component, u% , (m/s) (b) the bed-
normal form-induced velocity component, w% ,(m/s) (c) their product, u% w% , (m2/s2) and (d) form-

induced stress, uw
% % , (N/m2).

By breaking down form-induced stress into its components, the areas of faster than
average and slower than average fluid motions become evident. The pattern for
the streamwise component, u% [Figure 5.19(a)] is fairly symmetrically distributed
and lies totally within the gap between bars. In this case four clear areas dominate
and represent the relative magnitude of fluid velocities in both the downstream and
upstream directions. For example, an area of positive u% [the red areas in Figure

PhD Thesis 130 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
5.19(a)] can either signify that fluid is moving faster downstream than the double-
average, or faster upstream (if both the temporally and double-averaged velocities
are negative at that point). The upwards and downwards couple of w% [Figure
5.19(b)] is not as centrally located in the cavity as u% , because there are additional
small vortices occupying the region on the downstream face of each bar (recall
Figure 5.4). This slight offset in w% generates an extended chequerboard pattern
when the product of form-induced velocities, u% w% , is taken, as seen in Figure
5.19(c). The negative (blue) areas of u% w% shown in this figure contribute positive
form-induced stress. As these are generally larger in size and strength when
compared to the positive (red) regions of u% w% , the resultant form-induced stress
profile Figure 5.19(d) is largely positive. The local reduction in uw
% % around z =
3mm occurs because of the relatively slow moving fluid towards the centre of the
main cavity vortex.

This decomposition of form-induced stress to explain its distribution may seem


obvious, but it is important to understand the underlying patterns as although they
repeat to some degree for wider bar spacings the areas of positive and negative
u% and w% are not so distinct once the cavity vortex spreads. This is true for the wake
interference case below.

5.4.2 Fluid Shear Stress for Wake Interference Flow (case S400_H50_L5)

As in the skimming flow example, Reynolds stress for bar spacing L = 5 only
deviates from the linear profile close to the roughness crests because form-induced
stress becomes significant (Figure 5.20). Accordingly the form-induced stress is
negligible above the roughness layer.

The subplots in Figure 5.21 follow the same order as those in Figure 5.19, and
provide information about the underlying variation in the form-induced velocity
field (in terms of u% and w% ) that generates significant form-induced stress in the
roughness layer for wake interference flows. As the form-induced stress distribution
is similar to the skimming flow example, it is not discussed further. It is sufficient to
note that because the red areas of u% w% once again dominate in size and strength over

PhD Thesis 131 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
the blue zones [Figure 5.21(c)] the positive contribution to form-induced stress
outweighs the negative in the interfacial sub-layer; the net form-induced stress
distribution is given in Figure 5.21(d).

(a) (b)

u w / u*2 u% w% / u*2

Figure 5.20: Fluid stress profiles for wake interference flow case S400_H50_L5. (a) shows

spatially-averaged Reynolds stress, u w (b) shows form-induced stress, uw


% % . Both

are normalised with bed shear stress, o = u* . The dashed line indicates the level of the
2

roughness crests.

10-3
(a) (c)

(b) (d)

uw
% % (N/m2)

Figure 5.21: The flow patterns underlying form-induced stress for wake interference flow case
S400_H50_L5. (a) the streamwise form-induced velocity component, u% , (m/s) (b) the bed-
normal form-induced velocity component, w% ,(m/s) (c) their product, u% w% , (m2/s2) and (d) form-

induced stress, uw
% % , (N/m2).

PhD Thesis 132 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
Further comparison of Figures 5.19(d) and 5.21(d) reveals that in common with
roughness spacing L = 3, spacing L = 5 displays a slight negative bulge in the
uw
% % distribution immediately above the roughness crests. This signifies an
association between upwards-moving fluid ( w% > 0 ) meeting fluid moving faster than
average in the streamwise direction ( u% > 0 ) just above the roughness crests.
Although the feature is barely discernable at this point, it becomes increasingly
dominant as the roughness pitch grows, such that there is a marked change in the
form-induced stress distribution for most isolated roughness flows. The following
example highlights this behaviour.

5.4.3 Fluid Shear Stress for Isolated Roughness Flow (case S400_H50_L15)

In this case, the Reynolds stress begins to deviate from the linear trend at 2k above
the channel bed [Figure 5.22(a)]. Recall that for such isolated roughness flows the
roughness layer was estimated to be considerably thicker when compared to flows
over denser bar arrangements, with noticeable spatial variation in fluid velocities up
to 5k from the bed (Section 5.2.1). It is not surprising therefore that the peak in
form-induced stress no longer occurs in the interfacial sub-layer, but has instead
shifted upwards into the form-induced sub-layer (in this example the peak of
uw
% % lies 1-2mm above the roughness tops). What is perhaps surprising is that
the shape of the form-induced stress profile has completely changed [Figure 5.22(b)]
as compared to denser bar configurations.

(a) (b)

u w / u*2 u% w% / u*2

Figure 5.22: Fluid stress profiles for isolated roughness flow case S400_H50_L15. (a) shows

spatially-averaged Reynolds stress, u w (b) shows form-induced stress, uw


% % . Both are

normalised with bed shear stress, o = u* . The dashed line indicates the level of the roughness
2

crests.
PhD Thesis 133 Chapter 5.0
L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
The form-induced stress distributions for all bar spacings and flow depths are shown
in Appendix C (Figures C.6 to C.8). Reference to these figures reveals that the
change from a net positive to a net negative profile is a gradual one as roughness
classification shifts from wake interference to isolated roughness. Nevertheless,
the change in behaviour from having a very slight bulge in uw
% % (as discussed for
L = 3 and L = 5) to a noticeable negative spike just above the roughness crests
occurs precisely at the transition to isolated roughness flow. In other words this
shape of form-induced profile first appears for spacing L = 7, but is conspicuously
absent for L 6 . In common with tighter bar pitches however, spacing L = 7 retains
a maxima in the form-induced stress distribution just above z = k. In general, as bar
density decreases (i.e. we move through the spectrum of isolated roughness flows
from L = 7 to L = 20) the positive peak of the form-induced stress profile gradually
diminishes until ultimately, at L 15 negative form-induced stress dominates (see
Figures C.6 to C.8).

10-3
(a) (c)

(b) (d)

uw
% % (N/m2)

Figure 5.23: The flow patterns underlying form-induced stress for isolated roughness flow case
S400_H50_L15. (a) the streamwise form-induced velocity component, u% , (m/s) (b) the bed-
normal form-induced velocity component, w% ,(m/s) (c) their product, u% w% , (m2/s2) and, (d)

form-induced stress, uw
% % , (N/m2).

PhD Thesis 134 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
There is therefore an apparent distinction between form-induced stress in wake
interference and isolated roughness flows (Figures 5.20(b) and 5.22(b),
respectively). The former type has positive form-induced stress with the maximum
immediately below the roughness tops, while the latter exhibits negative form-
induced stress with the maximum value occurring above the roughness crests. In
effect this means that the form-induced momentum transfer (relative to average
flow) is directed towards the bed for wake interference flows, whilst for isolated
roughness it is directed away from the bed. Kanda et al. (2004) report similar
behaviour from an LES study of flow over cube arrays. Interestingly, to balance this
momentum transfer, the double-averaged Reynolds stress falls short of the linear
trend (i.e. is smaller) in wake interference flows, but shows a positive deviation from
the linear distribution for wider bar spacings. In other words the form-induced and
turbulent momentum fluxes are entirely complementary, and together maintain the
necessary force balance with gravity (down to the level of the roughness tops).

Opposite directions of form-induced momentum flux in wake interference and


isolated roughness flows indicate substantial differences in the interaction between
near-bed and mean flow in these two cases. As discussed there is comparably little
interaction when the transverse bars are sufficiently close together ( L 6 ). When
this condition is met the cavity vortices will occupy most, or all, of the gap between
the roughness elements leaving an insufficient length for the flow to reattach to the
channel bed. The double-averaged flow field is therefore relatively unaffected above
the bars, while local velocities below the crests carry more momentum downwards
than upwards. (For example, note that the negative bed-normal form-induced
velocity regions in Figures 5.19(b) and 5.21(b) have larger magnitude, and are often
larger in size than the positive areas). Conversely, if there is ample space for flow
reattachment (as in isolated roughness flows) the mean flow is able to interact freely
with the near-bed flow. The reattached stream then encounters subsequent roughness
elements that act to constrict the flow. Streamwise velocities are therefore locally
increased above the obstacle ( u% > 0 ) whilst flow is forced up and over the bar
( w% > 0 ). In overcoming this obstruction momentum is transferred upwards, yielding
a sharp negative peak in form-induced stress. These patterns are evident in Figure
5.23 for isolated roughness case S400_H50_L15.

PhD Thesis 135 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
5.4.4 Form-Induced Stress Levels

As has been seen in the preceding sections, peak form-induced stress may be
positive or negative, and occur above or below the roughness crests. Positive peaks
in form-induced stress tend to occur below or at the level of the roughness crests. In
contrast, and with the exception of skimming flow case L = 2, sub-zero peaks occur
exclusively above the bar tops (Figure 5.24). Although z-levels for peak form-
induced stress cannot be reliably interpreted as demarcating the upper bound of the
roughness layer, the fact that they increase fairly steadily as bar spacing widens
echoes the trend for greater roughness layer thickness as the square bars become
increasingly separated.

Figure 5.24: Normalised z-level for the dominant peak in the form induced stress profile,
uw
% % ( z ) , versus bar spacing, L for all three flow depths. Red data points represent a positive

peak, blue a negative peak.

Figure 5.25: Normalised z-level at which form induced stress, uw


% % ( z ) , becomes negligible,

versus bar spacing, L, for all three flow depths.

PhD Thesis 136 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
The level at which form-induced stress becomes negligible is instead a better
candidate for obtaining the roughness layer thickness as bar spacing varies. These
levels are shown in Figure 5.25, and show a similar trend to that shown in Figure
5.24 in that the level for negligible form-induced stress steadily increases as L
increases. The values plotted (Figure 5.25) would again suggest that the roughness
layer extends up to 5k from the channel bed, as was deduced when considering the
spatial variation in velocity and stress distributions. However, at this stage no further
consideration will be given to quantifying roughness layer thickness. This is
because, as shall be seen in the next chapter, it is possible to have significant local
velocity variation even when the form-induced stress is negligible. Taking this into
account, the roughness layer thickness is re-evaluated towards the end of the
following chapter (Section 6.3.3).

Peak form-induced stress levels measured in this study were consistently in the
range 15-20% of the total measured (turbulent and form-induced) stress. This places
the current results well within the range of previous findings in the literature,
acknowledging that existing estimates vary widely (from form-induced stress being
negligible to being the same order of magnitude as turbulent stress). For example,
when reviewing previous literature in Chapter 2 (Section 2.4.4), it was noted that
early studies concluded the form-induced stress was at most a few percent of the
turbulent stress, and therefore negligible (Mulhearn, 1978; Raupach et al., 1980).
Following these estimates advances in experimental techniques facilitated better
velocity measurements close to and within boundary roughness. Hence form-
induced stress levels have been detected far more accurately. Raupach et al. (1986,
wind tunnel) gauged the ratio between the maximum form-induced and Reynolds
stresses at 5%, Poggi et al. (2004, wind tunnel) find it to be nearly 40%, and Lien
and Yee (2004, numerical simulation) consider the form-induced stress to be of the
same order of magnitude as the turbulent stress (they do however consider the
superficial rather than the intrinsic spatial average). All of the above estimates come
from considering (three-dimensional) vegetative or urban canopies in the
atmospheric boundary layer. In such three-dimensional canopies the air stream is
more likely to flow around, rather than up and over the obstacles, and the main
contribution to form-induced stress will be uv
% % . Conversely, for transverse bar

PhD Thesis 137 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
roughness fluid is forced to flow up and over obstacles but is restricted from flowing
around in transverse planes, hence uw
% % is the dominant term.

As an endnote, natural sediment surfaces, e.g. gravel-bed rivers, are truly three-
dimensional such that fluid can flow up, over, under and around individual
obstacles. In simple terms they offer a combination of the flow paths from both
forest canopy and bar roughness geometries. Although the implication of such an
organic boundary roughness for the magnitude of form-induced stress may seem
uncertain, PIV measurements over an impermeable granular sediment bed
(d50 2mm) showed peak form-induced stress levels ( uw
% % ) to be 15-35% of the
recorded turbulent stresses in the roughness layer (Campbell et al. 2005). Hence,
whilst intuitively one may consider that two-dimensional bar roughness or cylinder
roughness to be very different and capable of generating far greater disturbance in
the time-averaged flow as compared to natural channel beds, this is clearly not the
case.

5.5 Chapter Closure

The first section of this chapter was devoted to demonstrating the need for double-
averaging the momentum equations when applied to flows over rough beds. In doing
so the extent of the roughness layer, i.e. the zone hydrodynamically influenced by
the surface roughness, was estimated to be up to five roughness heights from the
channel bed.

Section 5.3 then illustrated the results arising from implementing double averaging
methodology to the flow in the roughness layer. The key discovery was the
existence of a linear double-averaged velocity distribution in the gap between square
bars classified as isolated roughness. The reason for the gradual appearance of a
linear profile was explained in terms of the balance between separated and
reattached flow areas in the interfacial layer.

To examine the nature of fluid stress, Section 5.4 presented both turbulent and form-
induced stress profiles from a range of roughness spacings. Whilst form-induced

PhD Thesis 138 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
stress completely changed sign from positive to negative with increasing roughness
pitch, it always remained complementary to the Reynolds stress. As stated, the
roughness layer extended up to z = 5k, but the peak in the form-induced stress
profiles was always restricted to the level of the roughness crests (typically up to z =
2k). Therefore it is possible to have marked local variation in velocity throughout the
roughness layer ( u% , w% 0 ), but the spatial average of the product of form-induced
components, form-induced stress is negligible. The detailed variation in the time-
averaged flow in relation to the bed roughness is discussed further in the next
chapter.

It is clear that double averaging has the potential to help us ascertain the general
representative characteristics of flow close to rough boundaries. Further
investigation of such general rules will undoubtedly have applications for
understanding surface friction and numerical modelling. However, observing
double-averaged quantities in isolation can lead to the masking of the precise details
of the fundamental mechanics governing the detailed interaction between fluid flow
and a rough bed. The following chapter, Chapter 6, therefore engages with the
heterogeneity of flow characteristics in the roughness layer, by considering the
difference between time- and double-averaged flow distributions.

PhD Thesis 139 Chapter 5.0


L J Campbell Double-Averaged Flow Characteristics in the Roughness Layer
6.0 LOCAL FLOW
CHARACTERISTICS IN THE
ROUGHNESS LAYER

6.1 Introduction

Having ascertained the general characteristics of the near-bed region in open channel
flows over square bars (in terms of double-averaged velocity and shear stress
profiles), this chapter now examines the local flow patterns in the vicinity of the
roughness. This is achieved by subtracting the double averaged distribution from the
temporally averaged flow field. Doing so reveals the spatial variation in the time-
averaged flow, and makes it possible to chart the form-induced velocity components,
u% and w% . The chapter culminates in a new application for quadrant analysis
(Quadrant Mapping) based on plotting the form-induced velocity fluctuations ( u%
and w% ) in place of the turbulent fluctuations ( u and w ). Hence, the contents of
Chapter 6 are,

Section 6.2: A look at the local variation in fluid velocities close to the
square-bar-roughened beds, or, Local Fluid Motion in the Roughness
Layer.

Section 6.3: Visualising the form-induced disturbances: Quadrant


Mapping.

6.2 Local Fluid Motion in the Roughness Layer

6.2.1 A Preliminary Research Result

In order to provide some background as to why such a comprehensive range of


square bar spacings was investigated during this project, Figure 6.1 presents results
from a preliminary period of testing using the autocorrelation PIV system. It shows
the streamwise variation in form-induced momentum flux ( uw
% % ) in four different

PhD Thesis 140 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
levels over square bar roughness (L = 8, isolated roughness flow). The periodic
pattern, with 2 wavelengths per bar pitch is clear, and the amplitude of oscillations
gradually decreases as the distance from the bed increases. Whilst this pattern
appeared to be stable for bar spacing L = 8, its form posed many questions. For
example, how would the pattern alter when the gap between square bars had
insufficient length to support such periodic oscillations?

Figure 6.1: Local form-


induced momentum flux,

u% w% / u* , in planes 1.6-2.4 k
2

above the channel bed (where


k=6.35mm). This preliminary
result (from autocorrelation
PIV) is for spacing L=8 and
flow depth H=95mm, and
prompted the thorough and
systematic variation of square
bar pitches from L=2 to L=20.

6.2.2 The Form-Induced Velocity Components close to Square Bar Roughness

Explanation of the periodic pattern shown in Figure 6.1 requires an understanding of


the underlying spatial heterogeneity in the time-averaged flow close to a rough
boundary. This is most appropriately visualised via the form-induced velocity
components as they are the constituent parts of the form-induced momentum flux
( uw
% % ), and can highlight areas of differing momentum transfer (relative to average
flow).

Figures 6.2 to 6.5 show the streamwise variation of the form-induced velocity
components for a representative subset of bar spacings, L = 3, 5, 7 and 15. Data for
all plots come from a range of bed-parallel planes through the roughness layer;
Figures 6.2 (showing u% ) and 6.3 ( w% ) cover levels over the range z = 0.7 1.3k (i.e.
within, and just above, the interfacial sub-layer), whilst Figures 6.4 ( u% ) and 6.5 ( w% )

PhD Thesis 141 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
show behaviour higher up in the form-induced sub-layer ( z = 1.6 2.0k ). These
figures are discussed in turn.

The first point to note from case L = 3 is that the patterns of form-induced velocity
over successive bar pitches coincide extremely well, an indication that the flow in
the flume was fully developed [Figures 6.2(a) and 6.3(a)]. The general trend for u% in
the interfacial sub-layer is the same for all cases, and reflects areas of low
streamwise momentum in the separated flow immediately downstream and upstream
of each bar, where u% <0 (Figure 6.2). In these areas the behaviour of w% is also
governed by recirculating flow, such that w% >0 on the downstream face of each bar
(Figure 6.3) in the upwards portion of the separation vortex. w% may however be
positive or negative at the upstream face of each bar, largely depending on bar
spacing. For example, in case L = 3, w% dips negative at levels z/k = 0.7 and 1.0 (and
even slightly at z/k = 1.3) as the vortex cavity extends throughout the interfacial sub-
layer. At L = 5 and 7 only the lowest plotting plane, z/k = 0.7, shows w% <0 on the
upstream face of the roughness, whilst at the widest spacing shown, L = 15, w% >0 for
all planes z/k = 0.7, 1.0 and 1.3. The extent of the corner eddy formed ahead of each
bar diminishes with increasing bar spacing (Figure 6.6; also refer to Appendix D for
streamlines over a range of bar spacings to chart the evolution of different time-
averaged vortex patterns). Hence, fluid in the interfacial sub-layer over wider bar
pitches is forced up and over the obstacles, rather than becoming entrapped in stable
vortices.

It is possible to draw a further distinction in the pattern of w% in the interfacial sub-


layer, this time either side of flow reattachment. Bar spacing L = 5 generates wake
interference flow according to the classification of Morris (1955, Table 4.2), whereas
L = 7 yields isolated roughness flow. The lowest plotting plane, z/k = 0.7, in Figures
6.3 (b) and (c) show that over isolated roughness fluid can begin to move upwards
some distance ahead of the next roughness element, but is not free to do so in wake
interference flows. The quasi-sinusoidal form-induced stress patterns over isolated
roughness shown in Figure 6.1 are therefore governed primarily by the behaviour of
the vertical form-induced velocity component, which is in turn controlled by flow
separation and reattachment.

PhD Thesis 142 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
(a) (a)

(b) (b)

(c) (c)

(d) (d)

Figure 6.2: Streamwise variation (largely in the interfacial sub-layer) of Figure 6.3: Streamwise variation (largely in the interfacial sub-layer) of the
the streamwise form-induced velocity component, u% (recall u% = u u ), % (recall w% = w w ), in
bed-normal form-induced velocity component, w
in bed-parallel planes z=0.7, 1.0, and 1.3 k. Bar spacings (a) L=3; (b) bed-parallel planes z=0.7, 1.0, and 1.3 k. Bar spacings (a) L=3; (b) L=5; (c)
L=5; (c) L=7, and (d) L=15. Note: different plotting scales are used. L=7, and (d) L=15. Note: different plotting scales are used.
PhD Thesis 143Chapter 6.0
L J Campbell Local Flow Characteristics in the Roughness Layer
(a) (a)

(b) (b)

(c) (c)

(d) (d)

Figure 6.4: Streamwise variation (in the form-induced sub-layer) of the Figure 6.5: Streamwise variation (in the form-induced sub-layer) of the
streamwise form-induced velocity component, u% (recall u% = u u ), in % (recall w% = w w ),
bed-normal form-induced velocity component, w
bed-parallel planes z=1.6, 1.8, and 2.0 k. Bar spacings (a) L=3; (b) L=5; in bed-parallel planes z=1.6, 1.8, and 2.0 k. Bar spacings (a) L=3; (b)
(c) L=7, and (d) L=15. Note: different plotting scales are used. L=5; (c) L=7, and (d) L=15. Note: different plotting scales are used.
PhD Thesis 144Chapter 6.0
L J Campbell Local Flow Characteristics in the Roughness Layer
Before considering flow above the roughness crests, it is worth noting that in the
interfacial sub-layer the magnitude of the streamwise form-induced velocity
component is similar in all cases (Figure 6.2). For the bed-normal component
however, cases L = 3 and 15 have relatively high values (Figures 6.3 (a) and (d),
respectively). The former is typified by a strong recirculating cavity vortex and the
maximum lies in the plane plotted below the roughness crests, whilst the latter has a
localised maximum in the plane just above the crests when reattached flow
encounters the subsequent bar. The position of maximum disturbance in the time-
averaged flow (i.e. above or below crests) is entirely consistent with the observation
of the locations of peak form-induced stress in Section 5.4 for different roughness
classes.

(a)

(b)

Figure 6.6: Time-averaged streamlines for bar spacings (a) L=7 and (b) L=15 (H=50mm)
showing the diminished extent of the corner vortex (in L=15) in the angle between the solid bed
and the upstream face of the bar roughness. Subplots not drawn to same scale.

Moving beyond the immediate vicinity of the roughness into the upper zone of the
roughness layer, the form-induced sub-layer, Figures 6.4 and 6.5 show the variation
in u% and w% respectively at levels z/k = 1.6, 1.8 and 2.0 (and with data from the same

PhD Thesis 145 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
experiments as Figures 6.2 and 6.3). The distinction between different flow types
now becomes more apparent. It was discussed in Chapter 5 that there is negligible
disturbance above the roughness crests for skimming (quasi-smooth) flow types,
and little disturbance for wake interference flows. As such, the form-induced
velocity components for L = 3 and 5 are considerably smaller than those for isolated
roughness flows, and only approach 5% of the peak magnitudes in spacing L = 15
(Figures 6.4 and 6.5). The concentrated dips in u% and w% have disappeared as the
plotting planes are no longer intersected by the square bars. However there is a local
positive peak in u% as flow is forced over more widely spaced bars. Because the flow
is able to reattach to the solid bed in cases L = 7 and 15 the next bar represents a
substantial flow constriction (relative submergence, H/k = 7.9 for the examples
shown), so the streamwise velocity is enhanced [ u% >0, Figures 6.4(c) and (d)] whilst
the flow moves upwards [ w% >0, Figures 6.5(c) and(d)]. This gives the local product
% % >0 which is large enough to dominate the volume averaged value, i.e. form-
uw
induced momemtum flux, uw
% % <0 in the form-induced sub-layer, as seen in Section
5.4.3.

Aside from differences in magnitudes and patterns of the form-induced velocity


components in the upper part of the roughness layer for different bar spacings, the
most obvious feature of Figures 6.4 and 6.5 is the systematically decreasing effect of
the bed roughness. The same pattern for both u% and w% is continued through the flow
depth, but the amplitude of the periodic swings decreases as distance from the bed
increases. Furthermore, although in isolated roughness flows u% and w% each undergo
one full cycle per bar pitch (Figures 6.4 and 6.5), there is a systematic, roughness-
induced phase shift between the components. Hence the form-induced stress
completes 2 cycles per pitch as shown in Figure 6.1. This is discussed below.

6.2.2.1 Periodic Behaviour of the Form-Induced Velocity Components

Case S400_H50_L8 is representative of isolated roughness flows and has the same
bar spacing as the preliminary test case of Figure 6.1. As such it will be used to
further illustrate the periodic behaviour of the form-induced velocity components, u%
and w% , and the form-induced momentum flux uw
% % . The streamwise variation of u%

PhD Thesis 146 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
and w% at the level 2k from the solid bed is shown in Figure 6.7. Both components
complete one whole wavelength over each whole roughness pitch. As in the
isolated roughness examples above (Figures 6.4 and 6.5) streamwise motion is faster
than average both above, and for some distance downstream of the square bar, i.e.
u% >0. In this example streamwise velocity drops below average in the middle of the
interbar gap ( x / P 0.5 , where P is the roughness pitch as defined in Figure 4.13).
Over the same distance the direction of average bed-normal momentum switches
from positive to negative as the flow naturally tends towards the reattachment point
(Figure 6.7). Behaviour in the second half of the roughness pitch, 0.5 x / P 1.0 ,
mirrors the first half, but with w% >0 (flow moving up to overcome the next bar) and
u% switching from negative to positive (flow compensates for the constriction over
the next bar). Each form-induced velocity component therefore undergoes the same
positive-negative-positive sequence, albeit phase shifted relative to one another.

Figure 6.7: Streamwise

variation of u% and w
% (m/s)
for case S400_H50_L8,
showing that each component
undergoes one full cycle over
one whole roughness pitch, P.
Profiles for u% and w
% are
phase shifted by .

Cross-correlations of the distributions of u% and w% at all spatial lags make it possible


to quantify the degree of phase shift, , between the two components1. In other
words by shifting the position of the two profiles (for example, those shown in

{(u% (i ) mu% ) ( w% (i + d ) mw% )}


1
Where (circular) cross-correlation, r = i
; d is the lag or
(u% (i ) mu% ) 2 ( w% (i + d ) mw% ) 2
i i

spatial shift between the two profiles, and m refers to the mean value. Vectors u% and w % are
considered over the spatial interval of one roughness pitch, P, which is assumed to be a common
period of the two corresponded periodic signals. Vector w % is shifted circularly by d positions.
Strictly, mu% = mw% = 0 in the full function formula given above. This function is not to be confused
with cross-correlation PIV.

PhD Thesis 147 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
Figure 6.7) relative to each other and then calculating the correlation between the
profiles, the peak correlation gives the spatial shift between u% and w% . The result of
this exercise reveals that = 0.25P, i.e. u% and w% are at equivalent points of their
cycles a quarter of a roughness pitch apart (Figure 6.8). Perhaps more interestingly,
as the four plotting planes in Figure 6.8 show, this lag appears to remain constant
through the flow depth; w% always mimics u% , separated by one quarter of a
roughness pitch.

Figure 6.8: Cross-correlation plots


for 4 planes above L=8 rough bed
(H=50mm). These show the excellent
correlation, r, between u% and w
% at
various spatial shifts (refer to footnote
1 for details). The peaks for all planes
are clustered around x=0.25P, as
indicated by the shaded zone, thus
illustrating that the periodic forms of
u% and w% are shifted in space by a
quarter of one whole roughness pitch,
P. In this isolated roughness example,
this shift appears relatively insensitive
to vertical position over the range
2 z/k 5.

To further illustrate how these periodic variations in u% and w% , and the quarter-pitch
shift between them, affect the form-induced momentum flux one may approximate
the components as:

u% / u* = au cos(2 x / P ) (6.1a)

w% / u* = aw sin(2 x / P ) (6.1b)

The wavelength in Equation 6.1 corresponds to the length of a full orbit, and is
clearly equal to the roughness pitch, P. According to (6.1), over this length each

PhD Thesis 148 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
form-induced velocity component completes a full orbit, while their product, the
% % / u*2 = (au aw / 2)sin[2 x /( P / 2)] , will complete
form-induced momentum flux, uw

two orbits per pitch. In reality the amplitudes au and aw in Equation 6.1 are
dependent on z-level as they decrease when the influence of the roughness
diminishes through the flow depth. If required, this decrease may be simulated via a
simple relation of the form au = aw = ao exp(-cz/k), noted here solely for illustration
purposes. Somewhat less intuitively, au and aw are also dependent on streamwise
coordinate, x. This is because the upstream and downstream faces of each roughness
element exert a variable influence on the flow (to be discussed later in this chapter).
Additionally, the flow is extremely sensitive to small variations in roughness pitch
and height, as is evident in the discrepancy in peak form-induced momemtum flux
between successive pitches (Figure 6.1). A constant value for au and aw may
however be adopted for simplicity, and can be considered to be the mean amplitude
sampled over a number of roughness pitches.

Figure 6.9 compares the form of the analytical approximation of Equation 6.1 to real
data (that used for Figure 6.7), and shows that the simulated results capture the
general trends in u% , w% and uw
% % very well. Although the patterns in Figure 6.9 are
undoubtedly the consequence of flow over two-dimensional, fixed-period roughness,
they are most probably not geometry-specific. Furthermore, although the test case
used throughout this section is S400_H50_L8, a similar pattern is evident for all
isolated roughness cases (see Appendix D for equivalent figures for spacings L = 7,
10, 12, 15, and 20).

Although a similar pattern to that shown in Figure 6.9 persists for all isolated
roughness cases, reference to Appendix D (Figures D.12 to D.17) shows that the
extremes of form-induced velocity components become increasingly separated with
increasing bar pitch. This is to be expected, as these extremes are associated with
flow behaviour in close proximity to the roughness elements, and not with the zone
of reattached flow between the bars. As a direct consequence the peak and trough of
the cross-correlation function also become increasingly separated at wider bar
pitches.

PhD Thesis 149 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
a u cos(2 x / p )
a w sin(2 x / p )

SIMULATION [Eq.(6.1)]
2
uw
%% / u
*

(a) (b)

u% / u*
w% / u*

DATA [S400_H50_L8]
2
uw
% % / u*

(c) (d)

Figure 6.9: Form-induced velocities and momentum flux: comparison of suggested simulation
as given in Eq. (6.1) with real data from case S400_H50_L8. The plane considered is 2k from
the solid bed, streamwise coordinate is normalised with roughness pitch, P. (a) Variation in
% and uw
simulated values for u% , w % % , (b) cross-correlation of u% and w% versus spatial lag for
% and uw
simulated results, (c) Real variation in u% , w % % , and (d) cross-correlation of u% and
w% versus spatial lag for data.

In this section the characteristic variation in the time-averaged flow in the roughness
layer has been examined (by subtracting double-averaged velocity profiles). In all
figures the variation has been plotted against streamwise position. Attention now
turns to what the form-induced velocity components look like when plotted on
orthogonal axes, i.e. by applying a method analogous to quadrant analysis. To
differentiate from conventional quadrant analysis as applied to temporal velocity
fluctuations, application of the technique to the spatial disturbances (form-induced
velocities) will henceforth be referred to as quadrant mapping.

PhD Thesis 150 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
6.3 Quadrant Mapping

Quadrant mapping involves plotting the form-induced velocity components, u% and


w% , on orthogonal axes to gain further insight into the way fluid flow interacts with
the rough bed. It is important to recall that the form-induced velocity components are
calculated from time averaged velocities. Therefore, rather than indicating the
presence of short-lived turbulent events as in conventional quadrant analysis,
different quadrants in quadrant mapping relate to patterns of local deviations in time-
averaged velocities from the double average. As such, they may be interpreted as
roughness-induced structures or perturbations of the time averaged flow.

(a) w (b)
Q2 Q1
Outward
Ejection Interaction

u
Q3 Q4
Inward
Interaction Sweep

Figure 6.10: Quadrant diagrams. (a) Quadrant numbers and associated turbulent events,
(b) Quadrant diagram from experiment S400_H50_L5 generated from velocity fluctuations a point
located above the inter-bar gap at level z=2k. The clusters of data points in quadrants 2 and 4
indicate significant contributions to Reynolds stress.

6.3.1 Conventional Quadrant Analysis

Conventional quadrant analysis was introduced by Lu and Willmarth (1973) to study


the relationship between temporal fluctuations of orthogonal velocity components,
u and w , in particular their distribution between four quadrants as labelled in
Figure 6.10(a). Quadrant diagrams are of great interest as they provide a useful
means of understanding and visualising the turbulent structure of boundary layer
flows such that the burst-sweep cycle and primary Reynolds stress are both indicated

PhD Thesis 151 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
by the negative correlation of the velocity fluctuations. (Indeed, conventional
quadrant diagrams are utilised and presented in Chapter 7 to help explain the
transition phenomenon). Poor correlation between u and w is interpreted as an
absence of turbulent coherent structures, significant correlation between u < 0 and
w > 0 indicates ejection-like events (Quadrant 2), and good correlation between
u > 0 and w < 0 indicates the presence of sweep-like events (Quadrant 4) (Lu and
Willmarth 1973, Nezu and Nakagawa 1993). A typical quadrant diagram generated
from velocity data at a point in the turbulent flow above bar spacing L = 5 is shown
in Figure 6.10(b).

6.3.2 Spatial Quadrant Analysis: Quadrant Mapping

In this study a similar technique to that described above has been applied to the
spatial disturbances of the time-averaged velocities, i.e. the form-induced
components u~ and w~ . In this context a quadrant diagram shows time-averaged data

from the whole length of the bed-parallel averaging volume, rather than from a
single point through time (as for u and w ). The interpretation of each mapping
quadrant, including a colour code to enable cross-matching of specific flow regions
with quadrants 1 to 4, is shown in Figure 6.11 (Q1-green, Q2-blue, Q3-yellow, Q4-
red, Pokrajac et al., 2005). By plotting data from all z-levels this method allows one
to build a full picture, or map, of the recorded flow section in order to visualise the
spatial variation of local flow characteristics. Henceforth these plots are called
quadrant maps.

In conclusion to Chapter 5, it was stated that it is possible to have significant


variation in local time-averaged fluid velocities even in areas where the form-
induced stress is negligible. In other words the roughness layer comprises two
distinct zones; fluid flow in both zones is influenced by the presence of the bed
roughness, but whereas the lower zone exhibits significant form-induced stress, the
upper one does not. This distinction produces two quite different quadrant diagram
patterns. For reasons that will shortly become clear, the upper zone (with noticeable
form-induced velocity variation, but negligible form-induced stress) is called the
Zone of the Rings and the lower zone (with significant form-induced velocities and

PhD Thesis 152 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
stress) is referred to as the Zone of Irregular Orbits. Quadrant diagrams from each
zone are presented alongside quadrant maps in the following two subsections (in all
cases data come from the form-induced sub-layer, and not from the interfacial
portion of the roughness layer. This avoids areas of recirculation, which complicate
the interpretation of flow in each quadrant as described in the legend for Figure
6.11).

w%

Q2 Q1
Upwards ( w ) Upwards ( w )

Slower than Faster than


average ( u ) average ( u )

u%
Q3 Q4
Downwards ( w ) Downwards ( w )

Slower than Faster than


average ( u ) average ( u )

Figure 6.11: Interpretation of each Quadrant Mapping sector in terms of local (time-averaged)
fluid motion relative to average motion. Note that this account of each quadrant is correct for
all but the small areas of upstream fluid motion in recirculation zones, where an alternative
description may apply. For example, if at a point both u < 0 and u < 0 , u% ( = u u ) will
be positive when local streamwise velocity is slower than average, albeit in the upstream or
negative x direction.

6.3.2.1 The Zone of the Rings

The zone of the rings exists for all isolated roughness cases (L = 7 to L = 20) and lies
immediately above the level of significant form-induced stress (which is normally
one roughness height from the roughness crests at most, Section 5.4). Figure 6.12
shows typical quadrant diagrams from four levels in the range 1.5k z 3.0k
[6.12(a)] alongside the full quadrant map showing which flow areas correspond to

PhD Thesis 153 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
which quadrants [6.12(b)]. For completeness the form-induced stress distribution
makes up the third figure panel [6.12(c)]. Reference to Figure 6.12 reveals that form-
induced stress decreases rapidly and over shorter distances from the roughness crests
compared to the form-induced velocity fluctuations. Indeed, the spatial variability of
velocity is still significant at the height where the form-induced stress is already
negligible. Case S400_H50_L8 is used here, but additional plots from experiments
with different bar spacings are given in Appendix D.

(a) (b) (c)

Figure 6.12: Form-induced velocity behaviour in the zone of the rings for experiment
S400_H50_L8, colour code as given in Figure 6.11. (a) Quadrant Diagram charting the spatial
variation in the time-averaged flow for four levels, z/k=1.5, 2.0, 2.5, and 3.0. Each of these plots
make a ring-like pattern. (b) Quadrant Map showing which areas of flow correspond to which
quadrant, black dots illustrate plotting planes in subplot (a). (c) Form-induced stress distribution.

All velocities are normalised using friction velocity, u* .

The four quadrant diagrams plotted [Figure 6.12(a)] have a distinct ring-like,
pseudo-circular shape, in which the four quadrants are equally present. The rings do
not necessarily have to close precisely at each end of the averaging window (i.e. in
Quadrant 1, above the bar crest), and the fact that they do not in Figure 6.12(a)
probably signifies small discrepancies in bar height and position, and hence flow
field, in the streamwise direction. However, as the quadrant map [Figure 6.12(b)]
shows, the areas corresponding to each quadrant remain remarkably regular in the
bed-normal direction with increasing distance from the bed, indicating that the flow
patterns induced by the square bars persist well into the flow. This also dictates that
the ring shape must be similarly consistent throughout the height of the form-
induced sub-layer. Interestingly, the regular striped pattern only begins to break up

PhD Thesis 154 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
around z = 6k, which is above the conventionally accepted upper bound of the
roughness layer (5k, e.g. Raupach et al. 1991), and even above the levels suggested
in Chapter 5.

Upstream of each bar the longitudinal velocity decreases (relative to the spatial
average) and the vertical velocity increases, i.e. the spatial fluctuations of u~ (the
~ (the bed-normal component)
streamwise component) are negative whilst those of w
are positive. This combination corresponds to Quadrant 2 (blue, Figures 6.11 and
6.12). Following on from this position, the streamwise velocity increases over the
bar to allow flow up ( w% > 0 ) and over the obstacle, so the velocity fluctuation u~
becomes positive, and the quadrant diagram enters Quadrant 1 (green). Then
downstream of the bar streamwise velocity remains faster than average as it moves
over the region of separated flow, but w ~ has now become negative as the

reattachment point is approached, giving a series of quadrant diagram points in


Quadrant 4 (red). Finally, while the vertical velocity disturbance is still negative,
flow begins to decelerate as it adjusts to the full flow height between the roughness
elements and the quadrant diagram passes through Quadrant 3 (yellow). As noted,
this pattern is typical across the whole of the form-induced sub-layer (and beyond)
supporting the view that the flow structure in the roughness layer is strongly affected
by the geometry of the rough bed (Yaglom, 1979; Raupach et al., 1991).

Figure 6.13: Contours of time-averaged velocity showing the recirculation zones upstream and
downstream of the square bar in experiment S400_H50_L8. Together, these zones act to provide a
more aerodynamic profile, as approximated by the superimposed dashed line, to the otherwise
angular geometry of the bar.

PhD Thesis 155 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
The regular shape of both the quadrant diagrams and maps is undoubtedly due to the
regular spacing of the square bars throughout the flume, and is probably also
strongly affected by the roughness shape and size. For example, magnitudes of the
vertical form-induced velocity are comparable (and sometimes larger) than those of
the horizontal disturbances. This is a consequence of the rough bed geometry having
two-dimensional roughness elements, whose height is relatively large compared to
the spacing and water depth, extending across the whole width of the flume. Thus the
roughness geometry forces vertical form-induced velocity to change over a
significant range, similar to, or even greater than that of horizontal velocity. In
addition to being strongly influenced by the solid shape of the rib roughness, the
observed flow patterns are a product of the recirculation areas upstream and
downstream of each bar (Figure 6.13). The bubbles impose an aerodynamic shape to
the otherwise angular bar, and in doing so allow a smooth transition of flow between
decelerating and accelerating regions, i.e. transition through all four quadrants.
Beginning upstream of a roughness element, this transition is always in the same
clockwise order, Q2-Q1-Q4-Q3, and is typical for isolated roughness elements which
allow adequate space for reattachment between bars. Under isolated roughness
conditions the temporally-averaged flow is therefore locally disturbed in a consistent
manner throughout much of the flow depth by individual roughness elements.

The regular pattern evident in Figure 6.12 extends up to at least z = 6k from the solid
bed, it should however be noted that the form-induced velocities are extremely small
at this level ( 1mm/s). Figure 6.14 shows how rapidly the magnitude of the form-
induced velocities, and hence the radius of the rings, diminishes with distance from
the channel bed (note that colours in this figure are not used to link flow areas to
specific quadrants as in Figure 6.12, but instead show the level that each plotted ring
comes from). The size of the rings increases towards the bed as the influence of the
roughness elements on the spatial variability of the time-averaged flow grows. As
the observation plane moves away from the bed the quadrant rings shrink and
become very small [Figure 6.14(a)]. Coincidentally, this is opposite to how
conventional quadrant diagrams of instantaneous velocity fluctuations behave within
the roughness layer (as they increase in magnitude with increasing distance from the
bed).

PhD Thesis 156 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
(a)

(b)

Figure 6.14: Quadrant diagram showing rings from throughout the form-induced sub-layer (for
experiment S400_H50_L8). Colours in this figure are not used to link flow areas to specific
quadrants, but to indicate the level that each plotted ring comes from. (a) Quadrant diagrams
showing how rings collapse towards the origin as distance from the channel bed increases. (b)
Levels of plotted rings, starting closest to the bed these are: z/k=2.0, 2.5, 3.0, 4.0, 5.0 and 6.0.

The position vector (from the origin), r , for any point on the quadrant ring diagram
gives information about the total form-induced velocity; the magnitude shows the
magnitude of the total form-induced velocity, whilst the angle shows its direction
(Figure 6.15). Ring sectors in the lower two quadrants (Q3 and Q4) may have a
smaller radius (or, in other words, the position vector of each data point is shorter)
compared to the upper quadrants (Figures 6.12 and 6.14). This effect can be
explained by further reference to the shape of the recirculation bubble (Figure 6.13).
In the region of flow around this aerodynamic profile local fluid velocities are
forced to change more abruptly as they approach the steeper upstream face (ring
sections in Q2 and Q1, Figure 6.12) when compared to flow over the gentler slope of

PhD Thesis 157 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
the downstream portion (rings sections in Q4 and Q3). The upstream recirculation
region therefore causes greater local disturbance to the time-averaged flow. Thus, the
direct effect of the recirculation zones is felt up to 2k from the bed, as is evidenced
by the imbalance in the quadrant rings shown in Figures 6.12 and 6.14. Beyond this
level ( z > 2k ) the effect dissipates and quadrant rings tend to be well balanced.

w%
r = u% i + w% j

u% = r cos
r
w% w% = r sin

u%
u%
Figure 6.15: The meaning of the position vector, r , for data points plotted on quadrant diagrams.
i and j are unit vectors in the horizontal and vertical directions respectively.

The fact that certain quadrant rings shown in Figures 6.12 and 6.14 (especially those
from planes closer to the bed as noted above) have a larger upper radius and a lesser
point density compared to the lower quadrants, does not necessarily produce
considerable form-induced momentum flux. The form-induced momentum flux is
essentially the spatial average of all products, uw
% % , such that the contributions from
Q1 and Q3 are negative, but from Q2 and Q4 uw
%% > 0. In terms of the total
spatial average (through all four sectors) the imbalance in the quadrant orbit is
unimportant so long as there is symmetry about either quadrant diagram axis,
including symmetry of data point density. For example, in the planes located at z =
1.5k and z = 2k (Figure 6.12), inputs from Q1 and Q2 cancel each other out, as do
those from Q3 and Q4 because the ring exhibits reasonable symmetry about the
vertical axis of the quadrant diagram. Furthermore, as discussed, when the
observation plane moves further from the bed the rings become symmetrical about

PhD Thesis 158 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
both axes (for example, at z / k 2.5 , Figures 6.12, 6.14). The net result in either
case is negligible form-induced stress as the contributions from all four quadrant
sections add up to approximately zero. The form-induced velocity can however still
be significant. This true for all levels characterised by ring quadrant diagrams.

As the observation plane moves towards the channel bed, form-induced stress is
certainly not negligible [Figure 6.12(c)], and as the name would suggest the shape of
the quadrant diagrams changes accordingly in the Zone of Irregular Orbits.

6.3.2.2 The Zone of Irregular Orbits

For isolated roughness flows, the form-induced stress distribution typically has a
positive maximum just below the roughness crests, and a negative peak immediately
above the bar crests. Clearly in these regions, quadrant rings cannot exist as they
would indicate negligible form-induced stress. Instead, the orbit through all four
quadrants becomes distorted, and contributions from Q1, Q2, Q3 and Q4 no longer
act to cancel each other out. Below the tops of the bars, a positive peak in uw
% %
would suggest that contributions from quadrants 2 and/or 4 exert most leverage on
the form-induced stress balance. Conversely, just above the crests, quadrants 1
and/or 3 must dominate to produce negative form-induced stress. Using data from
the same experiment as in the previous section (S400_H50_L8), Figures 6.16 and
6.17 illustrate the quadrant behaviour in planes below and above the roughness tops,
respectively.

Below the Crests: Irregular Orbits in the Interfacial Sub-layer


Quadrant pathways below the roughness tops may be best described as a horizontal
figure of eight, though with markedly different sized loops [Figure 6.16(a)].
Beginning from the downstream face of a square bar, the path first completes the
smaller of the two loops in the figure of eight. This is a full clockwise rotation in
quadrant 2 and represents various degrees of slower than average, upwards motion
(as one would expect at this level in the stable recirculation zone). The route then
passes very briefly through quadrant 3 (downwards deceleration), before looping
across quadrants 4 and 1 back to the starting point in quadrant 2, thus completing the

PhD Thesis 159 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
larger loop of the pathway [Figures 6.16(a) and (b)]. Along the pathway Q2-Q3-Q4-
Q1, the largest magnitudes of the position vector, r (Figure 6.15), are to be found in
% % are in the upper
Q1 and Q2. Similarly, the greatest absolute values of the product uw
two quadrants. Quadrant 2 points are not readily understandable, as they come from
within the recirculation bubble, but the points in Q1 reflect significant fluid
displacement over a relatively short distance (compared to the inter-bar gap) at the
upstream face of each square bar. The higher planes at z/k = 0.8 and 0.9 experience
the greatest disturbance as they are closest to the crest. The contributions from Q1
and Q2 to form-induced stress are reasonably equal and of opposite sign, hence they
largely compensate for each other.

(a) (b) (c)


Q2 Q1

Q3 Q4

(d)

Figure 6.16: Behaviour of the form-induced velocity components in three planes located below the
roughness crests for experiment S400_H50_L8. (a) Sketch showing a typical quadrant pathway, (b)
Quadrant diagram for three levels, (c) Quadrant map in the near bed region plotted using the
quadrant colour code given in Figure 6.11, and (d) Levels of selected planes, z=0.7, 0.8,and 0.9k.
Red diamond indicates starting x-position of quadrant pathways, as illustrated on subplot (a).

PhD Thesis 160 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
Below the crests, the major imbalance in the quadrant diagram arises between the
lower two quadrants, Q3 and Q4 [Figure 6.16(b)]. There are comparatively few
points in Q3, only 11% of the total number across the whole averaging window (at z
= 0.8k). In contrast, Q4 contains the vast majority of data points (44%), indicating
that for around half of the inter-bar gap the fluid in this plane is moving downwards
and faster than average in the streamwise direction. This is unsurprising when one
considers that dead zones of fluid in recirculation areas act to suppress the double-
averaged velocity value. Hence Q4 provides the dominant contribution to the form-
induced stress profile at this level, and is primarily responsible for the positive peak,
uw
% % > 0 , in the interfacial sub-layer.

(a)
Q2 Q1

Q3 Q4

(b)

Figure 6.17: Behaviour of the form-induced velocity components in four planes located above the
roughness crests for experiment S400_H50_L8. (a) Quadrant pathways. The pseudo-circular orbits
follow the same clockwise rotation order, Q2-Q1-Q4_Q3 as those in Figure 6.16. (b) Levels of
selected planes, z=1.1, 1.2, 1.3 and1.4k.

PhD Thesis 161 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
Above the Crests: Irregular Orbits in the Form-Induced Sub-layer
In common with the quadrant diagrams from the interfacial sub-layer (Figure 6.16),
data in the lowest plotted plane above the bar crests, z = 1.1k, traces a figure of eight
[the dark blue route through the quadrants in Figure 6.17(a)]. Its structure is however
greatly altered, with the different sized loops in Q2 and Q4 reversing locations
completely. This type of pathway is peculiar to plotting locations very close to the
roughness crests, and is complicated by flow conditions immediately adjacent to the
surface. For example, the substantial negative disturbances to the average
alongstream velocity in Q2 ( u% 2u* ) occur immediately above the bar top, within
approximately 0.5mm of the solid crest surface, resulting in comparatively slow
streamwise fluid motion ( u% < 0 ). In other areas (in higher and lower planes), such
low points of streamwise momentum are usually accompanied by vertical
disturbances of similar magnitude. However, because the observation plane is so
close to a bed-parallel roughness face (i.e. the bar crest), the local time-averaged
bed-normal velocity ( w( x, z ) ) must be very small directly above a bar. Also the
double-averaged bed-normal velocity is necessarily negligible ( w ( z ) 0 ). Hence,
the difference between the two averages, w% , is also negligible. Therefore, although
there is significant local disturbance to the time-averaged flow in the streamwise
direction it is not transported in the bed normal direction, so Q2 makes little
contribution to form-induced stress at level z = k. Of the remaining three quadrants,
despite the large vertical disturbance at the upstream bar face (in Q1), Q1 and Q4 are
reasonably balanced, and Q3 contains no data points. Consequently, very close to the
roughness tops, form-induced stress is approximately equal to zero, and although
distinctive, the quadrant diagram around z = k bears little relation to the planes above
(or below). Indeed, fluid behaviour in the overlying planes has a far greater bearing
on the magnitude of the primary form-induced stress (Figures 6.12(c) and 6.17).

The negative peak in form-induced stress occurring around z = 1.25k (for experiment
S400_H50_L8) must be generated by significant contributions from quadrants 1
and/or 3 at this level, as noted in the opening paragraph to this section (6.3.2.2). The
irregular orbits of the middle two observation planes in Figure 6.17 at levels z = 1.2k
and z = 1.3k reveal that both Q1 and Q3 contribute. The orbits bulge noticeably on
their paths through Q1 giving substantial values for the product of the two form-

PhD Thesis 162 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
induced velocity components. In general quadrant rings in the form-induced sub-
layer tend have a larger radius in the upper quadrants, Q1 and Q2, when compared to
Q3 and Q4. Earlier in this chapter this was attributed to the steep upstream face of
the recirculation bubble generated by each bar (Section 6.3.2.1, and with reference to
Figure 6.13). This explanation holds for the planes currently under discussion lower
in the form-induced sub-layer where the effect is felt even more strongly because of
their proximity to the rough bed (Figure 6.17). The local disturbances in Q3 are
much smaller than in Q1 but the data point density is high, matched only by Q4. Q3
and Q4 share 65% of all points from the averaging window, with Q1 and Q2
splitting the remainder equally. As the orbits are bigger in the odd quadrants, Q1
easily outweighs Q2 and Q3 adds more to form-induced stress balance than Q4.
Hence both Q1 and Q3 control the magnitude of form-induced stress at this level,
and are responsible for the negative peak, uw
% % < 0 , in the form-induced sub-layer.

Figure 6.18: Contributions from each of the four spatial quadrant analysis quadrants (Q1 to Q4)
to the total form-induced stress (shown as a solid black line). This helps to illustrate that Q1 and
Q3 make larger contributions than Q2 and Q4, respectively to create a negative peak in form-
induced stress in the interfacial sub-layer (z=1.25k). This pattern is typical for isolated roughness,
or k-type flows.

PhD Thesis 163 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
Figure 6.18 further clarifies the above discussion of the contribution from each
quadrant to the total form-induced stress. This shows the vertical distribution of the
stress contributions from Quadrants 1 to 4, together with the total form-induced
stress profile (for the same experiment as plotted in Figures 6.16 and 6.17).

(a) z/k=3.94 w% / u* (g) z/k=1.49

u% / u*

z/k=2.97 z/k=1.36
(b) (h)

(c) z/k=2.46 (i) z/k=1.29

z/k=1.94 z/k=1.23
(d) (j)

(e) z/k=1.49 (k) z/k=1.16

(f) z/k=1.10 (l) z/k=1.10

Figure 6.19: Evolution of the quadrant rings (experiment S400_H50_L8). (a) to (f) show orbits from
planes throughout the form-induced sub-layer. (g) to (l) focus on the range 1.10 z / k 1.49 (i.e.
the range between subplots (e) and (f). All form-induced velocity components are normalised with
friction velocity.

PhD Thesis 164 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
6.3.2.3 Evolution of the Quadrant Orbit

This short section ties together many of the above quadrant diagrams from both the
Zone of the Rings (6.3.2.1) and the Zone of Irregular Orbits (6.3.2.2). Further
reference to Figures 6.16, 6.17 and 6.12 (in this order they show quadrant diagram
planes at increasing distance from the channel bed) holds some clues as to the
evolution of the well balanced quadrant ring above z = 2k. For example, the orbits
plotted in Figure 6.17(a) grow more ring-like as the distance from the bed increases.
It is useful to separate these rings onto separate subplots to illustrate the transition
from the zone of irregular orbits to the zone of the rings. A detailed evolution of the
quadrant ring for case S400_H50_L8 is provided in Figure 6.19.

The figure above (6.19) shows that the quadrant rings evolve over a short distance
from the roughness tops and are firmly established by z = 1.3-1.4k. The key features
of the orbits in Figure 6.19 have been discussed above. For example, the upper
quadrants are commonly larger than the lower two, and remain so until the rings
become well balanced from around z = 2k [6.19(d)] upwards. The negative peak in
form-induced stress for the illustrated experiment occurs at z = 1.23k. The quadrant
diagram [6.19(j)] demonstrates that a significant bulge in Q1, together with high data
point density in Q3, typifies the orbit at this level. Beyond z/k = 1.23 however the
radius of the orbit path through Q1 is gradually matched as the other three quadrants
fill out to complete the evolution of the ring.

6.3.2.4 The Effect of Bar Spacing on Quadrant Maps

The demonstration and discussion of quadrant mapping has thus far concentrated on
isolated roughness flows, as it is the undulating flow pattern over areas of separation
and reattachment generates quadrant rings. Indeed, for skimming flows it is difficult
to discern any repeating pattern in the quadrant maps beyond the immediate vicinity
of the roughness crests. This is to be expected, as there is little alongstream variation
in the time-averaged flow beyond the interfacial cavities of such flows (as presented
in Chapter 5). As bar spacing increases however, there is an increasing trend for
more organised fluid motion in the roughness layer, such that it is possible to chart

PhD Thesis 165 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
the development of quadrant map stripes in wake-interference flow types, L = 4, 5, 6
(Figure 6.20).

(a) (b)

Figure 6.20: Quadrant maps for wake


interference cases (a) S400_H50_L4,
(b) S400_H50_L5, and (c)
S400_H50_L6. Maps are colour
coded according to Figure 6.11. This
sequence of quadrant maps (i.e. for
bar spacings L 7 ) continues in
Appendix D, Figures D.18 to D.23. (c)

The quadrant maps for L = 4, 5, and 6 illustrate that the patterns in the time-averaged
flow field above the bar tops become progressively more coherent from L = 4
onwards. Quadrant map stripes are clear in wake interference case L = 6, which leads
the bar spacing series into the first true isolated roughness case L = 7 (the quadrant
map for L = 7 is given in Appendix D). Developing quadrant stripes with increasing
bar spacing indicate that the mean flow is able to penetrate more and more into the
growing gap between square bars, until ultimately it reattaches to the channel bed in
L = 7.

PhD Thesis 166 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
For isolated roughness cases ( L 7 ) at flow depth H = 50mm, quadrant stripes fill
most or all of the flow depth. Interestingly, at the same flow depth, Figures 6.20 (b)
and (c) show that although stripes are evident for both L = 5 and L = 6, the flow
pattern is not sufficiently strong to generate vertical banding right up to the water
surface.

6.3.2.5 The Effect of Flow Depth on Quadrant Maps

The quadrant orbits plotted thus far are all for flow depth H = 50mm. Yet, the most
distinctive quadrant orbit, the quadrant ring, certainly appears in the other two flow
depths studied as part of this project, H = 37mm and H = 85mm. To illustrate this
the plots in Figure 6.21 all come from the same isolated roughness spacing, L = 10,
and show time-averaged flow patterns via quadrant diagrams and maps at all three
measured flow depths.

Looking first at the quadrant diagrams [Figures 6.21(a), 6.21(c), and 6.21(e)] reveals
that the magnitude of the rings (plotted for z = 2, 3, 4, and 5k) is roughly equal for
any given level. This point is an important one to reiterate, and tells us that velocity
patterns and magnitudes in the zone of the rings are independent of flow depth. They
are instead governed by scales associated with the channel roughness, as one would
anticipate in the roughness layer. Despite consistency in the quadrant rings, the
quadrant maps [Figures 6.21(b), 6.21(d), and 6.21(f)] do reveal a trend with flow
depth. The uppermost row of subplots is for H = 37mm (relative submergence, H/k =
5.8), and shows that the striped pattern on the quadrant map extends throughout the
flow depth. Therefore, in this case the whole flow depth can be considered as the
roughness layer. The middle row (H = 50mm, H/k = 7.9) also shows that the
quadrant map banding remains coherent throughout the flow depth, although there
are signs that it is beginning to disintegrate close to the water surface [6.21(d)]. As
such, it is unsurprising that the regular quadrant map pattern has disbanded at around
z = 6k for the highest flow depth H = 85mm (H/k = 13.4). In this case the organised
near-bed flow patterns (visualised as quadrant rings) have sufficient headroom to
dissipate through the flow depth and the remainder of the quadrant map becomes
patchy (z>6k).

PhD Thesis 167 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
(a)

(b)

(c)

(d)

(e)

(f)

Figure 6.21: Quadrant rings and maps for bar spacing L=10. (a), (b) S400_H37_L10, i.e. flow
depth H=37mm, (c), (d) S400_H50_L10, i.e. flow depth H=50mm, (e), (f) S400_H85_L10, i.e.
flow depth H=85mm. Quadrant maps for all three flow depths are shown to indicate where the
Zone of the Rings ceases to exist.

PhD Thesis 168 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
The fact that there is sufficient flow depth in the H = 85mm subset of experiments to
chart the level at which coherent patterns in local time averaged velocity break down
suggests an alternative method for estimating the extent of the roughness layer. This
assessment could for example be based upon the average radius of the quadrant rings
dropping below some arbitrary threshold (e.g. r 0.05 u* ). Alternatively the extent

of the roughness layer could be easily assessed by visual examination of quadrant


map patterns, as is presented below.

6.3.3 Re-assessing the Extent of the Roughness Layer

In Chapter 5 the extent of the roughness layer was tentatively ascertained by


searching for the level at which streamwise variation in temporally-averaged profiles
(of both velocity and Reynolds stress) became insignificant. Using this method led to
a working estimate of the roughness layer extending up to (2-6)k from the channel
bed, with the higher levels typifying isolated roughness flows. Additionally,
although the general trend in roughness layer thickness with bar spacing was
corroborated (by plotting the z-levels for negligible form-induced stress, Figure 5.25,
Section 5.4.4), no attempt was made to accurately quantify this thickness. As has
now been demonstrated this was the correct decision, as for flows over two-
dimensional roughnesses it is clearly possible to have significant streamwise
variation even when form-induced stress levels have reduced to zero (i.e. in the
Zone of the Rings). Equipped with quadrant map plots, the extent of the roughness
layer can now be re-assessed for all bar spacings. This is done for flow depth H =
85mm, the only subset with sufficient depth to support more than just a roughness
layer.

Figure 6.22 charts the dependence of roughness layer thickness on square bar
roughness spacing (evaluated by observing the level at which the coherent striped
pattern of the quadrant maps, e.g. as in Figure 6.21, begins to disintegrate). The
observed trend echoes that seen in Figure 5.25 in so far as the effect of the roughness
is felt higher from the bed for isolated roughness types. In fact, there is a near-linear
(positive slope) relationship for bar spacings L = 6 to L = 12, before the trend
flattens. Patterns of fluid flow over the widest bar spacings, L = 12, 15, and 20 are

PhD Thesis 169 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
regulated by the roughness up to z = 5.5 to 7k from the solid surface. This is some
distance beyond the traditionally recognised upper bound of the roughness layer (z =
5k, Raupach et al., 1991), and exceeds the estimate proposed in Chapter 5 by the
same amount. Surfaces comprising two-dimensional square bars may therefore be
considered as very rough. This is partly attributable to the angular geometry of the
bars, and partly to the two-dimensionality of such strip roughness.
SKIMMING FLOW

WAKE
INTERFERENCE
FLOWS ISOLATED ROUGHNESS FLOWS

Figure 6.22: Roughness layer thickness versus square bar spacing. The extent of the roughness
layer is evaluated from quadrant maps of form-induced velocities. The levels plotted refer to the
level at which the coherent banding of the quadrant maps disappears, or in other words where
rings vanish from the quadrant diagrams. The dashed line is for skimming and wake interference
flows and indicates that the process of ascertaining the thickness of the roughness layer from
quadrant maps is not so straightforward in the absence of coherent striped patterns above the
roughness crests.

6.4 Chapter Closure

This chapter began by presenting and explaining the streamwise variation in the
form-induced velocity components, u% and w% over a selection of square bar
spacings. It was demonstrated that, most especially in the roughness layer for
isolated roughness flows, the two sinusoidally varying components had a quarter
pitch spatial lag between them.

The primary focus then shifted to plotting u% and w% on orthogonal axes, a technique
which has to date been successfully utilised to visualise and quantify the structure of

PhD Thesis 170 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
turbulent streams. Novel application of the method to the form-induced velocity
components was called Quadrant Mapping, and led to the following conclusions.

Firstly, in isolated roughness flows, quadrant orbits make pseudo-circular, ring-like


patterns on the quadrant diagram. The associated quadrant map areas are typified by
vertical banding, and form-induced stress in the Zone of the Rings is negligible.
Secondly, as the observation plane approaches the bed, the quadrant rings become
distorted in the Zone of Irregular Orbits, thus generating non-negligible form-
induced stress. Thirdly, by charting the level at which quadrant map banding (or
quadrant diagram rings) disappears, the upper bound of the roughness layer can be
tracked as bar spacing varies. This led to a re-evaluation of the roughness layer
thickness; for well-isolated roughness configurations ( L 12 ) the roughness layer
extends up to 7k from the channel bed. In common therefore with quadrant analysis
for temporal fluctuations, Quadrant Mapping for the spatial fluctuations in time-
averaged flow has proved to be a powerful tool for analysing flows over rough
boundaries.

As a slight aside, towards the end of Chapter 5 the similarity (in terms of a linear
double-averaged streamwise velocity profile) between the flow in the region close to
2-dimensional square bar roughness and in that close to 3-dimensional sphere
roughness was briefly discussed. Interestingly, there is also similarity in terms of
quadrant maps between simple bar roughness and a more complex bed topography,
in so far as distinct quadrant orbits appear in flows over gravel roughness (Pokrajac
et al., 2005). However, instead of being ring-shaped the orbits are elliptic, an
indication that the streamwise form-induced velocities are much larger than those of
the bed-normal component. This is a consequence of the three-dimensionality of the
local flow around the roughness elements, where a significant portion of flow takes
place around roughness elements rather than over them as happens in the case of 2-
dimensional bar roughness.

At a general level, this chapter has served to underpin the need for spatial averaging.
For example, it is only by first calculating, and then subtracting the double-averaged
profiles from the temporally-averaged flow field that the roughness induced spatial
variation becomes apparent.

PhD Thesis 171 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
The last two chapters have focussed firstly on the double-averaged flow
characteristics (Chapter 5) and then the local variation (Chapter 6) within the
roughness layer. The next chapter looks at how local variation in the near-bed region
can influence bulk flow properties. To do so, Chapter 7 introduces a highly novel
finding, and one that could not have been anticipated from the original project brief.
Specifically, the following sections are devoted to describing the Transition
Phenomenon whereby one particular square bar configuration markedly alters the
behaviour of the overlying fluid flow. The most striking outcome is the control that
bar spacing exerts over channel discharge.

PhD Thesis 172 Chapter 6.0


L J Campbell Local Flow Characteristics in the Roughness Layer
7.0 THE TRANSITION
PHENOMENON

7.1 Introduction

The previous chapters have addressed some of the original project objectives as
given in Section 1.1. (Broadly speaking Chapter 5 addressed objectives 2 and 3,
Chapter 6 tackled objective 4, and conducting and analysing results from the
experimental programme achieved the remaining objectives 1 and 5). This chapter
however focuses on a highly novel finding, and one that could not reasonably have
been anticipated from the initial project brief. More specifically, the approach
adopted in this study of systematically altering bar spacing has led to the discovery
of a transition phenomenon. In this context transition refers to the point at which
flows over wake interference roughness configurations develop into isolated
roughness flows or, in other words, the transition between d-type and k-type
roughness. The observed flow behaviour and necessary roughness conditions at the
transition, together with the effects of the phenomenon are presented in this chapter
as follows,

Section 7.2: A brief account of the phenomenon, or, What is the Transition
Phenomenon?

Section 7.3: A discussion of the key differences between flow at, and either
side of, the transition point: Fundamental Mechanisms Underlying the
Transition Phenomenon.

Section 7.4: How the phenomenon influences global flow characteristics,


i.e. The Effects of the Transition Phenomenon Drag Adjustment and
Other Considerations.

PhD Thesis 173 Chapter 7.0


L J Campbell The Transition Phenomenon
7.2 What is the Transition Phenomenon?

In short, the transition phenomenon refers to an instability at the transition between


wake interference and isolated roughness flows at which the overall properties of the
flow are dramatically altered by the interaction between the fluid and the wall
roughness (Campbell et al., 2005). The consequences of this interaction are profound
as it radically alters the nature and the behaviour of the flow; the most striking effect
of the phenomenon is that discharge capacity is extremely sensitive to bar spacing
(as will be discussed in Section 7.4).

Both classifications of bar roughness, those of Morris (1955) and Perry et al. (1969),
were discussed in the literature review. Although the two classifications are for all
intents and purposes equivalent (refer to Tables 4.2(a) to (c) for the categorization of
the current square bar roughness spacings under either scheme), there is a significant
distinction. In Morriss classification the use of terms such as skimming, wake
interference or isolated roughness in effect offers a ready spatial description of the
roughness-generated near-bed flow patterns over bars at different pitches. It was the
aim of Chapters 5 and 6 to describe and quantify the double-averaged and local near-
bed flow structure respectively, therefore Morriss classification was deemed most
appropriate. Now however attention turns to how specific bar spacings alter the
overall hydrodynamics of the flow, instead of just the near-bed flow patterns.
Therefore, although the transition phenomenon may be equally well-described within
either the wake interference versus isolated roughness or the d-type versus k-type
terminologies, the latter classification (Perry et al., 1969), will be adopted in this
chapter. As such, the transition phenomenon is henceforth referred to as the d-k-type
transition.

The importance of areas of flow separation and reattachment for determining flow
behaviour over regular rough beds has already been amply demonstrated in this
thesis. For example, the influence of the balance between recirculating and separated
flow zones for yielding a linear double-averaged velocity profile in the interfacial
layer (Chapter 5), or for generating stable undulating flow patterns and hence ring-
like quadrant orbits on quadrant diagrams for the form-induced velocity components
(Chapter 6). In common with the above examples, the reattachment length is

PhD Thesis 174 Chapter 7.0


L J Campbell The Transition Phenomenon
absolutely central to the transition phenomenon. This is because the reattachment
length may be considered to demarcate the precise transition point at which
roughness elements begin to behave as isolated obstacles, i.e. the transition between
d-type and k-type bar roughness.

The transition phenomenon occurs when uniformly spaced, laterally-continuous,


square bars are positioned such that the reattachment point of the flow from the
upstream roughness coincides with the front face of the next roughness element
(which is precisely at the transition between d-type and k-type roughness).
Reattachment lengths reported in this study lie in the range 4.7 to 5k (Section 5.2.1),
and are in good agreement with those reported in the literature. As discussed in the
literature review (Section 2.3.3), studies of eddy formation behind single roughness
elements show that the re-attachment point of the flow behind a square bar will occur
at a length of 4.8 5k downstream of the bar (e.g. Cui et al., 2000; Leonardi et al.,
2003). The inter-bar gap in roughness configuration L = 6 is 5k, which is wholly
consistent with the reattachment length for flow over an obstacle of height k. Hence,
bar spacing L = 6 most closely represents the transition point between wake-
interference, d-type conditions ( L 5 ), and isolated roughness, k-type configurations
( L 7 ).

Although understanding of the transition phenomenon is not yet complete, in an


attempt to provide a preliminary mechanistic explanation for the observed behaviour
the following section takes a closer look at the key features characterising the flow
over bar spacing L = 6. In most of the following considerations, roughness cases L =
5 and/or L = 7 are used as the primary comparison for the transition spacing L = 6 as
they demonstrate the contrast in flow behaviour at, and immediately either side of,
the transition point.

7.3 Fundamental Mechanisms underlying the Transition Phenomenon

In general, transitions (e.g. laminar/turbulent) often produce rich and unexpected


behaviours, and this transition phenomenon is no exception. One may expect to find
clues as to the nature of the flow dynamics at transition from both the mean velocity

PhD Thesis 175 Chapter 7.0


L J Campbell The Transition Phenomenon
distribution and the turbulent structure. These clues are now presented and discussed
in turn.

7.3.1 Indications from the Temporally-Averaged Flow Field

At the d-k-type transitional bar spacing, L = 6, the inter-bar gap is such that the
reattachment point of the flow from the upstream bar coincides with the front face of
the next roughness element. The overall consequence of this particular periodic
spacing is to induce a very strong upward velocity spike at the leading face of each
roughness element. To illustrate this, Figure 7.1 compares the time-averaged
streamline patterns at the upstream face of a square bar for cases L = 5 (d-type), L =
6 (d-k transition), and L = 7 (k-type).

L=5 L=6 L=7


(a) (b) (c)

Figure 7.1: Time-averaged streamlines for (a) d-type roughness, L=5, (b) the d-k-type transition,
L=6, and (c) k-type roughness, L=7. All are for flow depth H=50mm. Subplot (b) illustrates the
concentrated vertical fluid motion at the upstream edge of each square bar; this feature only exists
for the d-k-type roughness transition. [Note, to provide further insight into the frequency and extent
of the flow patterns shown in this figure, there are two series of ten consecutive time-steps showing
instantaneous streamlines from spacing L=5 and L=6 in Appendix E].

PhD Thesis 176 Chapter 7.0


L J Campbell The Transition Phenomenon
As has been seen in previous chapters, the gap length in the L = 5 case (which is
equal to 4k) is not quite sufficient to allow the flow to fully reattach itself to the solid
bed, and a relatively stable recirculation occupies the whole region between
roughness elements [Figure 7.1(a)]. However, by extending the inter-bar gap by just
one element height in spacing L = 6, there is room for the main cavity vortex to
divide in two, essentially producing discrete trailing and leading edge separation
eddies (refer to Figure D.5 in Appendix D). Importantly, for L = 6 the bulk flow
approaches the next square bar over this leading edge separation eddy approximately
parallel to the channel bed (i.e. the optimum approach angle for carrying streamwise
momentum), and then meets the vertical face some distance from the bar crest [at
z 4.5mm , Figure 7.2(b)]. This induces a large upwards velocity at the leading edge
of each roughness element. In effect, the mean flow encounters an impermeable
barrier placed close to its natural reattachment point and is consequently forced to
suddenly move upwards. It should be noted that this unique feature is not present to
this degree for any other bar configuration and may be fairly considered to be a
transition phenomenon. Indeed, extending the bar pitch again by one element height
changes the roughness geometry from the d-k-type transition into k-type, isolated
obstacle roughness. Hence, for spacing L = 7 the inter-bar cavity is just long enough
to allow the flow to reattach to the channel bed rather than the upstream face of the
next bar. Following this brief reattachment, the shape of the leading edge separation
eddy means that the main flow is guided up and over successive bars with relative
ease [Figure 7.1(c)]. Note however that there is residual evidence of the L = 6
transition feature in spacing L = 7. As Figure 7.1(c) shows, there remains a
reasonably focussed area of upwards fluid movement at the leading edge of the bars,
albeit as a diminished or weaker form of the flow behaviour at transition, and no
longer fixed so sharply in the bed-normal direction.

To further demonstrate the extent and magnitude of the vertical velocity spike
generated at the d-k-type transitional roughness, Figures 7.2 and 7.3 show contours
of time-averaged velocity in the near-bed region for L = 5 and L = 6 (in each case
streamwise, u , and bed-normal, w , mean velocities are shown in the top and
bottom windows respectively). Firstly, comparing the mean streamwise velocities
between L = 5 and L = 6 [Figures 7.2(a) and 7.3(a)] shows that there is an area of
near-zero streamwise momentum just above the crest at the upstream face of the

PhD Thesis 177 Chapter 7.0


L J Campbell The Transition Phenomenon
square bars in both cases. Although this penetrates around twice as far into the
overlying flow in the transitional spacing, the true extent of the difference between
bar pitches L = 5 and L = 6 only becomes clear when contours of mean bed-normal
velocity [i.e. Figures 7.2(b) and 7.3(b)] are compared. Whilst Figure 7.3(b) illustrates
a marked vertical disturbance at the upstream bar face in L = 6, this attribute is all
but absent in L = 5. More expressly, for the L = 5 roughness configuration the
maximum time-averaged vertical velocity anywhere in the flow field is
approximately 4% of the global mean velocity, U; at the L = 6 spacing this rises to
over 15%, sharply focussed at the leading edge of every square bar.

(a)
(a)

(b)
(b)

Figure 7.2: Contours of time-averaged Figure 7.3: Contours of time-averaged


(a) streamwise, and (b) bed-normal (a) streamwise, and (b) bed-normal
velocities (m/s) in the near-bed region velocities (m/s) in the near-bed region
for L=5 d-type spacing (flow depth, for L=6 d-k-type transitional spacing
H=85mm). Flow is from left to right. (flow depth, H=85mm). Flow is from left
to right.

Given that the strong vertical velocity spike will exist to some degree right across the
whole width of the flume (as the bars are laterally continuous), it has a profound

PhD Thesis 178 Chapter 7.0


L J Campbell The Transition Phenomenon
influence on the overall dynamics of the flow. In the region of the flow near the top
of the roughness elements, continuity demands that a compensatory downwards
velocity must exist. Otherwise, from very simple considerations, continuity would be
violated for the bed-parallel plane between the successive roughness tops. The
proximity of the compensatory downwards velocity to the upwards spike in the L = 6
case produces a velocity couple which causes rotation in an anti-clockwise direction
(when flow goes from left to right). Subtracting the double-averaged velocity
profiles, u ( z ) and w ( z ) , from the temporally-averaged flow field [to yield the
form-induced flow field, u% ( x, z ) and w% ( x, z ) ] serves to highlight this zone of
counter-clockwise rotation at the d-k-type transition point. By plotting form-induced
velocity vectors (i.e. the magnitude and direction of the vector sum of
u% ( x, z ) = u ( x, z ) u ( z ) plus w% ( x, z ) = w( x, z ) w ( z ) at each grid point), Figure 7.4
shows the large fixed region of counter-rotation induced by the L = 6 bar spacing.

Figure 7.4: Velocity vector plot of the temporally averaged (form-induced) flow field at the d-k-
type roughness transition (L=6) for experiment S400_H85_L6. Flow is from left to right. The
fixed large anti-clockwise eddy centred above x = 30 mm is formed at the leading edge of the
downstream bar as a consequence of the vertical velocity spike shown in Figure 7.3(b).

From Figure 7.4 the effect of the strong local flow disturbance at the upstream face of
the bar in terms of both flow pattern and extent becomes clear, in the form of a large,
fixed coherent vortex. The fact that the presence of this vortex is revealed by
subtracting the double-averaged flow field in each bed-parallel plane (i.e. a non-
constant frame of reference) tells something about the resulting momentum transfer

PhD Thesis 179 Chapter 7.0


L J Campbell The Transition Phenomenon
in bed-parallel flow layers. The uppermost portion of the vortex represents an area of
slower-than-average streamwise fluid motion ( u% < 0 ). As such, it acts as a
streamwise momentum sink for flow at that level. The extent of this flow behaviour
is particularly worthy of note as it is not simply restricted to the near-bed region.
Instead there is evidence of u% < 0 areas above the large vortex up to z 40mm, i.e.
effectively throughout the whole of the lower half of the flow depth. Furthermore, the
anti-clockwise vorticity demonstrated in Figure 7.4 opposes the natural clockwise
vorticity that is part-and-parcel of the shearing over a rough wall. Put simply, the
inherent shear-driven turbulent structure (e.g. burst and sweeps) is opposed by
counter-vorticity generated by the interaction of the flow and the wall geometry at the
d-k-type roughness transition.

The above observations would suggest that the natural turbulent structure might be
altered through a significant portion of the water depth in flows over the d-k-type
transitional square bar roughness. The following section explores whether or not this
is the case.

7.3.2 Turbulent Structure at the d-k-type Roughness Transition

The hypothesis that counter-vorticity generated at the d-k-type transition interferes


with the natural turbulent structure of the flow may be substantiated by examining the
velocity time-series. The practical approach adopted is to revisit a valuable tool
introduced in the last chapter, namely quadrant analysis. In this instance however,
instead of using quadrant analysis for the form-induced velocity components, it has
been applied in the conventional manner, i.e. to the instantaneous temporal velocity
fluctuations, u and w . As discussed (Section 6.3.1), quadrant diagrams are a useful
means of understanding and visualising the turbulent structure of the boundary layer
flow such that the burst-sweep cycle and shear Reynolds stress are both indicated by
the negative correlation of the velocity fluctuations.

Figure 7.5 shows three quadrant diagrams, one from each of roughness spacings L =
5, L = 6, and L = 7. All plots show data from a single velocity time-series (4096
time-steps, or around 300 seconds of real-time flow) from a point located at z = 2k

PhD Thesis 180 Chapter 7.0


L J Campbell The Transition Phenomenon
above the middle of the inter-bar gap. Best-fit correlation lines (i.e. w = au ) are
marked in red to indicate the general structure of the data. This shows that here is a
perceptible flattening of the slope at the d-k-type transitional roughness [Figure
7.5(b)]. In other words, the slope coefficient, a, becomes less negative. As this
flattening may be difficult to discern from the plots to the left, Figure 7.6 compares
contour maps of quadrant diagram point density for L = 5 and L = 6 [this shows the
same time-series plotted in Figure 7.5(a) and (b)].

(d)
* z=2k
(a)

Figure 7.5: Temporal quadrant analysis


diagrams for (a) L=5 (d-type), (b) L=6 (d-
k-type transition), and (c) L=7 (k-type)
(b) roughness configurations. Red lines show
best-fit correlation, i.e. w = au Velocities
are normalised with friction velocity. (d)
shows location from which temporal time
series were generated.

(c)

With reference to Figure 7.6, the flattening of the slope of the best-fit correlation line
at the d-k-type roughness transition [subplot (b)] is now clear. A distinct advantage of
obtaining PIV velocity data is that one can ascertain whether the flattened slope of the
quadrant diagram is extensive, or occurs at just a few spatial locations in the flow
above the square bars at the d-k-type roughness transition. Contour maps of this slope
(the coefficient, a) for roughness configurations L = 5, L = 6, and L = 7 are presented
in Figure 7.7.

PhD Thesis 181 Chapter 7.0


L J Campbell The Transition Phenomenon
(a) (b)

Figure 7.6: Contour maps of point density for the quadrant diagrams in Figures 7.5(a) and (b).
Subplot (a) is for d-type roughness spacing L=5, (b) shows the d-k-type roughness transition,
L=6. Warm colours indicate high numbers of data points, whilst cold colours indicate few.

(a) (b)

Figure 7.7: Contour maps showing


the spatial variation of slope
(c)
coefficient, a, of temporal quadrant
plots, i.e. w = au. These relate to
bar spacings (a) L=5 (d-type), (b)
L=6 (d-k-type transition), and (c)
L=7 (k-type). The large yellow area
in subplot (b) illustrates an extensive
zone of reduced quadrant map slope
at the d-k-type transition. Flow depth,
H=85mm.

PhD Thesis 182 Chapter 7.0


L J Campbell The Transition Phenomenon
Examination of Figure 7.7 (which takes data from H = 85mm) shows that the slope
coefficient, a, has perceptively flattened for the transitional roughness, L = 6,
through much of the flow depth. Indeed, the extent of this flattening (up to
z 40mm) is precisely consistent with the extent of the fixed counter-rotating eddy
in Figure 7.4, as discussed in Section 7.3.1. The flattening of the slope coefficient for
spacing L = 7 [Figure 7.7(c)] is simply a weaker version of the transition
phenomenon, and may be attributed to the residual vertical velocity effect shown in
Figure 7.1(c). Section 7.4 reports how channel discharge varies with bar spacing. In
this respect, it must be noted that the same trend of attenuation of vertical velocity
fluctuations was observed for the enhanced discharge conditions at the 50mm flow
depth (refer to Appendix E, Figures E.3-E.5, for equivalent contour maps for flow
depth H = 50mm), and to a lesser extent for H = 37mm. The pattern also emerged for
the resonant spacing L = 12. This signifies a systematic and extensive alteration to
the turbulent structure, as discussed below.

The flattening evident in Figure 7.7(b) is indicative of the decrease in the relative
magnitude of vertical velocity fluctuations with a corresponding increase in
streamwise velocity fluctuations (for L = 6). This demonstrates that turbulent
fluctuations are less influential in vertical momentum transfer at L = 6 and indicates
that the shear-driven turbulent boundary-layer flow has undergone some degree of
re-laminarisation in which the natural turbulence has been suppressed by the
interaction of the flow and surface geometry. Decreasing the amount of fluid mixing
in this manner reduces the energy required to overcome turbulence. Observe that this
change in turbulent structure is especially noticeable in the near-wall region in the
gaps between square bars, the very area where turbulent eddies make a significant
contribution to surface drag. In other words, the counter-vorticity generated by the
flow at the dk-type roughness transition is substantially disrupting the formation
and growth of turbulent eddies. Furthermore, Figure 7.6 showed that compared to L
= 5, flow over bars spaced at L = 6 is characterised by a large number of temporal
quadrant points clustered towards ( u , w ) = (0,0). Consequently, with fewer large
vertical velocity fluctuations, the streamwise momentum is not carried and mixed in
the vertical direction and the flow regime can be considered to be partially re-
laminarised. The most obvious consequence of re-laminarisation is of course to

PhD Thesis 183 Chapter 7.0


L J Campbell The Transition Phenomenon
reduce surface drag. This drag adjustment is discussed as part of the following
section.

7.4 Effects of the Transition Phenomenon Drag Adjustment and Other


Considerations

7.4.1 Drag Adjustment

Before presenting and discussing results from the current study showing how
channel discharge varies with square bar spacing, this section begins with a brief
introduction to a range of existing drag reduction techniques. Note that the term
drag reduction is used to imply a special case of drag adjustment, such that the
rough-wall drag level is less than that of a smooth wall.

7.4.1.1 Conventional Drag Reduction Techniques

In practice fluid friction or drag is generally governed by turbulence. Streamwise


momentum is taken from the mean flow and transferred to the vertical and cross-
stream directions by turbulent fluid mixing. It is estimated that in the near-wall
region of a flow larger eddy structures account for over 80% of the turbulent energy
such that the burst-sweep cycle associated with larger eddies makes a dominant
contribution to surface friction (Lu and Willmarth, 1973). Minimising the energy
transferred to turbulence is therefore the key to reducing drag. Consequently much
research effort has concentrated on reliable and practical techniques for drag
reduction. Drag reduction techniques function by controlling either small-scale
turbulence (e.g via longitudinal grooves or riblets), or by interrupting large turbulent
eddies (e.g. large eddy breakup devices). Although extensively studied, it should be
noted that in many cases the precise mechanism sustaining the drag reduction
phenomenon is not completely understood (Jacobson and Reynolds, 1998; Warholic
et al., 1999).

Drag reduction methods may be classified as either active or passive. Active


measures involve external energy input, typically in response to sensing large-scale
turbulent events as they form and develop. This triggers a mechanism geared

PhD Thesis 184 Chapter 7.0


L J Campbell The Transition Phenomenon
towards preventing further growth or fragmentation of the eddy. Examples of such
feedback mechanisms include breakup devices in the form of mechanical flaps
(Lumley and Blossey, 1998) and vibrating surfaces (Rathnasingham and Breuer,
1997a, 1997b). Or, input energy may be used to drive alternate suction and blowing
through a perforated wall, with the aim of countering large quasi-streamwise vortices
as they develop in the turbulent flow. Attempts to generate drag reduction by this
alternate sucking and blowing method have recently become relatively commonplace
(e.g. Tardu, 2001; Baker et al., 2002).

In contrast to the energy requirement of active measures, passive drag reduction


methods require no external energy source, and instead rely on altering fluid
properties or honing the surface topography. Reported techniques involving the
modification of fluid properties include viscous drag reduction via the addition of
polymer chains (e.g. Berman, 1978; Lumley, 1969), and the injection of
microbubbles (Xu et al., 2002). Methods that alter the surface roughness include
riblets (Walsh, 1990; Bechert et al., 1997) and wall protrusions (Sirovich and
Karlsson, 1997). Taking inspiration from tiny tooth-like protrusions covering
sharkskin, the dermal denticles, the practice of utilising riblets (typically triangular
grooves a few tens of microns deep) is arguably the most widely known and
accepted drag reduction technique. Indeed, riblets have been adopted in a number of
high-profile situations. They were employed on the hull of the 1997 Americas Cup
winning vessel, the Stars and Stripes, for which it was estimated they enabled an
overall 2% increase in speed (however, more controlled conditions suggest it may be
possible to gain 7-8% drag reduction). More recently, riblet-like additions have been
incorporated into the fabric of professional sports gear, for example speed-skating
suits and swimwear.

A small number of studies (Choi and Fujisawa, 1993; Matsumoto, 1993) have
concluded the possibility of drag reduction over d-type grooved walls by way of the
cavity vortex absorbing and reorganising the incoming turbulence. However, the
mechanism reported here occurs specifically at the d-k-type transition, and is
therefore quite different. Before proceeding to a discussion of the drag reducing
influence of the transition phenomenon it is worth noting that the downwards and
upwards velocity couple, or counter-vorticity, generated by the interaction of the

PhD Thesis 185 Chapter 7.0


L J Campbell The Transition Phenomenon
flow and the wall geometry at the d-k roughness transition (Figures 7.3(b) and 7.4) is
effectively equivalent to a roughness-driven sucking and blowing action occurring
alternately over the wall region at the level of the roughness tops. In this respect, the
transition phenomenon may be considered to mimic a significant feature of an
existing recognised drag reduction technique, that of alternate wall suction and
blowing. However, the crucial distinction is that whereas true wall suction and
blowing is an active drag reduction technique, the roughness-driven flow counter-
vorticity patterns at the d-k-type transition are generated without the need for energy
input, i.e. passively.

Note that as the discharge capacity of the square bar-roughened beds did not quite
exceed the smooth bed value for a given flow depth, the following section is written
in terms of drag adjustment, not absolute drag reduction.

7.4.1.2 Drag Adjustment at the d-k-type Roughness Transition

A central finding from this study is that channel conveyance is extremely sensitive to
roughness bar spacing. The contour maps in Figures 7.2(a) and 7.3(a) showed
markedly different streamwise velocities, with much faster flow evident in L = 6. To
illustrate this variation, Figure 7.8 shows how channel discharge (Q) varies with
roughness spacing (L) for all three measured flow depths H = 37, 50 and 85mm.

Channel conveyance in this study was evaluated by two distinct methods, by


integrating the double-averaged streamwise PIV velocity profile through the flow
depth, or, more directly by the v-notch weir in the holding tank at the downstream
end of the flume. Discharge data plotted in Figure 7.8 were as measured by the v-
notch weir (values are given in Appendix E), whilst for comparison Figure 7.9
provides the value calculated by integrating the velocity profiles. The two
independent measures produce essentially very similar results, noting however that
discharge deduced from the PIV centre-line velocity profiles tends to be higher than
the weir reading, as the former method ignores the effect of the channel side walls.

Although the PIV and weir measurements of channel conveyance show very similar
patterns there are certain discrepancies. For example, there are distinct differences in

PhD Thesis 186 Chapter 7.0


L J Campbell The Transition Phenomenon
the PIV and weir discharge patterns for bar spacings L = 2 and L = 3 at the lower two
flow depths. Also for spacing L = 20 it is known that the PIV camera was incorrectly
positioned such that the flow towards the top of the water column was not recorded,
and therefore direct comparison with the weir measurements is not possible for this
spacing. As part of the investigation the laboratory set-up was carefully checked and
there appeared to be no obvious problems or pitfalls with experimental and
equipment design. There is however a degree of cross channel variation (see Section
7.4.2) as evidenced by differences in the bar spacing yielding maximum discharge
between two measurement techniques. The investigated flow depths were chosen
such that the channel aspect ratio was maintained at B/H 4.5 (after e.g. Graf, 1998)
as this was deemed an acceptable limit for ensuring largely 2-dimensional flow in
the channel centreline. However, with the benefit of hindsight this limit was most
probably insufficient to validate the assumption of 2-dimensional flow. Knight
(1981) used the mean wall shear force as a criterion to aid in the definition of a
wide channel (i.e. one in which 2-dimensional flow may be reasonably assumed in
the centreline). He concluded that B/H = 12.6 22.6 (depending on the bed
roughness) was necessary to limit the shear force carried by the sidewalls to 5% of
the total boundary shear force. The clear downside of this constraint would of course
be that in a 400 mm wide flume, the maximum safe flow depth for 2-dimensional
flow in the channel midline is limited to around 30 mm.

In addition to cross-channel variation, other possible contributing factors to the


discrepancies between discharge measurements include problems with the weir and
the downstream holding tank capacity. The weir may be susceptible to error at
extremes of the discharge range (in terms of pump capacity). For example, at high
channel discharge the flow may be insufficiently calmed in the downstream holding
tank in the vicinity of the v-notch weir (a degree of calming is achieved by a pivoting
cylinder free to float on the water surface in the tank). At low discharges the
calibrated scale can become difficult to read accurately. Care must also be taken to
ensure that the downstream tank has sufficient water to supply the pump at high
discharges - if the water level drops too low the pump can have insufficient head to
perform consistently. Despite identifying these issues, the particular reasons for the
discrepancies between the PIV- and weir-evaluated patterns of discharge remain
largely unclear.

PhD Thesis 187 Chapter 7.0


L J Campbell The Transition Phenomenon
Figure 7.8: Variation in weir-measured channel discharge, Q (l/s), with roughness spacing, L, for
all three flow depths.

Figure 7.9: Variation in PIV-measured channel discharge, Q (l/s), with roughness spacing, L, for
all three flow depths. Note how there appears to be a resonant effect as discharge attains local
maxima at L=6 and L=12. This resonant behaviour is also reflected in discharge minima, with low
discharge for all three flow depths occurring at L=5, 10 and 15. The change in discharge between
L=5, 6, and 7 for flow depth H=85mm matches the trend in the mean flow and turbulence clues
underlying the transition phenomenon as illustrated in other figures (Figures 7.1 and 7.7). The
dashed lines to data at spacing L = 20 indicate uncertainty in the PIV discharge values due to
incorrect camera positioning (see main text).

From the v-notch weir data (Figure 7.8), the variation in conveyance is similar for all
flow depths with a notable rise in discharge occurring between L = 5 and L = 6 or 7.
The fact that such a peak in conveyance occurs at L = 6 (H = 85mm case), and at L =
7 (H = 37 and 50mm cases) implies either that the drag-reducing effect of the
roughness exhibits some sensitivity to flow depth, or that the roughness arrangement
has not been completely optimised, i.e. that the maximum discharge augmentation

PhD Thesis 188 Chapter 7.0


L J Campbell The Transition Phenomenon
actually occurs somewhere in the roughness spacing interval 6<L<7. At this point it
is useful to recall Figure 7.1, which illustrated some residual d-k-type transition flow
pattern behaviour persisting at L = 7. However, turning attention to discharge as
evaluated from the PIV centre-line velocity profiles (Figure 7.9) consistently shows
local peaks at bar spacing L = 6, regardless of flow depth. The slight discrepancy
between the two trends (weir versus PIV) suggests that the velocities in the mid-line
of the channel may not always be totally representative of the whole channel width.
The possibility of significant cross-channel variation as part of the transition
phenomenon remains one of the key unanswered question from this project, and is
discussed in Section 7.4.2.

The effect of the bar roughness on the mean flow can be expressed in terms of
Mannings n. The stage-discharge curves from repeated experiments (in different
hydraulic flumes) indicate that Mannings n for the transitional spacing, L = 6, is
equivalent to that for a smooth and clean excavated earth channel (as listed by The
Fluid Mechanics Calculation Website, 2000), with nL6 = 0.022. The Mannings n
value for spacing L = 5 is nL5 = 0.035.

The patterns of discharge variation shown in Figures 7.8 and 7.9 can also be
interpreted as charting the variation in friction factor with bar spacing. At first glance
this variation would appear to be at odds with much previous research. However,
many of the differences between the current study and previous work with square
bars were discussed in the literature review following Table 2.1. Issues such as the
limited range of bar spacings investigated, and more importantly a lack of open-
channel flow work with small relative submergence (the key focus of this study)
were highlighted. For example, the Engineering Sciences Data Item (ESDU)
Number 79014 presents many figures showing a peak in friction factor at around
spacing L = 10. Although the family of curves presented are smooth and continuous
over the range 2 L 100 , these curves are the result of a great deal of interpolation
between measured data points. Indeed, much of the underlying research cited as used
to create the friction factor curves is included in Table 2.1 (e.g. Webb et al., 1971,
who measured over spacings L = 10, 20, 40). The ESDU publication is also
specifically geared towards pipe flows, not free-surface flows. Similarly, Figure 2.13

PhD Thesis 189 Chapter 7.0


L J Campbell The Transition Phenomenon
illustrated how friction factor was found to vary with bar spacing by a number of
researchers (Knight and MacDonald, 1979a; Adachi, 1964; Raju and Garde, 1970;
Sayre and Albertson, 1963). Whilst this figure does indeed pertain to open-channel
flow, the relative submergence is not comparable to that used in this study (it shows
results for H/k = 32), and the roughness Reynolds number would suggest that flow
may not be fully hydraulically rough.

This study has focussed firmly on hydraulically rough shallow open-channel flow
over a broad range of roughness spacings. As has been seen in previous chapters, the
effect of the square bar roughness can extend many times the bar height into the
overlying flow. As such, it is perhaps not surprising that the friction factor behaviour
is altered to that previously observed in deeper flows. It is also worth stating that the
observed transition phenomenon (and indeed the overall trends of channel
conveyance variation with bar spacing) will most likely only occur across a limited
range in parameter space (e.g. flow and roughness Reynolds numbers, relative
submergence). The transition effect also occurs for a very specific bar spacing. As
such, bar spacings either side of the transition point would not exhibit the observed
behaviour to the same extent. A simple parametric study recording channel discharge
for a wide range of hydraulic and roughness parameters is required to elaborate
further on this point.

In conclusion to this section, it is useful to recapitulate the identified facts relating to


this new drag adjustment mechanism at the d-k-type roughness transition. The most
efficient roughness arrangement (i.e. the one that conveys maximum channel
discharge) comprises transverse square bars of height k positioned with an inter-bar
gap equal to the flow reattachment length (~5k). This configuration results in sharp
local disturbance in the near-wall flow at the back face of each roughness element,
which is induced by the reattachment of the wake shed by the upstream roughness
element [Figure 7.3(b)]. Given that the concentrated vertical velocity spike is
periodic in the longitudinal direction (happening at every leading edge of the
regularly spaced roughness bars) and occurs at all locations across the channel, it is
reasonable to hypothesise that this discernible local flow disturbance and associated
compensatory downwards velocity is markedly disrupting the large eddies in the
overlying flow. Indeed, the counter-vorticity generated by the interaction of the flow

PhD Thesis 190 Chapter 7.0


L J Campbell The Transition Phenomenon
and the bed roughness at the d-k-type transition does alter the natural turbulent
structure in the main stream [Figure 7.7(b)]. If large turbulent eddies are disrupted or
prevented from growing, the capacity of the flow to transfer momentum in the bed-
normal and cross-stream directions is impeded and the flow can be considered to be
partially re-laminarised. Consequently, the streamwise velocity is enhanced yielding
a marked increase in channel conveyance, relative to neighbouring bar spacings,
without any increase in flow depth at the d-k-type roughness transition (Figures 7.8
and 7.9).

To further investigate the transition phenomenon, two supplementary experimental


conditions were studied. The second involved recording PIV images from a bed-
parallel, rather than bed-normal, plane to evaluate cross-channel effects at the
transitional bar spacing, L = 6 (this work is presented in Section 7.4.2.1). The first
supplementary experiment was conducted to record discharge in the absence of
transverse bars to address to what degree channel conveyance over the transitional
bar spacing approaches that over a smooth bed.

7.4.1.3 Supplementary Experiment 1: The Smooth Bed

In addition to the rough bed experiments, channel discharge was recorded for a range
of flow depths without bars on the bed. The base roughness for the smooth bed case
was the same bed that was used to support the transverse bars, varnished (but slightly
grainy) marine plywood. Figure 7.10 shows the experimental data points plotted
alongside the theoretical discharge curve calculated using Mannings equation with
Mannings n = 0.012. This value corresponds to a smooth mortar surface, n = 0.011
to 0.013 (values taken from Chadwick and Morfett, 1998) and therefore provides a
physically sensible comparison for the varnished wooden surface of the hydraulic
flume.

From Figure 7.10 it may be seen that the experimental data points are in good
agreement with theoretical values for depth-uniform flow, and cover the range of
discharges measured over the rough beds (up to Q ~14l/s). Using these smooth bed
values, Figure 7.11 shows the smooth versus rough wall comparison by normalising
the weir-measured rough wall discharges (previously shown in Figure 7.8) with the

PhD Thesis 191 Chapter 7.0


L J Campbell The Transition Phenomenon
equivalent smooth wall discharge (Qo). On this figure therefore, a data point would
have to fall above the dashed line representing Q/Qo = 1 to indicate absolute drag
reduction with respect to the smooth wall.

Figure 7.10: Weir-measured discharge values for a smooth bed. Dashed line illustrates
discharge calculated using the Manning equation with n=0.012.

Reference to Figure 7.11 shows that several rough wall cases do approach the
smooth wall value. Firstly, the discharge over d-type bar spacing (L = 3) at the lower
two flow depths comes close to the smooth bed value. This result provides support
for the work of Choi and Fujisawa (1993) and Matsumoto (1993) which reported
drag reduction over d-type rough walls (Section 7.4.1.1), although the present data
does not suggest the rough wall drag is less than the smooth wall (the data points still
fall below the dashed line in Figure 7.11). Secondly, the discharge at the widest bar
spacing, L = 20, for the lower two flow depths has effectively recovered to that of
the smooth bed. This suggests that despite the significant local flow disturbance in
the vicinity of the obstacle for L = 20 (refer to Chapter 6), there is sufficient length
of plane bed between the bars that the flow discharge is able to match that of the
smooth bed. One may reasonably speculate that points for bar spacings wider than L
= 20 would also fall along the Q/Qo = 1 line.

Turning attention to the discharge at the transitional bar spacing (L = 6 or 7, weir


data) shows that the peak discharges reach around 70-90% of the smooth wall value.

PhD Thesis 192 Chapter 7.0


L J Campbell The Transition Phenomenon
In this respect, it is worth noting that there is most probably some potential for
increasing the peak transitional discharge by optimising the roughness geometry for
any given flow. The optimisation of bar spacing will almost certainly exhibit some
dependence on flow depth, or more specifically Reynolds number. The weir-
measured discharges (Figure 7.8) from the present data already suggest that the
optimal bar spacing varies with relative submergence, H/k. This may be related to
the brief discussion of reattachment lengths reported in the literature (Section 2.3.3).
In other words, although the wake shed from each bar predominantly scales with the
bar height, k, the precise reattachment length, and hence transitional spacing, may
vary further depending on impinging flow velocity.

Figure 7.11: Weir-measured rough-bed discharge values normalised with the smooth
bed discharge, Qo, for all bar spacings and flow depths. The dashed line is positioned
at Q/Qo=1.

It is also interesting at this point to speculate what may happen if the background
roughness underlying the transverse bars is rough, not smooth. Figure 7.11
effectively compares the discharge over bar-roughened beds to a hydraulically
smooth surface. One may expect the gap between rough and smooth bed discharges
to narrow if the base roughness provided hydraulically rough flow conditions. It is
not inconceivable that the conveyance of a channel with optimally spaced transverse
bars fixed to an already hydraulically rough surface may actually increase compared
to the bar-free condition. Further laboratory experiments are needed to confirm
whether or not this is the case.

PhD Thesis 193 Chapter 7.0


L J Campbell The Transition Phenomenon
7.4.2 Lateral Effects at the d-k-type Roughness Transition

All of the experiments routinely conducted as part of this study (i.e. those listed in
Table 4.2) have recorded centre-line velocities. As the data do not include any
measurements away from the mid-line of the channel, the role of cross-channel
variation cannot be fully assessed. However, there are some indications, from an
increase in the non-linearity of the Reynolds stress profiles, that the transition
phenomenon induces significant lateral effects. This behaviour deserves thorough
investigation as it may hold the key to forming a deeper understanding of the
underlying drag reduction mechanism. Whilst this is not entirely possible from the
dataset as it currently stands, the preliminary evidence from centre-line
measurements, together with one supplementary transverse PIV experiment, is
presented and discussed here.

Figure 7.12: Double-averaged


Reynolds stress profiles for
roughness spacings L=5, L=6,
L=7, L=8 (H=50mm). All d-type
(L=5) and k-type (L=7, L=8)
roughnesses, adhere to a linear
distribution. However, the profile at
the d-k-type transition deviates
from linear, especially through the
lower half of the flow depth. All
profiles deviate from linear towards
the free surface, under the
influence of the glass sheet.

u w / u*2

Taking Reynolds stress as a starting point, Figure 7.12 shows the double-averaged
turbulent stress distribution for roughness configurations at, and either side of, the d-
k-type roughness transition. Excluding minor deviation towards the water surface
(which occurs as a consequence of the surface glass sheet that was used to pass the
laser sheet through the water surface) the profiles for both d-type (L = 5) and k-type

PhD Thesis 194 Chapter 7.0


L J Campbell The Transition Phenomenon
(L = 7 and L = 8) arrangements are satisfactorily linear down to the level of the
roughness layer. In contrast, the stress distribution at the d-k-type transition departs
from linear, most notably through the lower half of the flow depth.

The non-linear L = 6 Reynolds stress distribution is consistent with interference from


secondary currents. Interestingly however, the data in Figure 7.12 come from
experiments with a channel aspect ratio, B/H = 8 (flume width, B = 400mm, flow
depth, H = 50mm). As such, one would not expect the effects of coherent cross-
channel variation to significantly influence flow in the channel mid-line. For
example, Graf (1998) stated that the limiting value for flume aspect ratio for the
maintenance of 2-dimensional flow conditions in the flume centreline is B/H = 5;
flows with values lower than this may be susceptible to secondary circulation. Song
et al. (1994) consider the 2-dimensional flow assumption to be valid in the centreline
of their flume for aspect ratios, B/H3.5. Therefore, the aspect ratio for data in
Figure 7.12 equal to 8, was above Grafs recommended safe value, and well above
the value proposed by Song et al. for rough beds. Nevertheless, it appears that the
transitional spacing L = 6 induces lateral variation even at aspect ratios
conventionally considered to be free from such effects (although, as discussed in
Section 7.4.1.2, the analysis of Knight (1981) would suggest otherwise).

In addition to the turbulent stress distribution, further indication that the transition
phenomenon may actively alter the natural flow patterns come from the double-
averaged vertical velocity profiles, w ( z ) . Figure 7.13 shows how w behaves
throughout the flow depth, with the aim of demonstrating that the natural secondary
circulation is disrupted at the d-k-type transition. For d-type spacings L = 2 to L = 5,
w is slightly positive for most of the flow depth above the roughness crests. At the
transition point however, there is a step-change in w as it switches in the main
stream from positive to negative, before slowly recovering through spacings L = 7,
and L = 8 back to positive w for L 8 .

Alongside the flume aspect ratio, the overall magnitude of the recorded vertical
velocities in this study would not suggest any persistent interference from secondary
currents. Song et al. (1994) consider 2-dimensional flow conditions to exist (albeit

PhD Thesis 195 Chapter 7.0


L J Campbell The Transition Phenomenon
with weak secondary currents) when their mean vertical velocities are up to 3.8% of
the maximum streamwise velocity. In this project the mean vertical velocities are, at
most, 5 mm/s, which constitute up to 1% of the maximum streamwise velocity.
Clearly however, the Reynolds stress profile for L = 6, when taken in conjunction
with the mean vertical velocity distribution, would indicate that the transition
phenomenon does indeed induce significant lateral effects.

Figure 7.13: Double-averaged bed-normal (vertical) velocity distribution, w ( z ) for bar


spacings L=2 to L=15. Velocities in the main stream switch from positive to negative at the d-
k-type transitional spacing L=6.

In light of this, a single supplementary experiment was conducted to gain further


qualitative insight into the potential cross-channel effects at the d-k-type transition.

PhD Thesis 196 Chapter 7.0


L J Campbell The Transition Phenomenon
7.4.2.1 Supplementary Experiment 2: PIV Evidence from a Transverse (x, y) Plane

The supplementary experiment involved gathering 2-component velocity data (u and


v) from a transverse cross-section in the x, y plane, i.e. aligned parallel to the channel
bed. The laser light sheet was directed through the side wall of the flume to be
centred in the form-induced sublayer 9mm above the bar crests (at z 15 mm).
Digital cross-correlation PIV images were recorded as for the main experimental
programme (4096 time-steps at 13Hz), but with the camera situated above the flume,
looking downwards onto the bed-parallel measurement plane. The images captured
flow from a central 90mm of the 400mm wide flume, with the centre of the image
located around 15mm from the flume midline; the uniform flow depth was just over
50mm.

(a) (b)

Figure 7.14: Results from a bed-parallel plane


located 9mm above the bar crests. (a) contours of
(c)
time-averaged streamwise velocity, u ( x, y ) , (b) the

double-averaged streamwise velocity profile, u ( y ) ,

and (c) digital image of the underlying bed


topography. The dark cluster in subplot (a), and the
corresponding notch in the profile in subplot (b), is
the result of poor PIV data due to surface glare. y=0
is approximately located along the centre-line of the
flume.

Figure 7.14 shows contours of time-averaged, and the double-averaged profile of


streamwise velocity (subplots (a) and (b) respectively). Subplot 7.14(c) presents an

PhD Thesis 197 Chapter 7.0


L J Campbell The Transition Phenomenon
image of the underlying section of channel bed, from which it is evident that the
recording area traversed two square bars. Note that the area of poor data evident in
Figure 7.14(a), and the associated notch in the profile shown in subplot (b) is most
probably the result of surface glare and should be ignored.

The fact that there is marked variation in mean streamwise velocity in the x-direction
is no surprise, and again underlines the need for double-averaging over rough beds.
As has been shown in previous chapters, the downstream velocity increases as the
flow area constricts over each square bar, and slows between the bars. There is also
however considerable variation in the y-direction, with faster fluid motion towards
the centre of the flume (around y = 0). This is consistent with the above observation
that the double-averaged bed-normal velocity is substantially negative in the main
stream above the L = 6 bar spacing (Figure 7.13). In other words, in the mid-line of
the flume faster moving fluid is being directed down from the outer flow layer
towards the channel bed, augmenting the streamwise motion. This effect dissipates
as the observation point moves away from the channel centre-line, such that the
double-averaged profile, u ( y ) [Figure 7.14(b)], shows that there is a 17%
reduction from the peak double-averaged streamwise velocity towards the left bank
( y > 0 ), and a 10% reduction towards the right ( y < 0 ).

(a) (b)

Figure 7.15: The transverse velocity component. (a) contour map of time-averaged transverse
velocity, v ( x, y ) (colour bar marked in units of mm/s), and (b) the double-averaged transverse

velocity distribution, v ( y ) . For y>25mm and y<-40mm, v begins to decrease, indicating the

possibility of further secondary cells distributed across the channel.

PhD Thesis 198 Chapter 7.0


L J Campbell The Transition Phenomenon
Further information about the likely cross-channel circulatory structure of the d-k-
type transitional flow comes from the transverse velocity component, v. A contour
map of time-averaged transverse velocity, together with the corresponding double-
averaged profile is shown in Figure 7.15. Values of v > 0 indicate fluid motion
towards the left flume wall, and v < 0 towards the right. This shows that in the plane
15mm from the channel bed, fluid disperses away from the flume centre-line to both
the left and the right. The maximum double-averaged transverse velocity is between
4 and 5 mm/s. Although this is only around 2% of the peak average downstream
velocity, the effect is significant, producing marked cross-channel variation as
demonstrated in Figure 7.14(b).

(a)
B/H=1

B/H=2

B/H=3

B/H=4

(b)

z
y

Figure 7.16: (a) Secondary flow patterns (looking along the main channel axis) for different
rectangular duct aspect ratios as measured by Knight and Patel (1985). (b) The pattern of
secondary circulation cells suggested for the current dataset above transitional bar spacing L=6
(H 50mm, B/H 8). Solid lines indicate cells supported by data from the bed-parallel PIV data,
whereas dashed lines indicate further likely secondary cells (for the central part of the flume only).

PhD Thesis 199 Chapter 7.0


L J Campbell The Transition Phenomenon
Collecting together the evidence about the average fluid motion over L = 6 from
Figures 7.13 ( w < 0 ), 7.14 ( u faster in the channel midline), and 7.15 ( v directed
away from the midline in both directions) suggests a persistent rotation in the
transverse plane. This fluid motion is sketched in Figure 7.16, together with
secondary cell patterns as inferred by Knight and Patel (1985) from bed shear stress
and velocity measurements for comparison. The fact that v begins to decrease for
y>25mm and y<-40mm [Figure 7.15(b)] indicates the possibility of further
secondary currents distributed across the channel, each one scaling with flow depth.
Hence, the 90mm wide measurement window for a flow depth of 55mm spans the
best part of two complete counter-rotating secondary cells. The observed direction of
rotation is entirely consistent with the rectangular duct flow results of Knight and
Patel (1985) for a flow with an even aspect ratio (as B/H 8).

The structure of secondary currents in straight rectangular ducts and open channels is
reasonably well understood (e.g. Knight et al., 1982; Knight and Patel, 1985; Nezu
and Nakagawa, 1993). In contrast, the fundamental mechanics underlying the
appearance of such strong lateral effects specifically at the d-k-type transitional
roughness for relatively high aspect ratios remains unclear. One may speculate that
the concentrated spike of vertical momentum at the leading edge of the bars in the
transitional spacing could play an important role. The strong vertical velocity must
be present across the entire flume width, as the square bars are laterally continuous.
However, it is reasonable to assume that the streamwise velocity of the main flow is
the driving force behind the vertical spike [refer to Figure 7.1(b)], and that this
streamwise velocity will vary over the channel width as the side walls are
approached (i.e. u must decrease to zero for no slip at the flume walls). The logical
consequence is that the strength of the vertical velocity spike will vary across the
channel; this may potentially be a generation mechanism for enhanced secondary
currents. Obviously, this suggested mechanism requires further experimental
investigation, and as yet it is difficult to reconcile enhanced secondary currents with
conditions of increased discharge. It is also true that the direction of rotation of the
secondary cells for the supplementary experiment is at odds with what one may
intuitively expect given the likely cross-channel variation of the leading edge vertical

PhD Thesis 200 Chapter 7.0


L J Campbell The Transition Phenomenon
velocity spike. As such, the effect of the transition phenomenon in terms of cross-
channel variation remains the central unanswered question evoked by this study.

7.5 Chapter Closure

This chapter was devoted to introducing, and providing a preliminary mechanistic


explanation of, the transition phenomenon. The phenomenon occurs when the inter-
bar gap is approximately equal to the natural reattachment length of the rough-bed
flow (~5k), and is characterised by a marked local disturbance in the near-bed flow.
This in turn exerts significant control over the turbulent structure of the flow such
that it may be considered to be partially re-laminarised. The most striking outcome
of the special flow behaviour at transition is the increase in channel conveyance
which is reflected in the decreased friction factor and Mannings n values for bar
spacing L = 6. In this respect it must be noted that the roughness Reynolds number
( Re* ) based on the roughness height, k, or the equivalent sand grain roughness, ks,
easily exceeded the threshold value for hydraulically rough flow ( Re* = 70, see
Section 2.2.1) in all cases ( Re* 175). However, it is possible that these values for
Re* may be misleading for certain square bar flows. For example, any dependence
of friction factor on Reynolds number would suggest the flow regime is transitional,
not fully rough (and that Re* has therefore been overestimated).

The literature reviewed in Chapter 2 showed that bar roughness has been extensively
studied (e.g. Wood & Antonia, 1975; Knight & MacDonald, 1979a,b; Tani, 1987;
Cui et al., 2000; Leonardi et al. 2003) and its general properties are well known.
However, its special behaviour at the dk-type roughness transition has not hitherto
been recognised. This is probably because the resulting phenomenon occurs for a
very particular roughness spacing and experiments or simulations focussing on either
side of the transition give no hint of the surprising behaviour that occurs (refer to
Table 2.1 for a list of bar spacings from the existing literature). As was noted at the
outset of this chapter, transitions often produce rich and unexpected behaviours and
this phenomenon is no exception. With an eye to the future, the potential
applications for the transitional bar spacing may be numerous. The ability to effect
significant changes to Mannings n by using fixed or mechanically deployable ribbed

PhD Thesis 201 Chapter 7.0


L J Campbell The Transition Phenomenon
roughness has much potential. Ducts, pipes and open channels, for example
ventilation shafts in industrial units, culverts and storm water drainage channels, oil,
gas and water supply pipelines, and canals and rivers, convey many fluids in both
our built and natural environments. Indeed, some of the above man-made conduits
are already manufactured with internal ribs for engineering design reasons (e.g. for
strengthening); potential energy savings could be realised if this rib roughness
geometry is optimised to reduce surface drag.

The evidence presented in this chapter for the mechanism underlying the transition
phenomenon is convincing. However, it must be remembered that this finding arises
from an extensive and highly complex dataset of flows over square bars, and is
subject to ongoing analysis. Whilst it has been possible to deduce information about
the turbulent structure and cross-channel effects from clues in the instantaneous,
temporally-averaged and double-averaged velocity fields, further experimental
results (most notably pertaining to the cross-channel direction) are undoubtedly
required to advance our understanding of the phenomenon.

PhD Thesis 202 Chapter 7.0


L J Campbell The Transition Phenomenon
8.0 CONCLUSION

Much research effort has been invested in clarifying the velocity and turbulence
characteristics of hydraulically rough open-channel flows at levels beyond the direct
influence of the bed topography (i.e. above the roughness layer). The majority of this
analysis has been founded on the temporally-averaged momentum equations (the
Reynolds-Averaged Navier-Stokes equations, or RANS). However, the structure of
shallow open-channel flows (i.e. those in which the roughness height is of the same
order of magnitude as the flow depth) over hydraulically rough beds, together with
the near-bed region of deeper open-channel flows with rough beds awaits
clarification. This is because in these regions the effect of surface roughness on the
flow properties cannot be neglected, so that the information gained from a RANS
approach is dependent on the measurement location relative to the bed topography
(e.g. whether upstream or downstream of an obstacle). To overcome this difficulty,
recent advances in the analysis of such flows have suggested that temporal averaging
of the momentum equations should be supplemented with spatial averaging (in
planes or volumes orientated parallel to the mean bed), thus yielding the double-
averaged momentum equations. The literature review presented in this thesis has
identified the need for more detailed (spatial) measurements of velocity in the
roughness layer specifically geared towards assessing the double-averaged flow
equations for hydraulically rough, shallow open-channel flows.

------------------------------------------------------------------------------------------------------

As stated in Chapter 1 of this thesis, the objectives of the project were,

1) to conduct a major experimental program to further our understanding


of the fundamental mechanics underlying the interaction between fluid
flow and a rough surface; this involves detailed measurements of
temporally, and spatially varying fluid velocities;

PhD Thesis 203 Chapter 8.0


L J Campbell Conclusion
2) to evaluate the general applicability of the double-averaged equations
for describing flow over a range of relative roughnesses and roughness
configurations;

3) to test specific relationships derived by double-averaging for the near-


bed velocity distribution over a range of relative roughnesses and
roughness types;

4) to quantify the extent of non-uniformity in the near-bed flow for a range


of relative roughnesses and roughness configurations;

5) to establish an extensive dataset of detailed velocity measurements for


steady, uniform open-channel flows over geometrically-simple bed
roughnesses, which may be readily transferred to other researchers
concerned with the development and benchmarking of numerical
models.

------------------------------------------------------------------------------------------------------

Proper assessment of the double-averaged equations for rough open-channel flow


requires comprehensive measurement of fluid velocities as they vary in both time
and space. As such, one of the main achievements of the PhD is the creation of a
comprehensive dataset for shallow open-channel flows over simple rough beds
(Objective 5). The dataset comprises experiments with three flow depths, all of small
relative submergence, over square bar strip roughness covering a wide range of near-
bed flow patterns, from skimming or d-type flow, to isolated roughness or k-type
flow. In all cases, comprehensive and synchronous measurements have been
obtained of the streamwise and bed-normal velocity components, including their
variation in both temporal and spatial domains over a representative cross-section
orientated along the centre-line of the flume. The level of detail in the near-bed
velocity measurements, especially recording their variation in both time and space,
enabled the accurate evaluation of individual terms in the double-averaged equations
for open-channel flows (Objective 1). Analysis of this new data in the context of
double averaging has led to the following main conclusions.

PhD Thesis 204 Chapter 8.0


L J Campbell Conclusion
Double-Averaged Velocity and Fluid Stress Profiles (Chapter 5): Objectives 2 and 3

To date the description of velocity and fluid stress profiles has largely relied on
temporal averaging of the flow field, and hence the precise nature of such profiles in
the vicinity of bed roughness has remained unclear. The use of PIV to gain good
spatial resolution over the full flow depth made it possible to investigate double-
averaged velocity and stress distributions, including the form-induced stress. The
findings from this stage of the analysis were:

1) In the gaps between successive square bars, areas of recirculation mean that the
bed-normal distribution of double-averaged streamwise velocity, u ( z ) , is S-
shaped for skimming and wake interference flows.
2) As the bar pitch is extended and the bars become increasingly isolated, the
streamwise velocity profile becomes increasingly linear in the interfacial sub-
layer (and becomes linear for L 15). This was explained in terms of the balance
between S-shaped and semi-smooth velocity profiles, generated from areas
dominated by separated and reattached flow respectively.
3) Above either the S-shaped or linear profile, the streamwise velocity distribution
shows good agreement with the logarithmic law, even in areas of the water
column dynamically influenced by the square bar roughness. The variation of
roughness lengths and displacement heights with bar spacing in most cases make
sound physical sense. However, large displacement heights for isolated
roughness arrangements together with a broad range of the von Karman constant
suggest that the log-law may not be valid for such shallow open-channel flows.
4) Form-induced stress cannot be neglected as it contributes up to 20% of the total
measured (turbulent plus form-induced) stress. Peak form-induced stress
changed from positive to negative as roughness pitch increased, whilst the
distribution remained entirely complementary to the Reynolds stress. The peak in
the form-induced stress profiles was always restricted to the level of the
roughness crests (typically up to z = 2k).
5) The extent of roughness layer was tentatively estimated as between 2.5k-5k, with
greater thicknesses linked to isolated roughness conditions. This estimate was
based on observation of the spatial variation in both (temporal) mean streamwise

PhD Thesis 205 Chapter 8.0


L J Campbell Conclusion
velocity and stress profiles, and on charting the level at which form-induced
stress became negligible.

Spatial Variation within the Roughness Layer (Chapter 6): Objective 4

Double averaging enabled the general representative characteristics of flow close to


square bar-roughened boundaries to be ascertained. Additionally, the fundamental
characteristics of how the flow interacted with the bed roughness were clarified by
subtracting the double-averaged profiles from the temporally-averaged flow field.
This analysis led to the following conclusions:

1) The form-induced velocity components, u% and w% behave periodically, and are at


equivalent points of their cycles a quarter of a roughness pitch apart (i.e. they are
phase shifted by = 0.25p).
2) The form-induced velocity components may still be significant even in areas
where their spatial average, the form-induced stress, is negligible. The roughness
layer therefore comprises two distinct zones. Fluid flow in both zones is
dynamically influenced by the presence of the bed roughness, but whereas the
lower zone exhibits significant form-induced stress, the upper one does not.
3) A new application of quadrant analysis for the spatial velocity fluctuations is a
useful tool for visualising the mean flow structure over rough beds. In the region
of the roughness layer characterised by negligible form-induced stress all four
quadrants are equally represented and the quadrant diagrams trace rings (the
Zone of the Rings). Associated quadrant maps are vertically banded. In the
lower part of the roughness layer, the rings become distorted (the Zone of
Irregular Orbits) and form-induced stress is not negligible.
4) On the basis of charting the disintegration of the vertical banding of spatial
quadrant maps, the extent the roughness layer was re-evaluated and found to be
up to z = 7k from the channel bed for isolated roughness or k-type arrangements.

The Influence of Bar Spacing: Transitional Effects (Chapter 7)

The approach adopted in this study of systematically altering bar spacing has led to
the discovery of a transition phenomenon, an instability at the d-type to k-type

PhD Thesis 206 Chapter 8.0


L J Campbell Conclusion
roughness transition (L = 6). Although the precise mechanisms underlying the
phenomenon require additional clarification, the following preliminary explanations
were presented:

1) For bar spacing L = 6 the reattachment point occurs on the upstream vertical face
of each square bar. This induces a strong vertical velocity spike focussed at the
leading edge of each bar, which in turn generates a region of counter-vorticity (as
seen in the form-induced velocity field).
2) There is evidence that the turbulent structure is altered at the d-k-type roughness
transition by a marked reduction in the slope coefficient, a, of the temporal
quadrant plots (i.e. w = au ).
3) The most striking result of the transition phenomenon is drag adjustment, such
that peaks in channel discharge occur at bar spacing L = 6.

8.1 Thesis Closure

This thesis has addressed the applicability of the double-averaged equations for the
analysis of fluid flow over rough beds. Detailed velocity measurements of shallow
open-channel flows over a wide range of square bar configurations have been
analysed and presented. The results form a comprehensive dataset that may be used
for the development and validation of numerical models over simple roughness
types. The presented analysis demonstrates that double-averaging is a powerful tool
for the analysis of hydraulically rough open-channel flows and significantly
advances understanding of the fundamental mechanisms underlying the interaction
of a fluid flow and a rough surface.

PhD Thesis 207 Chapter 8.0


L J Campbell Conclusion
REFERENCES

Adachi, S. (1964), On the artificial strip roughness, Bulletin No. 69, Disaster
Prevention Research Institute, Kyoto University, Tokyo, Japan
Antonia, R. A. & Luxton, R. E. (1971), The response of a turbulent boundary layer
to an upstanding step change in surface roughness, Transactions of the ASME
Journal of Basic Engineering, 22-34
Ayotte, K. W., Finnigan, J. J. & Raupach, M. R. (1999), A second-order closure
for neutrally stratified vegetative canopy flows, Boundary Layer Meteorology, 90,
189-216
Baker, J., Myatt, J., & Christofides, P. D. (2002), Drag reduction in flow over a
flat plate using active feedback control, Comp. & Chem. Engrg., 26, 1095-1102
Bandyopadhyay, P. R. (1986), Drag reducing outer-layer devices in rough wall
turbulent boundary layers, Experiments in Fluids, 4, 247-256
Brandyopadhyay, P. R. (1987), Rough-wall turbulent boundary layers in the
transition regime, J. Fluid Mech., 180, 231-266
Barrantes, A. I. & Madsen, O. S. (2000), Near-bottom flow and flow resistance for
currents obliquely incident to two-dimensional roughness elements, J. Geophysical
Res., 105, No. C11, 26253-26264
Bechert, D. W., Bruse, M., Hage, W., Van Der Hoeven, J. G. T., & Hoppe, G.
(1997), Experiments on drag-reducing surfaces and their optimization with an
adjustable geometry, J. Fluid Mech. 338, 59-87
Berger, F. P., Berger, K.-F. & Hau, F.-L. (1979), Local mass/heat transfer
distribution on surfaces roughened with small square ribs, Int. J. Heat Mass
Transfer, 22, 1645-1656
Berman, N. S. (1978), Drag reduction by polymers. Annual Rev. Fluid Mech. 10,
47-64
Bhm, M., Finnigan, J. J. & Raupach, M. R. (2000), Dispersive fluxes and canopy
flows: just how important are they?, American Meteorology Soc., 24th Conference on
Agriculture and Forest Meteorology, 106-107
Brunet, Y., Finnigan, J. J. & Raupach, M. R. (1994), A wind tunnel study of air
flow in waving wheat: single point velocity measurements, Boundary Layer
Meteorology, 70, 95-132
Campbell, L. J., McEwan, I., Nikora, V., Pokrajac, D., Gallagher, M. & Manes,
C. (2005), Bed-load effects on hydrodynamics of rough-bed open-channel flows, J.
Hydr. Engrg., 131, 7, 576-585
Campbell, L. J., McEwan, I., Pokrajac, D., Manes, C. & Lazic, R. (2005),
Passive drag adjustment by very rough surfaces, accepted by J. Hydr. Engrg.
Chadwick, A. & Morfett, J. (1998), Hydraulics in Civil and Environmental
Engineering, Third edition, E & FN Spon, London, UK
Cheng, H. & Castro, I. P. (2002), Near wall flow over urban-like roughness,
Boundary Layer Meteorology, 104, 229-259

PhD Thesis 208 REFERENCES


L J Campbell
Choi, K-S. & Fujisawa, N. (1993), Possibility of drag reduction using d-type
roughness, Appl. Scientific Res. 50, 315-324
Clauser, F. H. (1954), Turbulent boundary layers in adverse pressure gradients, J.
Aero. Sci., 21, 91-105
Coleman, N. (1981), Velocity profiles with suspended sediment, J. Hydr. Res., 19,
3, 211-229
Coles, D. (1956), The law of wake in the turbulent boundary layer, J. Fluid Mech., 1,
15-29
Crapiste, G. H., Rotstein, E. & Whitaker, S. (1986), A general closure scheme for
the method of volume averaging, Chem. Eng. Sci., 41, 2, 227-235
Cui, J., Patel, V. C. & Lin, C.-L. (2000), Large-eddy simulation of turbulent flow
over rough surfaces. IIHR Technical Report No. 413, Iowa Institute of Hydraulic
Research
Cui, J., Patel, V. C. & Lin, C.-L. (2003), Large-eddy simulation of turbulent flow
in a channel with rib roughness, Int. J. Heat & Fluid Flow, 24, 372-388
Cunningham, G. (2000), Predicting entrainment of mixed size sediment grains by
probabilistic methods, PhD thesis, University of Aberdeen, Scotland, UK
Dittrich, A. & Koll, K. (1997), Velocity field and resistance of flow over rough
surfaces with large and small relative submergence, Int. J. Sediment Res., 12, No 3,
21-33
Djenidi, L., Anselmet, F. & Antonia, R. A. (1994), LDA measurements in a
turbulent boundary layer over a d-type rough wall, Experiments in Fluids, 16, 323-
329
Djenidi, L., Elavarasan, R. & Antonia, R. A. (1999), The turbulent boundary layer
over transverse square cavities, J. Fluid Mech., 395, 271-294
Engineering Sciences Data Unit Item Number 79014 (1979), Losses caused by
friction in straight pipes with systematic roughness elements, ESDU, London
Finnigan, J. J. (1983), Turbulent heat and mass transport in flexible plant canopies,
8th Australian Fluid Mechanics Conference, University of Newcastle, New South
Wales, 1A.9-1A.13
Finnigan, J. J. (1985), Turbulent transport in flexible plant canopies, The Forest-
Atmosphere Interactions, B. A. Hutchinson & B. B. Hicks (eds.) Reidel, Dordrecht,
The Netherlands, 443-480
Finnigan, J. J. (2000), Turbulence in plant canopies, Ann. Rev. Fluid Mech., 32,
519-571
Garrat, J. R. (1980), Surface influence upon vertical profiles in the atmospheric
near surface layer, Quart. J. Roy Meteorol. Soc., 104, 199-212
Gimenez-Curto, L. A., & Corniero Lera, M. A. (1996), Oscillating turbulent flow
over very rough surfaces, J. Geophys. Res., 101 (C9), 20745-20758
Graf W. H. (1998), Fluvial Hydraulics: Flow and Transport Processes in Channels
of Simple Geometry, John Wiley & Sons Chichester, UK

PhD Thesis 209 REFERENCES


L J Campbell
Grass, A. J. (1971), Structural features of turbulent flow over smooth and rough
boundaries, J. Fluid Mech., 50, 2, 233-255
Grass, A. J., Stuart, R. J. & Mansour-Tehrani, M. (1991), Vortical structures and
coherent motion in turbulent flows over smooth and rough boundaries, Royal Soc.,
Phil. Transactions: Phys. Sciences & Eng., 336, No. 1640, Part 1, 35-65
Gray, W. G. (1975), A derivation of the equations for multiphase transport, Chem.
Eng. Sci., 30, 229-233
Gray, W. G. & Lee, P. C. Y. (1977), On the theorems for local volume averaging of
multiphase systems, Int. J. Multiphase Flow, 3, 333-340
Islam, O. & Logan, E. (1976), Channel flow over a smooth-to-rough surface
discontinuity with zero pressure gradient, Transactions of the ASME Journal of
Fluids Engineering, 626-634
Jacobson, S. A. & Reynolds, W. C. (1998), Active control of streamwise vortices
and streaks in boundary layers. J. Fluid Mech., 360, 179-211
Jimenez, J. (2004), Turbulent flows over rough walls. Ann. Rev. Fluid Mech., 36,
173-196
Kanda, M., Moriwaki, R. & Kasamatsu, F. (2004), Large-eddy simulation of
turbulent organized structures within and above explicitly resolved cube arrays,
Boundary Layer Meteorology, 112, 343-368
Keane, R. D. & Adrian, R. J. (1992), Theory of cross-correlation analysis of PIV
images, Appl. Sci. Res., 49, 191-215
Keirsbulck, L., Labraga, L., Mazouz, A. & Tournier, C. (2002), Surface
roughness effects on turbulent boundary layer structures, J. Fluids Eng., 124, 127-
135
Knight, D. & MacDonald, J. A. (1979a), Hydraulic resistance of artificial strip
roughness, J. Hydr. Div., ASCE, 105, HY6, 675-690
Knight, D. (1981), Boundary shear in smooth and rough channels, J. Hydr. Div.,
ASCE, 107, HY7, 839-851
Knight, D. & MacDonald, J. A. (1979b), Open channel flow with varying bed
roughness, J. Hydr. Div., ASCE, 105, HY9, 1167-1183
Knight, D. & Patel, H. S. (1985), Rectangular ducts, J. Hydr. Eng., 111, No. 1, 29-
46
Knight, D. W., Patel, H. S., Demetriou, J. D. & Hamed, M. E. (1982), Boundary
shear stress distributions in open channel and closed conduit flows, Euromech 156,
Mechanics of Sediment Transport, Istanbul, 33-40
Leonardi, S., Orlandi, P., Smalley, R. J., Djenidi, L., & Antonia, R. A. (2003),
Direct numerical simulations of turbulent channel flow with transverse square bars
on one wall. J. Fluid Mech. 491, 229-238
Lien, F.-S. & Yee, E. (2004), Numerical modelling of the turbulent flow developing
within and over a 3-D building array, Part 1: A high-resolution Reynolds-averaged
Navier-Stokes approach, Boundary Layer Meteorology, 112, 427-466
Lien, F.-S. , Yee, E. & Wilson, J. D. (2005), Numerical modelling of the turbulent
flow developing within and over a 3-D building array, Part 2: A mathematical

PhD Thesis 210 REFERENCES


L J Campbell
foundation for a distributed drag force approach, Boundary Layer Meteorology, 114,
245-285
Lien, F.-S. & Yee, E. (2005), Numerical modelling of the turbulent flow developing
within and over a 3-D building array, Part 3: A distributed drag force approach, its
implementation and application, Boundary Layer Meteorology, 114, 287-313
Liou, T.-M., Hwang, J.-J. & Chen, S.-H. (1993a), Simulation and measurement of
enhanced turbulent heat transfer in a channel with periodic ribs on one principal
wall, Int. J. Heat Mass Transfer, 36, No. 2, 507-517
Liou, T.-M., Wu, Y.-Y & Chang, Y. (1993b), LDV measurements of periodic fully
developed main and secondary flows in a channel with rib disturbed walls, J. Fluids
Eng., 115, 109-114
Liu, C. K., Kline, S. J., & Johnston, J. P. (1966), An experimental study of
turbulent boundary layer on rough walls, Report MD-15, Department of Mechanical
Engineering, Stanford University
Lohsz, M., Rambaud, P. & Benocci, C. (2003), LES stimulation of ribbed square
duct flow with Fluent and comparison with PIV data, Conference on Modelling
Fluid Flow, 12th Int. Conference on Fluid Flow Technologies, Budapest, Hungary
Lpez, F. & Garca, M. H. (2001), Mean flow and turbulence structure of open-
channel flow through non-emergent vegetation, J. Hydr. Eng., 127, No. 5, 392-402
Lu, S. S. & Willmarth, W. W. (1973), Measurements of the structure of the
Reynolds stress in a turbulent boundary layer, J. Fluid Mech., 60, 481-511
Lumley, J. & Blossey, P. (1998), Control of turbulence, Annual Rev. Fluid Mech.,
30, 311-327
Lumley, J. L. (1969), Drag reduction by additives, Annual Rev. Fluid Mech., 1, 367-
384
MacDonald, R. W. (2000), Modelling the mean velocity profile in the urban canopy
layer, Boundary Layer Meteorology, 97, 25-45
Maddux, T. B., Nelson, J. M. & McLean, S. R. (2003), Turbulent flow over three-
dimensional dunes: 1. Free surface and flow response, J. Geophysical Res., 108, F1,
10-1 10-19
Massey, B. S. (1990), Mechanics of Fluids, 6th edition, Chapman & Hall, London,
UK
Matsumoto, A. (1993), Some features of turbulent boundary layers over grooved
rough walls. Trans. Japan Soc. Aero. Space Sci., 37, 115, 27-41
McLean, S. R., Wolfe, S. R. & Nelson, J. M. (1999), Spatially averaged flow over
a wavy boundary revisited, J. Geophysical Res., 104, C7, 15743-15753
Millikan, C. M. (1938), A critical discussion of turbulent flows in channels and
circular tubes, Proc. 5th Int. Congress Appl. Mech., Wiley, New York, USA, 386-392
Monin, A. S. & Yaglom, A. M. (1977), Statistical Fluid Mechanics: Mechanics of
Turbulence Volume 1, MIT Press, Cambridge, Massacheusetts, USA
Morris, H. M. (1955), Flow in rough conduits, Trans. ASCE, Paper 2745, 373-410

PhD Thesis 211 REFERENCES


L J Campbell
Mulhearn, P. J. (1978), Turbulent flow over a periodic rough surface, Phys. Fluids,
21, 1113-1115
Mulhearn, P. J. & Finnigan, J. J. (1978), Turbulent flow over a very rough random
surface, Boundary Layer Meteorology, 15, 109-132
Nezu, I. & Nakagawa, H. (1993), Turbulence in Open Channel Flow, Edited by A.
A. Balkema, IAHR, Rotterdam, The Netherlands
Nikora, V., Goring, D., McEwan, I., Griffiths, G. (2001), Spatially-averaged open-
channel flow over rough bed, J. Hydr. Eng., 127, 2, 123-133
Nikora, V., Koll, K., McLean, S., Dittrich, A. & Aberle, J. (2002), Zero-plane
displacement for rough-bed open-channel flows, River Flow 2002, Proc. Int. Conf.
on Fluvial Hydraulics, Louvain-la-Neuve, Belgium
Nikora,V., Koll, K., McEwan, I., McLean, S. & Dittrich, A. (2004), Velocity
distribution in the roughness layer of rough-bed flows, J. Hydr. Engrg., 130, No. 10,
1036-1042
Okamoto, S., Seo, S., Nakaso, K. & Kawai, I. (1993), Turbulent shear flow and
heat transfer over the repeated two-dimensional square ribs on ground plane, J.
Fluids Eng., 115, 631-637
Perry, A. E. & Joubert, P. N. (1963), Rough-wall boundary layers in adverse
pressure gradients, J. Fluid Mech., 17, 193-211
Perry, A. E., Schofield, W. H. & Joubert, P. N. (1969), Rough wall turbulent
boundary layers J. Fluid Mech., 37 (2), 383-413
Poggi, D., Porporato, A., Ridolfi, L., Albertson, J. D. & Katul, G. G. (2004a),
The effect of vegetation density on canopy sub-layer turbulence, Boundary Layer
Meteorology, 111, 565-587
Poggi, D., Katul, G. G. & Albertson, J. D. (2004b), Momentum transfer and
turbulent kinetic energy budgets within a dense model canopy, Boundary Layer
Metereology, 111, 589-614
Poggi, D., Katul, G. G. & Albertson, J. D. (2004c), A note on the contribution of
dispersive canopy fluxes to momentum transfer within canopies, Boundary Layer
Metereology, 111, 615-621
Pokrajac, D., Campbell, L. J., Nikora, V. I., McEwan, I. K. & Manes, C. (2005),
Spatially-averaged flow over artificial roughness: a new application of quadrant
analysis, submitted to J. Fluid Mech.
Raffel, M., Willert, C. E. & Kompenhans, J. (1998), Particle Image Velocimetry,
A Practical Guide, Springer-Verlag, Berlin, Germany
Raju, K. G. R. & Garde, R. J. (1970), Resistance to flow over two-dimensional
strip roughness, J. Hydr. Div., Proc. ASCE, 96, HY3, 815-834
Rathnasingham, R. & Breuer, K.S. (1997a), Coupled fluid-structural
characteristics of actuators for flow control. AIAA J., 35 (5), 832-837
Rathnasingham, R. & Breuer, K.S. (1997b), System identification and control of a
turbulent boundary layer. Phys. Fluids, 9 (7), 1867-1869
Raupach, M. R. & Shaw, R. H. (1982), Averaging procedures for flow within
vegetation canopies, Boundary Layer Meteorology, 22, 79-90

PhD Thesis 212 REFERENCES


L J Campbell
Raupach, M. R., & Thom, A. S. (1981), Turbulence in and above plant canopies,
Ann. Rev. Fluid Mech., 13, 97-129
Raupach, M. R., Thom, A. S. & Edwards, I. (1980), A wind-tunnel study of
turbulent flow close to regularly arrayed rough surfaces, Boundary Layer
Meteorology, 18, 373-397
Raupach, M. R., Coppin, P. A., & Legg, B. J. (1986), Experiments on scalar
dispersion within a plant canopy 1, the turbulence structure, Boundary Layer
Meteorology, 35, 21-52
Raupach, M. R., Antonia, R. A. & Rajagopalan, S. (1991), Rough-wall turbulent
boundary layers, ASME Appl. Mech. Rev., 44 (1), 1-25
Rotach, M. W. (1993), Turbulence close to a rough urban surface Part 1: Reynolds
stress, Boundary Layer Meteorology, 65, 1-28
Sayre, W.W & Albertson, M. L. (1961), Roughness spacing in rigid open channels,
J. Hydr. Div., ASCE, 87, HY3, 121-150
Sayre, W.W & Albertson, M. L. (1963), Roughness spacing in rigid open channels,
Transactions ASCE, 128, 343-427
Schlichting, H. (1979), Boundary-Layer Theory, 7th edition, McGraw-Hill, New
York
Sirovich, L. & Karlsson, S. (1997), Turbulent drag reduction by passive
mechanisms, Nature, 388, 753-755
Siuru, W. D. Jr. & Logan, E. Jr. (1977), Response of a turbulent pipe flow to a
change in roughness, Trans. ASME, J. Fluids Eng., 99, 548-553
Slattery, J. C. (1967), Flow of viscoelastic fluids through porous media, Am. Inst.
Chem. Eng. J., 13, 1066-1071
Smith, J. D., & McLean, S. R. (1977), Spatially-averaged flow over a wavy
surface, J. Geophys. Res., 83 (12), 1735-1746
Song, T., Graf, W. H. & Lemmin, U. (1994), Uniform flow in open channels with
movable gravel bed, IAHR J. Hydr. Res., 32 (6), 861-876
Stoesser, T. & Rodi, W. (2004), LES of bar and rod roughened channel flow, 6th
Int. Conf. on Hydrosci. and Eng. (ICHE-2004), May 30 June 3, Brisbane, Australia
Tani, I. (1987), Turbulent boundary layer development over rough surfaces, In
Perspectives in Turbulence Studies (ed. H. U. Meier & P Bradshaw), 223-249
Tardu, S. F. Active control of near-wall turbulence by local oscillating blowing. J.
Fluid Mech., 439, 217-253 (2001)
Mannings n Coefficients for Open-Channel Flow: The Fluid Mechanics Calculation
Website http://www.lmnoeng.com/manningn.htm (2000)
Thom, A. S. (1971), Momentum absorption by vegetation, Quart. J. Royal
Meteorological Soc., 97, 414-428
Townes, H. W. & Sabersky, R. H. (1966), Experiments on the flow over a rough
surface, Int. J. Heat Mass Transfer, 9, 729-738
Walsh, M. J. (1990), in Progress in Astronautics and Aeronautics Vol. 123 (eds.
Bushnell, D. & Hefner, J.). AIAA, Reston, VA, USA

PhD Thesis 213 REFERENCES


L J Campbell
Warholic, M. D., Massah, H., & Hanratty, T. J. (1999), Influence of drag-
reducing polymers on turbulence: effects of Reynolds number, concentration and
mixing, Experiments in Fluids. 27, 461-472
Webb, R. L., Eckert, E. R. G. & Goldstein, R. J. (1971), Heat transfer and friction
in tubes with repeated-rib roughness, Int. J. Heat Mass Transfer, 14, 601-617
Westerweel, J. (1994), Efficient detection of spurious vectors in particle image
velocimetry data, Exper. Fluids, 16, 236 247
Whitaker, S. (1967), Diffusion and dispersion in porous media, Am. Inst. Chem.
Eng. J., 13, 420-438
Wilson, N. R., & Shaw, R. H. (1977), A higher order closure model for canopy
flow, J. Appl. Meteor., 16, 1198-1205
Wood, D. H. & Antonia, R. A. (1975), Measurements in a turbulent boundary layer
over a d-type surface roughness, J. Appl. Mech., 42, 591-597
Xie, Z., Voke, P. R., Hayden, P. & Robins, A. G. (2004), Large-eddy simulation of
turbulent flow over a rough surface, Boundary Layer Meteorology, 111, 417-440
Xu, J., Maxey, M. R., & Karniadakis, G. E. (2002), Numerical simulation of
turbulent drag reduction using micro-bubbles, J. Fluid Mech. 468, 271-281
Yaglom, A. M. (1979), Similarity laws for constant-pressure and pressure-gradient
turbulent wall flows, Ann. Rev. Fluid Mech., 11, 505-540
Yalin, M. S. (1977), Mechanics of Sediment Transport, 2nd edition, Pergamon Press,
Oxford, UK
Zippe, H. J., & Graf, W. H. (1983), Turbulent boundary layer flow over permeable
and non-permeable rough surface, J. Hydr. Res., 21 (1), 300-316

PhD Thesis 214 REFERENCES


L J Campbell
APPENDIX A: PROOF OF INTEGRAL THEOREM A

[The Spatial Averaging Theorem (e.g. Whitaker, 1967; Slattery, 1967), as


given in Section 3.4.3, Equation (3.11)]

1 1
Proof for:
xi
=
xi

Vf n dS
Sint
i (A.1)

Where is some tensorial quantity, vector or scalar, defined only in fluid (note also
that V must be additive over the averaging volume); Sint is the solid-fluid
interface surface; ni is the component of the unit normal vector in the i-direction;
angle brackets denote volume averaging, and i = 1,2,3, where x1,x2,x3 are the
coordinate axes of the Cartesian coordinate system.

Consider an averaging volume intersected by solid surfaces (e.g. sediment grains):

n Vf = volume of fluid
in averaging region
(Vo)
n Sff: Fluid-Fluid
interface
Vs = volume of
nfs solids in averaging
Vo: Total averaging region (Vo)
region (volume)
Vo = Vf + Vs

Vf
=
Sfs: Fluid-Solid Vo
interface
nfs = unit normal
vector pointing into
fluid from solid
Sss: Solid-Solid
interface surfaces

Figure A.1: Definition sketch of averaging n = outward unit


volume (or area) showing different bounding normal vector
surfaces between solid and fluid portions.
Includes unit normal vectors (n) used with pointing out from
Gauss Theorem fluid-fluid interface

Superficial average, t , considered over whole volume, Vo


Intrinsic average, , correctly considered in fluid, Vf, only

Averages related by: t =

PhD Thesis 215 APPENDIX A


L J Campbell Proof of Integral Theorem A
(The proof continues by using the superficial average, and converts to the intrinsic
average at the end).

By definition of the volume averaging operator,

1
t = dV Vo t = dV = dV (A.2)
Vo Vo Vo Vf

(as is defined within the fluid portion of Vo only)

Similarly, for the spatial derivative,


Vo
xi
= x dV = x dV + x dV = x dV (A.3)
t Vo i Vf i Vs i Vf i


Gauss Theorem (i.e. x dV = n dS ) is then implemented to switch between
V S
volume and surface integrals. Application (to the fluid volume only) gives:


Vo
xi
= x V = ni dS ni dS (A.4)
t Vf i S ff S fs

The negative sign in front of the second surface integral arises because the unit
normal vector, nfs, points into the fluid, but Gauss Theorem as defined above is for
an outward pointing unit normal. The first surface integral has positive sign, as n
points from the fluid. Both unit normals for surfaces Sfs and Sff are defined in Figure
A.1.

Note that the surface integral over Sfs is in fact the integral over the fluid-solid
interface. Hence this term is complete in terms of the Theorem under
consideration; Sfs = Sint.

For the fluid-fluid surface integral, Sff, term, we first consider the definition of a
derivative:

lim xi +xi xi
= (A.5)
xi xi 0 xi

PhD Thesis 216 APPENDIX A


L J Campbell Proof of Integral Theorem A
Applying (A.5) to

Vo t
, with Vo t = dV [i.e. (A.2)] (A.6)
xi Vf

gives,

dVxi +xi dVxi


Vo t lim Vf Vf
= (A.7)
xi xi 0 xi

For some small shift in space, xi , the averaging volume is transformed as in Figure
A.2:

Vf- Vfo Vf+


= averaging region, Vo, at
position xi = Vfo + Vf-

= averaging region, Vo, at


position xi + xi = Vfo + Vf-

Vf- = volume of fluid left behind


after small shift in space

Vfo = shared volume between old


and new positions
xi
Vf+ = new portion of volume
n resulting from small shift in
position

Figure A.2: Definition sketch of averaging n = outward pointing unit normal


volume (or area) before and after small shift vector
in position (xi).

Substituting the volume integrals based on Figure A.2 into Equation (A.7) gives,

Vo t lim 1
dV + dV dV dV
xi 0 xi V
=
xi
f+ V fo V fo Vf

PhD Thesis 217 APPENDIX A


L J Campbell Proof of Integral Theorem A
lim1
dV dV
xi 0 xi V
= (A.8)
f+ Vf

Looking at the geometry of the new volume, Vf+, with reference to Figure A.3 which
focuses on a small portion of the surface, dS:

l
ni = n cos = cos and cos = l = ni dS (A.9)
dS

dS n n = outward
dVf+ = dS ni xi
pointing unit

dS ni
normal, n = 1
l ni
ni = component of
unit normal in ith
direction
xi
Figure A.3: Geometry of a small portion of the surface, as shown in A.2

Hence, once again this enables the move from a volume to surface integral as,

dV = xi ni dS (A.10)
Vf + Sf+

Obviously, this result also holds for the Vf- volume in Figure A.2, and as the
direction of the outward pointing unit normal takes care of sign (i.e. ni < 0 ), the
surface integrals for Sf- and Sf+ can be collated into one surface integral (Sff) term,
which in the limit gives:

Vo t lim 1
xi ni dS xi ni dS =
xi 0 xi Sf +
= + ni dS
xi
Sf S ff

(A.11)

Looking back to Equation (A.4) we had:


Vo
xi
= ni dS ni dS [(A.4)]
t S ff S fs

PhD Thesis 218 APPENDIX A


L J Campbell Proof of Integral Theorem A
Substituting for the Sff surface integral from Equation (A.11) into Equation (A.4),
and noting as before that Sfs is in fact just Sint yields,

Vo t t
Vo
xi
=
xi
ni dS = Vo
xi
ni dS (A.12)
t S int S int

Finally, convert (A.12) from the superficial to the intrinsic average using:

Vf Vf
t = ; = ; Vo = (A.13)
Vo

To give,

Vf V f
= + ni dS
xi xi Sint

1 1

xi
=
xi
+
Vf ni dS
Sint

The Spatial Averaging Theorem, (A.1)

PhD Thesis 219 APPENDIX A


L J Campbell Proof of Integral Theorem A
APPENDIX B: APPLICATION OF INTEGRAL THEOREMS A
and B TO MOMENTUM CONSERVATION and CONTINUITY
EQUATIONS

[To produce the double (volume)-averaged continuity equation as given in


Section 3.4.4, Equation (3.14), and the double-averaged momentum equations,
as given in Sections 3.4.2 and 3.4.3, Equation (3.9) and (3.13), respectively,]

Part 1 Averaging Rules and Theorems


Part 2 Mass Conservation (Continuity) Equation
Part 3 Momentum Conservation Equation
Part 4 Supplementary Working for Equations (3.9) and (3.13)

PART 1. AVERAGING RULES and THEOREMS

AVERAGING RULES

Consider two variables, f and g. The Reynolds Averaging Rules (e.g. Monin and
Yaglom, 1977) state that:

f + g = f + g; f = f ; f g= f g (B.1)

Using these rules, further expressions may be simply derived, e.g. f g = f g + f g


(where f = f + f ; g = g + g ).

THEOREMS

Double-averaging the continuity and Reynolds Averaged Navier-Stokes (RANS)


equations requires the following theorems (note that Theorem A was proved in
Appendix A):

Integral theorem A, The Spatial Averaging Theorem (Eq. (3.11), Section 3.4.3):

1 1
xi
=
xi

Vf n dS
Sint
i (B.2)

PhD Thesis 220 APPENDIX B


L J Campbell Application of Integral Theorem A
Integral theorem B, The General Transport Equation (Eq. (3.12), Section 3.4.3):

1 1
t
=
t
+
Vf v n
Sint
j j dS (B.3)

PART 2. MASS CONSERVATION (CONTINUITY) EQUATION

Beginning with the time-averaged continuity equation for an incompressible fluid,

ui
=0 (B.4)
xi

Development of the volume-averaged continuity equation follows by straightforward


application of integral Theorem A, which gives,

ui 1 ui 1
xi
=
xi

Vf u n dS = 0
Sint
i i (B.5)

1 ui 1
i.e.
xi
=
Vf u n dS
Sint
i i (B.6)

For frozen roughness,

1 1 ui
Vf u n dS = 0
Sint
i i , therefore
xi
=0

ui
=0 (B.7)
xi

PhD Thesis 221 APPENDIX B


L J Campbell Application of Integral Theorem A
PART 3. MOMENTUM CONSERVATION EQUATION

The appropriate integral theorem, A or B, will be applied to each term of the


Reynolds Averaged Navier-Stokes (RANS) equation in turn (Nikora, 2004), i.e.,

ui ui 1p ui uiuj
+ uj = gi + (B.8)
t xj xi x j x j xj

Averaging individual terms,

ui
1. Local fluid acceleration, (applying Theorem B, with ui ):
t

ui 1 ui 1
t
=
t
+
Vf u
Sint
i v j n j dS (B.9)

ui
2. Convective fluid acceleration, u j (applying Theorem A, with ui u j ):
xj

ui u j 1 ui u j 1
xj
=
xj

Vf u u n dS
Sint
i j j (B.10)

Then substitute,

ui u j = ( ui + u%i ) ( u j + u% j ) (B.11)

and, using one of the Reynolds averaging conditions (refer to B.1),

a b = a b (B.12)

with the result from applying Theorem A to the continuity equation [see Equations
(B.5) and (B.6)],

PhD Thesis 222 APPENDIX B


L J Campbell Application of Integral Theorem A
1 u j 1
xj
=
Vf u n dS
Sint
j j (B.13)

and the chain rule for derivatives of products gives,

ui u j 1 ui u j 1
xj
=
xj

Vf u u n dS
Sint
i j j

1 ui u j 1 u%i u% j 1 ui u% j 1 u%i u j 1
=
xj
+
xj
+
xj
+
xj

Vf u u n dS
Sint
i j j

1 ui u j 1 u%i u% j 1
=
xj
+
xj

Vf u u n dS
Sint
i j j

ui u 1 u%i u% j 1
= uj
xj
+ i
Vf u n dS +
Sint
j j
xj

Vf u u n dS
Sint
i j j

(B.14)

3. Gravity, gi :

gi = g i (B.15)

p
4. Pressure gradient, (applying Theorem A, with p ):
xi

p 1 p 1
xi
=
xi

Vf pn dS
Sint
i (B.16)

ui u
5. Viscous shear term, (applying Theorem A, with i ):
xj xj x j

PhD Thesis 223 APPENDIX B


L J Campbell Application of Integral Theorem A
ui 1 u 1 ui
=
xj xj xj
i
xj

Vf x j n j dS
Sint
(B.17)

uiuj
6. Turbulent shear term, (applying Theorem A, with ui u j )
xj

uiuj 1 uiuj 1
xj
=
xj

Vf uu n dS
Sint
i j j (B.18)

Collecting all double-averaged terms [Equations (B.9), (B.14), (B15), (B16), (B17),
and (B.18)] gives:

1 ui 1 ui u 1 u%i u% j
t
+
Vf u
Sint
i v j n j dS + u j
xj
+ i
Vf u n dS +
Sint
j j
xj

1 1 p 1

Vf u u n dS = g
Sint
i j j i
xi
+
V f pn dS
Sint
i

1 u 1 ui
+
xj
i
xj

Vf x
Sint
n j dS
j

1 uiu j 1

xj
+
Vf uu n dS
Sint
i j j (B.19)

Applying the chain rule to the first term in (B.19), and rearranging gives,

ui ui
+ uj
t xj

1 p 1 u 1 uiu j 1 u%i u% j
= gi + i
xi xj xj xj xj

1 1 ui
+
V f pn dS V x
Sint
i
f Sint
n j dS
j

PhD Thesis 224 APPENDIX B


L J Campbell Application of Integral Theorem A
1 ui 1 1 ui

Vf u v n
Sint
i j j dS
Vf u n dS + V u u n dS + V uu n dS
Sint
j j
f Sint
i j j
f Sint
i j j
t

(B.20)

If the roughness is rigid (i.e. frozen), no-slip, and solid (i.e. non-porous) then the
terms in the last line of (B.20) disappear. However, some, or all of these terms may
play a significant role in flows over mobile boundaries (for example, in sediment
transport cases). The two remaining surface integral terms (on line 3) represent form
and viscous drag, and only appear when averaging below roughness crests.

In the absence of the last line, Equation (B.20) can be verified as identical to
Equation (3.13) the version of the full double-averaged momentum conservation
equation given in Section 3.4.3. A slightly different version was given in Equation
(3.9), Section 3.4.2. However, Part 4 of Appendix B demonstrates how Equations
(3.9) and (3.13) are in fact equivalent.

PART 4. SUPPLEMENTARY WORKING TO DEMONSTRATE THE


EQUIVALENCE BETWEEN DIFFERENT VERSIONS OF
THE VOLUME-AVERAGED RANS EQUATION [i.e. those
given in Sections 3.4.2 and 3.4.3, Equations (3.9) and (3.13)],
INCLUDING FULL DEFINITON OF FORM AND VISCOUS
DRAG TERMS

Equation (3.9) stated:

ui ui 1 p 2 ui 1 uiu j
+ uj = gi +
t x j xi x j x j x j

1 u%i u% j 1 p% 2u%i
+ (3.9)
x j xi x j x j

whilst Equation (3.13) gave:

u i ui 1 p
+ u j = gi
t x j xi

1 uiu j 1 u%i u% j 1 u
+ i
x j x j x j x j

PhD Thesis 225 APPENDIX B


L J Campbell Application of Integral Theorem A
1 1 ui
+
V f pni dS V x n j dS (3.13)
Sint f Sint j

The discrepancy between (3.9) and (3.13) lies in the representation of drag
components following double averaging of the pressure and viscous terms in the
RANS equations (B.8). In order to demonstrate equivalence of terms involving
pressure and viscosity, it is necessary to show that:

1 p 1 p p%
xi

Vf pni dS =
xi
+
xi
(B.21)
Sint

(equating terms from (3.9) and (3.13) and multiplying by 1 fluid density)

Also that,

1 u 1 ui 2 ui 2u%i
i x n j dS = + (B.22)
x j x j Vf Sint j x j x j x j x j

Pressure Terms

Considering Equation (B.21) first, which contains terms arising from double
averaging the pressure gradient in the RANS equations ( p / xi ). This term can be
volume averaged in two ways, (1) by first substituting p = p + p% and then applying
integral Theorem A, or (2) by direct application of integral theorem A.

The first route gives,

p ( p + p% ) p p%
= = + (B.23)
xi xi xi xi

Then applying integral Theorem A [Equation (B.2)] to (B.23),

p 1 p 1 1 p% 1
xi
=
xi

Vf p ni dS +
xi

Vf pn
% i dS
Sint Sint

PhD Thesis 226 APPENDIX B


L J Campbell Application of Integral Theorem A
1 p p 1
=
xi

Vf ni dS
Vf pn
% i dS (B.24)
Sint Sint

(as obviously p = p , and p% =0)

And noting that if we put = 1 into integral Theorem A (B.2), we have (Gray, 1975;
Crapiste et al., 1986),

1 1
xi
=
Vf n dS
Sint
i (B.25)

Using (B.25) to substitute for part of the middle term of (B.24),

p 1 p p 1
xi
=
xi

xi

Vf pn
% i dS
Sint

Where, by the chain rule,

1 p p p
= + (B.26)
xi xi xi

Giving,

p p p p 1
xi
=
xi
+
xi

xi

Vf pn
% i dS
Sint

p 1
=
xi

Vf pn
% i dS (B.27)
Sint

It is interesting to note here as an aside, that following terms through from Equation
(B.24) to (B.27) shows that:

p p p% 1
xi
=
xi
, and
xi
=
Vf pn
% i dS (B.28)
Sint

PhD Thesis 227 APPENDIX B


L J Campbell Application of Integral Theorem A
which certainly helps to give a greater degree of physical meaning to the form drag
term appearing in Equation (3.9).

The second averaging route (direct application of Theorem A to the time-averaged


pressure gradient term) produces the result already given in Equation (B.16).

Hence, equating (B.16) and (B.27),

p 1 p 1 p 1
xi
=
xi

Vf pn dS
Sint
i =
xi

Vf pn
% dS
Sint
i (B.29)

and rearranging to have form drag on one side,

FORM DRAG

1 1 p p 1
Vf p n dS =
Sint
i
xi

xi
+
Vf p% n dS
Sint
i

p 1
=
xi
+
Vf p% n dS
Sint
i [by (B.26)]

p p%
= [by (B.28)] (B.30)
xi xi

Note that (B.30) gives the correct interpretation of form drag, and clarifies the
statement in Section 3.4.3 that p% / xi is only one component of form drag.

Finally, to demonstrate that Equation (B.21) is indeed true, substitute the expression
for form drag (B.30) into (B.21),

1 p p p% p p%
+ = + (B.31)
xi xi xi xi xi

and apply the chain rule to the first term, i.e. result (B.26),

PhD Thesis 228 APPENDIX B


L J Campbell Application of Integral Theorem A
p p% p p%
+ = + (B.32)
xi xi xi xi

Thus pressure terms are equal between Equations (3.9) and (3.13).

Viscosity Terms

Demonstration that Equation (B.22) holds for viscosity terms follows the same
method as for the pressure terms.

As above, route 1 for double-averaging the viscous shear term in the RANS
equations involves firstly substituting ui = ui + u%i for velocities,

ui ui 2u%i
= + (B.33)
x j x j x j x j x j x j

and then applying integral Theorem A once,

ui 1 ui 2u%i

xj

xj

Vf x j j

Sint
n dS +
x j x j
(B.34)

then once more,

1 ui 1

x j x j

Vf i j
u n dS
Sint

1 ui 2u%i

Vf x j n j dS + x j x j
Sint
(B.35)

rearranging, taking constant terms outside the surface integrals, and using relation
(B.25) to re-express both surface integral terms gives,

ui ui ui 2u%i
2

= ui +
x j x j x j x j x j x j x j x j x j x j

(B.36)

PhD Thesis 229 APPENDIX B


L J Campbell Application of Integral Theorem A
applying the chain rule to the second term on the right side yields,

ui ui 2 ui ui 2 2u%i
2

= +
x j x j x j x j x j x j x j x j x j x j
(B.37)

The second route for double-averaging the viscous shear term is by direct application
of integral Theorem A (B.2) (i.e. without first decomposing temporally-averaged
velocities). This produces the result already given in Equation (B.17),

ui 1 u 1 ui
=
xj xj xj
i
xj

Vf x
Sint
n j dS
j
[(B.17)]

applying Theorem A (B.2) again to the remaining average of a spatial derivative,

1 1 ui 1 1 ui

x j x j

Vf u n
Sint
i j dS
V f
x
Sint
n j dS
j

ui ui
2
1 1
=
xj xj x j x j

x j V f ui n j dS

Sint

1 ui

Vf x
Sint
n j dS
j
(B.38)

Hence, equating (B.37) and (B.38),

ui ui 2 ui ui 2 2u%i
2

= +
x j x j x j x j x j x j x j x j x j x j

ui
2
1 1 1 ui
=
x j x j

x j V f ui n j dS
Vf x j n j dS
Sint
Sint

(B.39)
and cancelling terms and rearranging gives the full expression for viscous drag as,

PhD Thesis 230 APPENDIX B


L J Campbell Application of Integral Theorem A
VISCOUS DRAG

1 ui 2u%i 1 1

Vf x j j
Sint
n dS =
x j x j
+
x j V f u ni j dS

Sint

ui 2 2 ui
(B.40)
x j x j x j x j

Which again helps explain the statement that ( 2u%i / x j x j ) is only one
component of viscous drag (made in Section 3.4.3). Note also that for rigid (i.e.
frozen), no-slip, and solid (i.e. non-porous) bed roughness then the surface
integral term on the right hand side disappears.

Finally, to show equivalence between the viscous terms in Equations (3.9) and (3.13)
(i.e. that (B.22) is true) if we briefly return to Equation (B.37),

ui ui 2 ui ui 2 2u%i
2

= +
x j x j x j x j x j x j x j x j x j x j

[(B.37)]

applying the chain rule to (B.37) to remove all products from derivative terms gives,

ui 2 2
= ui ui
x j x j x j x j x j x j

ui ui
+ 2 2
x j x j x j x j

2 ui 2u%i
+ +
x j x j x j x j

ui 2 ui 2u%i
= + (B.41)
xj xj x j x j x j x j

Therefore equating (B.17) to (B.41) proves that (B.22) is true and that the viscous
terms in Equations (3.9) and (3.13) amount to the same:

PhD Thesis 231 APPENDIX B


L J Campbell Application of Integral Theorem A
ui 1 u 1 ui
=
xj xj xj
i
xj

Vf x
Sint
n j dS
j

2 ui 2u%i
= + (B.42)
x j x j x j x j

Thus all pressure and viscosity terms in Equation (3.9) correspond exactly to those in
Equation (3.13), hence the two versions of the double-averaged momentum
equations are indeed fully equivalent.

PhD Thesis 232 APPENDIX B


L J Campbell Application of Integral Theorem A
APPENDIX C: SUPPLEMENTARY PLOTS FOR CHAPTER 5.0

C.1 Contours of Time-Averaged Streamwise Velocity (H=37, 85mm)


C.2 Double-Averaged Velocity Distributions in the Roughness Layer
C.3 Form-Induced Stress Distributions in the Roughness Layer
C.4 Instantaneous Streamlines for Skimming Flow, L=2

C.1 CONTOURS OF TIME-AVERAGED STEAMWISE VELOCITY

Flow depth, H=37mm


Velocities are
normalised by bulk
mean velocity (U =
Q/A)
Slope, depth,
roughness parameters
in experiment tag as
described in Section
4.6

Figure C.1: Contours of time-averaged streamwise velocity, H=37mm. Note: subplots are not
drawn to the same horizontal scale.

PhD Thesis 233 APPENDIX C


L J Campbell Supplementary Plots for Chapter 5.0
Flow depth, H=85mm
Velocities are
normalised by bulk
mean velocity (U =
Q/A)
Slope, depth, roughness
parameters in
experiment tag as
described in Section 4.6

Figure C.2: Contours of time-averaged streamwise velocity, H=85mm. Note: subplots are not
drawn to the same horizontal scale.

PhD Thesis 234 APPENDIX C


L J Campbell Supplementary Plots for Chapter 5.0
C.2 DOUBLE-AVERAGED VELOCITY DISTRIBUTION

Flow depth, H=37mm

Velocities are
normalised by bulk
mean velocity
(U = Q/A)

Slope, depth, roughness


parameters in
experiment tag as
described in Section 4.6

Figure C.3: The double-averaged velocity distribution in the roughness layer for H=37mm.
Velocities normalised with bulk mean velocity (U=Q/A).

PhD Thesis 235 APPENDIX C


L J Campbell Supplementary Plots for Chapter 5.0
Flow depth, H=50mm

Velocities are
normalised by bulk
mean velocity
(U = Q/A)

Slope, depth, roughness


parameters in
experiment tag as
described in Section 4.6

Figure C.4: The double-averaged velocity distribution in the roughness layer for H=50mm.
Velocities normalised with bulk mean velocity (U=Q/A).

PhD Thesis 236 APPENDIX C


L J Campbell Supplementary Plots for Chapter 5.0
Flow depth, H=85mm

Velocities are
normalised by bulk
mean velocity
(U = Q/A)

Slope, depth,
roughness parameters
in experiment tag as
described in Section
4.6

Figure C.5: The double-averaged velocity distribution in the roughness layer for H=85mm.
Velocities normalised with bulk mean velocity (U=Q/A).

PhD Thesis 237 APPENDIX C


L J Campbell Supplementary Plots for Chapter 5.0
C.3 FORM-INDUCED STRESS DISTRIBUTION

Flow depth,
H=37mm

Slope, depth,
roughness parameters
in experiment tag as
described in Section
4.6

Figure C.6: Form-induced stress (N/m2) distributions for H=37mm.

PhD Thesis 238 APPENDIX C


L J Campbell Supplementary Plots for Chapter 5.0
Flow depth,
H=50mm

Slope, depth,
roughness parameters
in experiment tag as
described in Section
4.6

Figure C.7: Form-induced stress (N/m2) distributions for H=50mm.

PhD Thesis 239 APPENDIX C


L J Campbell Supplementary Plots for Chapter 5.0
Flow depth,
H=85mm

Slope, depth,
roughness parameters
in experiment tag as
described in Section
4.6

Figure C.8: Form-induced stress (N/m2) distributions for H=85mm.

PhD Thesis 240 APPENDIX C


L J Campbell Supplementary Plots for Chapter 5.0
C.4 INSTANTANEOUS STREAMLINES FOR L=2

1 6

2 7

3 8

4 9

5 10

Figure C.9: Sequence of ten representative images of instantaneous streamlines from d-type
experiment S400_H50_L2. Shows degree of interaction between persistent counter-rotating cavity
vortex and the overlying flow. Flow is from left to right.

PhD Thesis 241 APPENDIX C


L J Campbell Supplementary Plots for Chapter 5.0
APPENDIX D: SUPPLEMENTARY PLOTS FOR CHAPTER 6.0
D.1 Streamlines of Time-Averaged Streamwise Flow
D.2 Simulated and Measured Form-Induced Velocity Components
D.3 Form-Induced Quadrant Analysis for Isolated Roughness Flows

D.1 CONTOURS OF TIME-AVERAGED STEAMWISE VELOCITY


Figure D.1: S400_H50_L2:
time-averaged streamlines
Figure D.2: S400_H50_L3:
time-averaged streamlines

PhD Thesis 242 APPENDIX D


L J Campbell Supplementary Plots for Chapter 6.0
Figure D.5: S400_H50_L6: Figure D.4: S400_H50_L5: Figure D.3: S400_H50_L4:
time-averaged streamlines time-averaged streamlines time-averaged streamlines

PhD Thesis
L J Campbell
243
Supplementary Plots for Chapter 6.0
APPENDIX D
Figure D.8: S400_H50_L10: Figure D.7: S400_H50_L8: Figure D.6: S400_H50_L7:
time-averaged streamlines time-averaged streamlines time-averaged streamlines

PhD Thesis
L J Campbell
244
Supplementary Plots for Chapter 6.0
APPENDIX D
Figure D.11: S400_H50_L20: Figure D.10: S400_H50_L15: Figure D.9: S400_H50_L12:
time-averaged streamlines time-averaged streamlines time-averaged streamlines

PhD Thesis
L J Campbell
245
Supplementary Plots for Chapter 6.0
APPENDIX D
D.2 SIMULATED AND MEASURED FORM-INDUCED VELOCITY
COMPONENTS (Simulation by Equation 6.1 versus measured data. Layout
is as for Figure 6.9).

SIMULATION [Eq.(6.1)]
a u cos( 2 x / p )
a w sin(2
2
x / p)
uw
%% / u
*
L=7

DATA [S400_H50_L7]

u% / u*
w% / u*2
uw
%% / u
*

Figure D.12: S400_H50_L7


SIMULATION [Eq.(6.1)]
L=8

DATA [S400_H50_L8]

Figure D.13: S400_H50_L8

PhD Thesis 246 APPENDIX D


L J Campbell Supplementary Plots for Chapter 6.0
SIMULATION [Eq.(6.1)]
L=10

DATA [S400_H50_L10]

Figure D.14: S400_H50_L10


SIMULATION [Eq.(6.1)]
L=12

DATA [S400_H50_L12]

Figure D.15: S400_H50_L12

PhD Thesis 247 APPENDIX D


L J Campbell Supplementary Plots for Chapter 6.0
SIMULATION [Eq.(6.1)]
L=15

DATA [S400_H50_L15]

Figure D.16: S400_H50_L15


SIMULATION [Eq.(6.1)]
L=20

DATA [S400_H50_L20]

Figure D.17: S400_H50_L20

PhD Thesis 248 APPENDIX D


L J Campbell Supplementary Plots for Chapter 6.0
D.3 FORM-INDUCED QUADRANT ANALYSIS: PLOTS FOR ALL
ISOLATED ROUGHNESS FLOWS (H=50mm)

In all figures, subplots show:

(a) Quadrant diagams (from the Zone of the Rings)


(b) Quadrant maps up to z=6k

All form-induced velocities are normalised with friction velocity, and all (a) subplots
are plotted using the same scale. Refer to Chapter 6, Section 6.3.2 for further details.

(a)

(b)

Figure D.18: S400_H50_L7

(a)

(b)

Figure D.19: S400_H50_L8

PhD Thesis 249 APPENDIX D


L J Campbell Supplementary Plots for Chapter 6.0
(a) (b)

Figure D.20: S400_H50_L10

(a) (b)

Figure D.21: S400_H50_L12

PhD Thesis 250 APPENDIX D


L J Campbell Supplementary Plots for Chapter 6.0
(a) (b)

Figure D.22: S400_H50_L15

(a) (b)

Figure D.23: S400_H50_L20

PhD Thesis 251 APPENDIX D


L J Campbell Supplementary Plots for Chapter 6.0
APPENDIX E: SUPPLEMENTARY PLOTS FOR CHAPTER 7.0

E.1 Sequences of Instantaneous Streamlines


E.2 Contour Maps of Slope Coefficient
E.3 Tables of Measured Channel Discharge

PhD Thesis 252 APPENDIX E


L J Campbell Supplementary Plots for Chapter 7.0
E.1 SEQUENCES OF INSTANTANEOUS STREAMLINES

1 6

2 7

3 8

4 9

5 10

Figure E.1: Ten representative time-steps showing instantaneous streamlines from flow over d-
type roughness spacing L=5. Note how the whole cavity is neatly filled by at least one large vortex
at all times. Subplot 2 shows the effect of an ejection from the upstream cavity, which can then be
charted as it passes through the measurement window in subplots 3 and 4. Flow is from left to
right.

PhD Thesis 253 APPENDIX E


L J Campbell Supplementary Plots for Chapter 7.0
1 6

2 7

3 8

4 9

5 10

Figure E.2: Ten representative time-steps showing instantaneous streamlines from flow over the
d-k-type roughness transition L=6. Note the relative instability (compared to L=5) of the cavity
flow structure at the upstream face of the bars. This is particularly evident in subplots 2 and 7, and
is deemed responsible for the transition phenomenon. Flow is from left to right.

PhD Thesis 254 APPENDIX E


L J Campbell Supplementary Plots for Chapter 7.0
E.2 CONTOUR MAPS OF SLOPE COEFFICIENT, a, FOR DEPTH
H=50mm

Figure E.3:
d-type
roughness
spacing,
L=5

Figures E.3-E.5:
Contour maps showing
the spatial variation of
slope coefficient, a, of
temporal quadrant
plots, i.e. w = au. The
large, clear yellow area
Figure E.4:
in Figure E.4 illustrates d-k-type
an extensive zone of roughness
transition,
reduced quadrant map L=6
slope at the d-k-type
transition. Flow depth,
H=50mm.

Figure E.5:
k-type
roughness
spacing,
L=7

PhD Thesis 255 APPENDIX E


L J Campbell Supplementary Plots for Chapter 7.0
E.3 TABLES OF MEASURED CHANNEL DISCHARGE (v-notch weir)

Experiment Q (l/s)
S400_H37_L2 3.2
S400_H37_L3 6.6
S400_H37_L4 2.4
S400_H37_L5 2.2
S400_H37_L6 2.6
S400_H37_L7 6.1
S400_H37_L8 2.5
S400_H37_L10 2.1
S400_H37_L12 2.6
S400_H37_L15 1.6
S400_H37_L20 7.1

Table E.1: Weir-measured channel discharge (Q) for flow depth H = 37 mm

Experiment Q (l/s)
S400_H50_L2 6.0
S400_H50_L3 8.7
S400_H50_L4 4.5
S400_H50_L5 3.4
S400_H50_L6 5.8
S400_H50_L7 7.2
S400_H50_L8 3.8
S400_H50_L10 3.3
S400_H50_L12 4.4
S400_H50_L15 2.7
S400_H50_L20 9.2
Table E.2: Weir-measured channel discharge (Q) for flow depth H = 50 mm

Experiment Q (l/s)
S400_H85_L2 12.1
S400_H85_L3 11.3
S400_H85_L4 12.0
S400_H85_L5 6.7
S400_H85_L6 13.8
S400_H85_L7 10.0
S400_H85_L8 7.2
S400_H85_L10 6.6
S400_H85_L12 10.5
S400_H85_L15 6.2
S400_H85_L20 13.8

Table E.3: Weir-measured channel discharge (Q) for flow depth H = 85 mm

PhD Thesis 256 APPENDIX E


L J Campbell Supplementary Plots for Chapter 7.0

You might also like