Vous êtes sur la page 1sur 382

C S James

Hydraulic Structures
Hydraulic Structures
C S James

Hydraulic Structures

123
C S James
School of Civil and Environmental
Engineering
University of the Witwatersrand
Johannesburg, South Africa

ISBN 978-3-030-34085-8 ISBN 978-3-030-34086-5 (eBook)


https://doi.org/10.1007/978-3-030-34086-5
© Springer Nature Switzerland AG 2020
This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt from
the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, expressed or implied, with respect to the material contained
herein or for any errors or omissions that may have been made. The publisher remains neutral with regard
to jurisdictional claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface

This textbook is intended to provide for the needs of senior undergraduate and
postgraduate students in civil engineering programmes and graduate engineers in
civil engineering practice. It aims to provide a solid grounding in the theory
underlying the design and analysis of hydraulic structures, including underflow
gates, transitions, spillways, culverts, energy dissipators, flow measuring structures
and river intakes, as well as the prediction and prevention of scour.
Well-established theory and procedures are presented, as well as recent develop-
ments gleaned from the research literature with a design-oriented perspective. As a
learning resource, the book is not intended to provide complete design details, but
rather to develop understanding and competence in applying basic theoretical
concepts; the reader is directed to many freely available design guides for more
detailed treatment. There is, however, sufficient detail for preliminary design of
many structures and complete design of some standard structures such as measuring
weirs and flumes and stilling basins. Worked examples are presented in each
chapter and exercise problems are provided.
The content of the book is based on a postgraduate course in hydraulic structures
presented at the University of the Witwatersrand, Johannesburg over many years. At
this level, mastery of basic hydraulic theory is assumed. For revision, the intro-
ductory chapter provides an overview of the basic concepts relevant to the subse-
quent content, particularly covering steady rapidly and gradually varied flow in open
channels, and flow resistance. At postgraduate level, students and practitioners
should be aware of the incompleteness of current knowledge, and be adept at
keeping abreast of new developments. Knowledge of some topics presented in the
book is clearly not yet definitive, but their inclusion exemplifies the evolving state
of the art and the references cited provide first recourse for pursuing advancements.
I am indebted to the authors of the many books and articles from which I have
gained the knowledge that is presented in this work. I have acknowledged all direct
sources as far as possible, but there is much accumulated influence that is implicit

v
vi Preface

and difficult to source and attribute. I am especially indebted to Dr. Heinz Weiss for
inspiring my interest in open channel hydraulics as an undergraduate student so
many years ago, and to my many students since who have challenged my under-
standing and explanations.

Johannesburg, South Africa C S James


Contents

1 Basic Hydraulic Concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1


1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Flow Classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 The Conservation Laws in Hydraulics . . . . . . . . . . . . . . . . . . . . . 7
1.3.1 Conservation of Mass—The Continuity Equation . . . . . . . 7
1.3.2 Conservation of Energy . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.3.3 Conservation of Momentum . . . . . . . . . . . . . . . . . . . . . . . 12
1.4 Steady Uniform Flow and Flow Resistance . . . . . . . . . . . . . . . . . 14
1.4.1 The General Resistance Equation . . . . . . . . . . . . . . . . . . . 15
1.4.2 The Chézy Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.4.3 The Darcy–Weisbach Equation . . . . . . . . . . . . . . . . . . . . . 18
1.4.4 The Manning Equation . . . . . . . . . . . . . . . . . . . . . . . . . . 25
1.5 Steady Rapidly Varied Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
1.5.1 Application of Energy and Momentum Conservation . . . . . 28
1.5.2 The Control Concept . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
1.6 Steady Gradually Varied Flow . . . . . . . . . . . . . . . . . . . . . . . . . . 49
1.6.1 The Gradually Varied Flow Equation . . . . . . . . . . . . . . . . 50
1.6.2 Classification of Gradually Varied Profiles . . . . . . . . . . . . 51
1.6.3 Gradually Varied Flow Computation . . . . . . . . . . . . . . . . . 52
Reference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
2 Underflow Gates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
2.2 Unsubmerged Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
2.3 Submerged Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
2.4 Hysteretic Behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

vii
viii Contents

3 Open Channel Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
3.2 Subcritical Flow Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
3.3 Supercritical Flow Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
3.3.1 Straight Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
3.3.2 Curvilinear Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
3.3.3 Suppression of Standing Wave Propagation . . . . . . . . . . . . 94
3.4 Dual Stable States and Hysteresis . . . . . . . . . . . . . . . . . . . . . . . . 97
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
4 Spillways . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
4.1 Introduction to Conveyance Structures . . . . . . . . . . . . . . . . . . . . . 105
4.2 Spillway Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
4.2.1 The Overflow Spillway . . . . . . . . . . . . . . . . . . . . . . . . . . 108
4.2.2 Labyrinth and Piano Key Weirs . . . . . . . . . . . . . . . . . . . . 117
4.2.3 The Side-Channel Spillway . . . . . . . . . . . . . . . . . . . . . . . 119
4.2.4 The Side Weir . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
4.2.5 Shaft (Morning Glory) Spillways . . . . . . . . . . . . . . . . . . . 130
4.2.6 Siphon Spillways . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
4.2.7 Chutes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
4.2.8 Stepped Chutes and Spillways . . . . . . . . . . . . . . . . . . . . . 146
4.3 Cavitation and Aeration on Spillways and Chutes . . . . . . . . . . . . . 150
4.3.1 Cavitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
4.3.2 Aeration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
5 Culverts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
5.2 Inlet Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
5.3 Outlet Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
6 Energy Dissipation Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
6.2 The Hydraulic Jump . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
6.2.1 Hydraulic Jump Characteristics . . . . . . . . . . . . . . . . . . . . . 184
6.2.2 Controlled Hydraulic Jumps . . . . . . . . . . . . . . . . . . . . . . . 195
6.3 Standard Stilling Basins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
6.4 Other Energy Dissipators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
6.4.1 Bucket-Type Dissipators . . . . . . . . . . . . . . . . . . . . . . . . . 225
6.4.2 Impact-Type Dissipators . . . . . . . . . . . . . . . . . . . . . . . . . . 228
6.4.3 Baffled Spillways . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
Contents ix

6.4.4 Stepped Chutes and Spillways . . . . . . . . . . . . . . . . . . . . . 231


6.4.5 Spillway Splitters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
7 Flow-Measuring Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
7.2 Weirs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
7.2.1 Sharp-Crested Weirs . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
7.2.2 Broad-Crested Weirs . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
7.2.3 Advantages and Disadvantages of Weirs for Flow
Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
7.3 Flumes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
7.3.1 Throated (Venturi) Flume . . . . . . . . . . . . . . . . . . . . . . . . . 257
7.3.2 The Parshall Flume . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
7.3.3 The Cutthroat Flume . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
7.4 Long-Throated Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277
7.5 Errors and Measuring Ranges . . . . . . . . . . . . . . . . . . . . . . . . . . . 277
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
8 Intake Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
8.1.1 Reservoir Intakes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
8.1.2 River Intakes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
8.2 River Intake Design for Sediment Control . . . . . . . . . . . . . . . . . . 285
8.2.1 Vertical Sediment Distribution . . . . . . . . . . . . . . . . . . . . . 286
8.2.2 Bed Load Movement Around Bends . . . . . . . . . . . . . . . . . 290
8.2.3 Sediment Exclusion Structures . . . . . . . . . . . . . . . . . . . . . 292
8.3 Pump Sumps and Intakes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
8.3.1 Desirable Flow Conditions . . . . . . . . . . . . . . . . . . . . . . . . 307
8.3.2 Intake and Sump Design . . . . . . . . . . . . . . . . . . . . . . . . . 309
8.3.3 Model Testing for Intakes . . . . . . . . . . . . . . . . . . . . . . . . 315
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
9 Scour and Scour Protection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
9.2 Theoretical Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 320
9.3 Empirical Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
9.4 Design Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 327
9.4.1 Critical Shear Stress Design . . . . . . . . . . . . . . . . . . . . . . . 327
9.4.2 Permissible Velocity Design . . . . . . . . . . . . . . . . . . . . . . . 336
9.4.3 Protection of Underlying Material . . . . . . . . . . . . . . . . . . . 339
x Contents

9.5 Scour Around Bridge Piers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341


9.5.1 Scour Depth Estimation . . . . . . . . . . . . . . . . . . . . . . . . . . 342
9.5.2 Bridge Scour Countermeasures . . . . . . . . . . . . . . . . . . . . . 351
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361

Postscript . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367
Symbols

a Acceleration (m/s2)
a Bridge pier width (m)
a Empirical coefficient
a Reference height for sediment concentration profile (m)
a Sediment particle weight lever arm (m)
a Sluice gate opening (m)
ap Projected width of bridge pier (m)
a* Effective bridge pier diameter (m)
A Cross-sectional area (m2)
A Half horizontal ellipse axis for spillway profile (m)
A Plan area of settling basin (m2)
Ar Channel expansion area ratio
b Bridge abutment width (m)
b Channel width (m)
b Sediment particle drag force lever arm (m)
b Siphon barrel width (m)
b Weir crest width (m)
B Average width of trapezoidal channel (m)
B Box culvert/cutthroat flume width (m)
B Channel width (m)
B Flow surface width (m)
B Half vertical ellipse axis for spillway profile (m)
c Centrifugal pressure adjustment (m)
c Empirical aeration-bulked flow coefficient
c Sediment particle lift force lever arm (m)
c Wave celerity (m/s)
ci Contraction loss coefficient
co Expansion loss coefficient
C Air concentration
C Chézy resistance coefficient (m1/2/s)

xi
xii Symbols

C Empirical spillway discharge coefficient (m1/2/s)


C Sediment concentration
Ca Reference sediment concentration
Cc Contraction coefficient
Cd Weir/spillway/culvert/flume discharge coefficient
Ce Concentration of entrained air
Ce∞ Equilibrium concentration of entrained air
Cf Energy loss coefficient
Cf Stability factor adjustment for riprap stability
Ch Contraction coefficient
CL Channel expansion loss coefficient
Co Shaft spillway discharge coefficient (m1/2/s)
Co Air concentration at chute surface
Cp Pressure coefficient
Csg Density correction for riprap stability
C1 Coefficient for submerged cutthroat flume equation
d Outlet flow depth (m)
d Siphon barrel depth (m)
D Circular bridge pier diameter (m)
D Culvert barrel height (m)
D Flow depth (m)
D Inlet bellmouth diameter (m)
D Offset height for cavitation (mm)
D Pipe diameter (m)
D Sediment particle size (m)
Dr Riprap stone size (m)
Ds Grain size of material underlying filter layer (mm)
Dv Depth of Crump weir vee (m)
D15 upper 15% grain size of upper filter layer (mm)
D15 lower 15% grain size of lower filter layer (mm)
D16 16% sediment grain size (mm)
D50 Median grain size (mm)
D84 84% sediment grain size (mm)
D85lower 85% grain size of lower filter layer
E Sediment settling efficiency
E Specific energy (m)
Eo Limiting sediment settling efficiency
f Darcy–Weisbach friction factor
f Submergence correction for Crump weir
fa Aerated flow friction factor
fe Effective friction factor for stepped chutes
fs Riprap factor of safety
f 1, f 2, f 3 Factors in bridge scour equation
F Force (N)
F Force on bank particle (N)
Symbols xiii

FB Force on baffle blocks per unit width (N/m)


Fd Densimetric particle Froude number
FD Drag force (N)
FD Pump inlet Froude number
Ff Friction force
FH Hydrostatic pressure force (N)
FL Lift force (N)
FR Shear resistance force (N)
Fr Froude number
Fr0 Approach flow Froude number
FW Weight component (N)
F1, F2 Variables in submergence correction for Crump weir
F2 Downstream hydrostatic force for hydraulic jump
F*2 Downstream hydrostatic force for simple hydraulic jump
g Gravitational acceleration (m/s2)
G Modified Froude number in equation for hydraulic jump on slope
Go Radial gate opening (m)
GV Gradually varied flow
h Baffle block height (m)
h Height of energy level above weir crest (m)
h Height of stepped spillway step (m)
h Height of water level above base of spillway (m)
h Height of water surface above shaft spillway throat (m)
h Pressure head (m)
hb Bend head loss (m)
hd Height of downstream water level above weir crest (m)
hdam Dam height (m)
he Entrance head loss (m)
hexit Exit head loss (m)
hf Friction loss (m)
hi USBR stilling basin feature heights (m)
hL Local head loss (m)
ho Outlet head loss (m)
hs Hydrostatic pressure head (m)
hv Velocity head (m)
H Reservoir water level (m)
H Total energy (m)
H/ Energy level above weir crest (m)
Ha Head on shaft spillway throat (m)
Ha Minimum head for spillway splitter operation (m)
Hc Energy at siphon crest (m)
Hc Maximum head for spillway splitter operation (m)
Hd Design head on spillway (m)
Hdam Height of dam structure (m)
Hmax Total energy upstream of dam structure
xiv Symbols

Ho Design head for shaft spillway (m)


Hs Potential flow velocity head over spillway (m)
Ht Total head across shaft spillway (m)
H1, H2 Horizontal bed gradually varied profiles
HGL Hydraulic grade line
HW Headwater, reservoir water level (m)
I Intensity of sediment motion (t−1)
k Exponent in settling efficiency equation
k Roughness height (m)
ke Entrance loss coefficient
kn Proportionality constant for Manning n related to sediment size
ks Nikuradse roughness (m)
K Coefficient in sediment incipient motion equation
K Empirical spillway profile coefficient
K, K1 Discharge coefficients for Parshall/cutthroat flume
K Laminar flow friction coefficient
Ka Spillway abutment contraction coefficient
Kp Spillway pier contraction coefficient
K1 Transverse slope factor for riprap stability
l Distance of hydraulic jump from slope junction (m)
l Length of stepped spillway step (m)
L Aeration inception distance (m)
L Cutthroat flume/spillway/weir crest length (m)
L Length (m)
L Length of settling basin (m)
L Length of spillway splitter (m)
L Wavelength (m)
L/ Effective spillway crest length (m)
Lb Distance from toe of hydraulic jump to baffle block (m)
LB Stilling basin length (m)
Leff Effective weir crest length (m)
Lj Hydraulic jump length (m)
Li Distance from chute blocks to sill in stilling basin (m)
LII Length of USBR Basin II
Lr Hydraulic jump roller length (m)
L*r Simple hydraulic jump roller length (m)
Ls Distance from hydraulic jump toe to sill (m)
Ls Width of spillway splitter shelf (m)
m Mass (kg)
m Number of sediment particles displaced in time interval
m Shape factor for throated flume
M Momentum function (m3 or m2 for unit width)
M 1, M 2, M 3 Mild slope gradually varied profiles
n Crump weir side slope (1:n)
n Empirical spillway profile exponent
Symbols xv

n Manning resistance coefficient (s/m1/3)


n Number of concentric rows of riprap stones
n1, n 2 Cutthroat flume equation exponents
N Length factor in hydraulic jump equation
N Number of piers on spillway
N Number of sediment particles in observed area
N Particle weight component normal to bank (N)
N Shape factor for bridge pier
p Pressure (Pa)
pa Atmospheric pressure (Pa)
pc Pressure at siphon crest (Pa)
po Reference pressure (Pa)
pv Vapour pressure (Pa)
P Location of spillway splitters below crest (m)
P Net force (N)
P Wetted perimeter (m)
P Spillway/weir approach depth (m)
q Unit width discharge (m3/s/m)
Q Discharge (m3/s)
Qa Air discharge (m3/s)
Qf Free flow discharge for gated spillway (m3/s)
Qi Pump sump inflow discharge (m3/s)
Qp Sump pumping rate (m3/s)
Q* Weir discharge corrected for submergence (m3/s)
r Channel expansion width ratio
r Radius of curvature (m)
ri Inner radius of siphon spillway barrel (m)
ro Outer radius of siphon spillway barrel (m)
R Force resisting motion of bank particle (N)
R Hydraulic radius (m)
R Reaction force on sediment particle (N)
R0 Approach flow hydraulic radius (m)
Re Reynolds number
Re* Shear Reynolds number
Rmax Distance of outer spray surface from spillway with splitters (m)
Rmin Distance of inner spray surface from spillway with splitters (m)
Rs Shaft spillway crest radius (m)
RV Rapidly varied flow
s Expansion flow depth ratio
s Sill or step height (m)
si USBR stilling basin feature spacings (m)
S Pump intake submergence (m)
S Slope
S Spacing between spillway splitters (m)
S Submergence ratio
xvi Symbols

S s/y1
Sf Energy gradient
So Bed slope
Ss Channel side slope
Ss Sediment specific gravity
St Transition submergence ratio
Sw Water surface slope
SF Stability factor for riprap
S1, S2, S3 Steep slope gradually varied profiles
t Riprap layer thickness (m)
t Time or time interval (s)
te Time to equilibrium bridge pier scour (days)
tr Normalizing reference time for scour evolution (s)
t90 Time of scour to 90% of equilibrium depth (days)
T Dimensionless time variable
T Distance between top of spillway splitters and shelf (m)
T Particle weight component down bank slope (N)
T Sump pump cycle time (s)
TW Tailwater height (m)
u Discharge exponent for Parshall flume
u* Shear velocity (m/s)
u*c Critical entrainment shear velocity (m/s)
v Local velocity (m/s)
vc Velocity at siphon crest (m/s)
V Average velocity (m/s)
V Pump sump volume (m3)
Va Actual velocity at foot of spillway (m/s)
Vc Critical average velocity for sediment incipient motion (m/s)
Vo Approach velocity to weir (m/s)
Vo Reference velocity (m/s)
Vth Theoretical velocity at foot of spillway (m/s)
V0 Average approach velocity (m/s)
V0p Approach velocity for maximum live-bed scour (m/s)
V0p1 Approach velocity for maximum live-bed scour (m/s)
V0p2 Approach velocity for maximum live-bed scour (m/s)
V Average velocity over transverse distance (m/s)
w Baffle block width (m)
w Sediment settling velocity (m/s)
wi USBR stilling basin feature widths (m)
W Cutthroat flume throat width (m)
W Weight or submerged weight of sediment particle (N)
W Weight (N)
W Weight of sediment leaving settling basin (N)
W Weir height (m)
Symbols xvii

W Width (m)
W Width of spillway splitter (m)
Wb Radial gate width (m)
Wo Weight of sediment entering settling basin (N)
x Coordinate direction
x Depth ratio for throated flume
x Distance (m)
x Distance to critical section in side channel (m)
x Location of spillway splitters below reservoir level (m)
x* Distance from aeration inception point (m)
X Distance between chute blocks and baffle piers in stilling basin (m)
y Coordinate direction
y Flow depth (m)
y Height of spillway splitters above tailwater (m)
yb Aeration-bulked flow depth (m)
yc Critical flow depth (m)
yo Uniform flow depth (m)
yR Normalizing reference length for scour depth (m)
ys Scour depth (m)
yse Equilibrium scour depth (m)
yt Tailwater depth (m)
y0 Approach flow depth (m)
y1 Flow depth upstream of hydraulic jump
y2 Flow depth downstream of hydraulic jump
y*2 Downstream flow depth for simple hydraulic jump (m)
Y Sequent depth ratio for hydraulic jump
Y* Sequent depth ratio for simple hydraulic jump
z Baffle block spacing (m)
z Coordinate direction
z Depth to flow section centroid (m)
z Elevation (m)
z/ Height of bed above datum (m)
a Bed or bank slope (degrees)
a Coefficient in aeration inception equation
a Coriolis coefficient; kinetic energy correction factor
a Flow approach angle for bridge pier
a Labyrinth weir side wall angle (degrees)
a Triangular weir angle (degrees)
a Water surface slope in USBR Basin II
a/ Pressure coefficient for curvilinear flow
b Bed shear correction factor in hydraulic jump equation
b Exponent in aeration inception equation
b Momentum correction factor
b Radial gate angle (degrees)
b Ratio of pier diameter to channel width
xviii Symbols

b Wave disturbance angle (degrees)


b/ Pressure force correction for curvilinear flow
b1 Angle component in curved transition solution (degrees)
c Fluid specific weight (N/m3)
cs Sediment specific weight (N/m3)
C1 Variable in equation for hydraulic jump on slope
d Boundary layer thickness (m)
d Ratio of riprap size to underlying material size
DYf Hydraulic jump sequent depth correction to account for wall friction
DYs Hydraulic jump sequent depth correction to account for sill
Dz Change in bed level (m)
Dh Side wall deflection angle (degrees)
e Multiple of riprap stone size for distance specification
eh Indication error
eQ Discharge error
es Sediment diffusivity (m2/s)
η =w/(w + z)
η0 Riprap stone stability number
h Angle of underflow gate lip (multiples of 90o)
h Bed slope angle (degrees)
h Varying side wall angle (degrees)
h1 Integration constant for curved transition (degrees)
j Von Karman constant
K =Ls/L*r
l Absolute fluid viscosity (kg/m/s)
m Kinematic fluid viscosity (m2/s)
q Fluid density (kg/m3)
q1 Ratio of water to water + air discharge
r Cavitation index
r Sediment geometric standard deviation
so Boundary shear stress (N/m2)
sc Critical boundary shear stress at incipient sediment motion (N/m2)
scb Critical shear stress on bank particle (N/m2)
s* Dimensionless shear stress (Shields parameter)
s*c Critical dimensionless shear stress (critical Shields parameter)
u Angle of repose of sediment material (degrees)
u Sediment particle pivot angle (degrees)
w1 Size factor in hydraulic jump calculation with baffles
w2 Blockage factor in hydraulic jump calculation with baffles
Chapter 1
Basic Hydraulic Concepts

1.1 Introduction

Hydraulic structures can be broadly classified into four groups: conveyance struc-
tures are designed to pass a specified discharge safely from one location to another;
regulation structures control flow by inducing required water levels or discharges;
measuring structures enable discharges to be determined from measured water levels
and energy dissipation structures cause excess flow energy to be expended safely.
The design and analysis of such structures require establishing relationships between
flow characteristics such as discharge, pressure, flow depth and boundary shear
stress, and the bounding geometry of the structure. Most commonly, it is required to
predict the flow depth (y) around and through a structure for a particular design
discharge (Q). (For measuring structures, it is required to predict the discharge
corresponding to a measured flow depth or depths.)
As an example, a simple structure installed in a channel in the form of a smooth
rise and fall (Fig. 1.1) could affect the flow depths over and around it in a number of
ways; possible water surface profiles are shown in Fig. 1.1. There may be only a local
effect, with the water surface rising or dipping over the structure (Figs. 1.1a, c), or the
effect could extend considerable distances upstream and downstream, involving a
hydraulic jump (Figs. 1.1b, d).
Only one of these profiles can actually occur for a particular discharge and
structure size, and which it is depends on some characteristics of the flow. Before
developing the relationships required for describing flow responses to boundary
geometry, it is therefore necessary to characterize the different flow ‘types’ that
influence its behaviour.

© Springer Nature Switzerland AG 2020 1


C S James, Hydraulic Structures,
https://doi.org/10.1007/978-3-030-34086-5_1
2 1 Basic Hydraulic Concepts

(a) (b)
Q Q

(c) (d)
Q Q

Fig. 1.1 Possible water surface responses to a simple structure

1.2 Flow Classification

Both free surface and closed conduit flow are encountered in hydraulic structures.
The following classification is presented for the more general case of free surface
flow (Fig. 1.2). Some characteristics and definitions are the same for closed conduit
flow, while some are unique to free surface flows.
The first distinction between flow types is their classification as steady or
unsteady according to the variation with time. If flow conditions (such as velocity
and depth) do not change with time, then the flow is steady. If they do change with

Free Surface Flow

Steady Unsteady

Uniform Nonuniform (Uniform)

Gradually Rapidly
Varied Varied

Fig. 1.2 Basic types of free surface flow


1.2 Flow Classification 3

time, then the flow is unsteady. The small-scale variations of flow associated with
turbulent fluctuations and slow variations over long periods are not here considered
to be violations of steadiness.
Flow can also be classified as uniform or nonuniform according to its variation
with position. If the flow conditions are the same everywhere in the flow, then the
flow is uniform. If the flow conditions are different at different locations, then the
flow is nonuniform. In most practical applications, flow is assumed to be uniform if
average flow conditions do not change along a conduit, so that changes of direction
through bends and the variation of velocity through the boundary layer are not
considered as violations of uniformity.
Unsteady nonuniform flow is the general type, and commonly needs to be
described fully, particularly in natural situations. For many problems, the flow can
be considered to be steady and sometimes also uniform without introducing sig-
nificant errors in practical solutions.
Steady uniform flow is the simplest flow type, and its analysis provides results
which have useful application to the more general types as well. Unsteady uniform
flow is practically impossible; it would require a water surface to change with time
but remain parallel to its original form, which is impossible for an incompressible
fluid like water.
Nonuniform flow may be steady or unsteady. In either case, it can be further
classified as gradually varied (GV) or rapidly varied (RV). In gradually varied flow,
the changes in flow condition take place over long distances, such as the backing up
which occurs upstream of a dam or weir. In rapidly varied flow, the changes take
place over comparatively short distances, such as those over a structure, through a
constriction, or in a hydraulic jump. Rapidly and gradually varied changes are
fundamentally different in nature, the former being associated primarily with local
changes in boundary geometry with no significant influence of surface resistance,
and the latter being controlled primarily by flow resistance. The profile shown in
Fig. 1.1b can be classified for steady flow in these terms, as shown in Fig. 1.3.
Unsteady gradually varied flow is the most general flow type in nature, and
needs to be analysed for describing the passage of flood waves along rivers.
Unsteady rapidly varied flow also occurs as surges in rivers (as, for example, the

nonuniform

uniform
GV RV GV RV
Q

uniform

Fig. 1.3 Steady water surface profiles induced by a simple structure in a channel
4 1 Basic Hydraulic Concepts

famous tidal bore in the River Severn) and canals (for example, as resulting from
rapid closure of gates, particularly in hydroelectric power installations). Steady
rapidly varied flow is the typical situation for hydraulic structure analysis and
design.
Another flow classification which applies to open channel flow as well as to
closed conduit flow distinguishes between laminar and turbulent flow. The dis-
tinction can be defined in terms of the Reynolds number (Re), which expresses the
relative significance of inertial and viscous forces in determining the flow beha-
viour. Laminar flow is not common in practical applications of open channel flow
and will not be given much emphasis. It could occur in physical models or in sheet
flow such as drainage of paved surfaces, however, and this possibility should be
recognized and its occurrence established when it occurs. This can be done by
checking the value of the Reynolds number against established values defining the
ranges of each flow type. The Reynolds number is defined for pipes as

VD
Re ¼ ð1:1Þ
m

in which V is the average flow velocity, D is the pipe diameter and m is the
kinematic viscosity of the fluid. The threshold values for laminar and turbulent flow
for pipes are

laminar flow Re \ 2000


turbulent flow Re [ 4000

Between these values, there is an uncertain transition range.


Obviously, D is not defined for open channels and Re is usually defined in terms
of a parameter known as the hydraulic radius, R. This is defined as

A
R ¼ ð1:2Þ
P

in which A is the cross-sectional area and P is the length of wetted perimeter.


If D is replaced by R in the definition of Re, the threshold values for laminar and
turbulent flows will be different because, as can easily be shown, R = D/4. The
threshold values, and other important relationships, can also be applied unaltered to
open channels if the Reynolds number is defined as

VD 4VR
Re ¼ ¼ ð1:3Þ
m m

The inclusion of the factor 4 in the definition of the Reynolds number simplifies
the use of results obtained from pipe data considerably, and is widely used.
There are different types of turbulent flow, depending on the structure of the
boundary layer. Close to the solid boundary the flow is very slow and a viscous
sub-boundary layer can develop with a thickness (d) determined by the magnitude
1.2 Flow Classification 5

Fig. 1.4 Types of turbulent


flow past a solid boundary k δ

(a) Hydraulically smooth

k δ

(b) Transitional

k
δ
(c) Hydraulically rough

of the boundary shear stress, the fluid viscosity and the size of the roughness
elements on the boundary, k (Fig. 1.4).
Hydraulically smooth flow occurs if the roughness elements are well submerged
by the viscous flow layer, i.e. k < d. Transitional flow occurs when the roughness
elements are about the same size as the viscous layer thickness, and starts
influencing the flow further away, i.e. k  d. Hydraulically rough flow occurs when
the roughness elements are more than about five times the viscous layer thickness
appropriate to the prevailing flow conditions, and break it up completely.
The transition between laminar and turbulent flow can be defined by the
Reynolds number but the transition between the different types of turbulent flow
takes place at different values of Re, depending on the surface roughness. The
Reynolds number is therefore not sufficient to characterize the different types of
turbulent flow. An appropriate parameter can be determined for doing this by
considering quantitatively the physical reason for the differences, i.e. the relation-
ship between the thickness of the viscous sublayer and the size of roughness
elements, k/d.
The thickness of the viscous sublayer can be calculated as

11:6 m
d¼ ð1:4Þ
u

in which u* is the shear velocity (or friction velocity) defined by


rffiffiffiffiffi sffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffi
so cRS
u ¼ ¼ ¼ gRS ð1:5Þ
q q
6 1 Basic Hydraulic Concepts

in which so is the boundary shear stress, c (=qg) is the fluid specific weight, q is the
fluid density and g is gravitational acceleration. The ratio of the roughness element
size to the viscous sublayer thickness is therefore

k k u
¼
d 11:6 m

Removing the constant 11.6, and replacing k by the Nikuradse roughness, ks,
(which is related to, but not necessarily equal to the actual roughness size), means
that

k k s u
a ¼ Re ð1:6Þ
d m

in which Re* is a dimensionless number, very similar in form to the Reynolds


number, known as the shear Reynolds number, Re*. The shear Reynolds number
can be used to define the regimes of turbulent flow as follows:

Re \ 5 hydraulically smooth flow


5\Re \ 70 transitional flow
Re [ 70 hydraulically rough flow

Note that flow in a particular channel may be laminar or may be turbulent in any
of these regimes, depending on the prevailing discharge. These are hydraulic
concepts, and the flow type cannot be classified according to the channel charac-
teristics only—the discharge must be known as well.
A further very important flow classification for open channel flow reflects the
relative importance of gravitational and inertial forces on the flow, which is sig-
nificant because of the presence of the free surface and can be characterized by the
Froude number. This is defined as

Q
Fr ¼ qffiffiffiffiffiffiffi ð1:7Þ
g AB

in which g is gravitational acceleration and B is the flow width at the surface. For
rectangular channels, the surface width is equal to the channel width, so

V
Fr ¼ pffiffiffiffiffiffi ð1:8Þ
gy

where y is the flow depth.


If the value of Fr is <1, then the flow is called subcritical, if Fr > 1 it is
supercritical, and if Fr = 1, it is critical. This is a very important classification for
open channel flow because the flow behaviour of subcritical and supercritical flow
is quite different. The flow in Fig. 1.1a is wholly subcritical, in Fig. 1.1c it is
1.2 Flow Classification 7

subcritical critical
Q

supercritical subcritical

Fig. 1.5 Subcritical, critical and supercritical flow in a composite profile

wholly supercritical and in Figs. 1.1b, d all types occur, as shown in Fig. 1.5 for the
profile of Fig. 1.1b.
The following sections present the essential results for analysing steady uniform
and nonuniform flow conditions.

1.3 The Conservation Laws in Hydraulics

Most practical problems in hydraulics are solved by application of the three basic
conservation laws, viz., conservation of mass, energy and momentum. These are
expressed in equations with various forms. The most complete forms are obviously
the most general, but they are also the most complicated. By making certain
assumptions, particularly of steady and incompressible flow, simpler forms can be
derived which are adequate for many practical applications. Only the simple forms
necessary for the applications considered in this text are presented here; more
complete forms and their derivations are presented in standard introductory
hydraulics texts. These simple forms do have limitations resulting from their
underlying assumptions, which should be noted.

1.3.1 Conservation of Mass—The Continuity Equation

Under the assumptions of steady, incompressible flow, the form of the mass con-
servation equation is particularly simple. The incompressibility assumption enables
variations of density to be ignored and the law considered in terms of volume. The
simplest form of the continuity equation applies where the discharge does not vary
in the flow direction, and concern is with flow in one direction only. Then

Q ¼ V A ¼ constant ð1:9Þ

This form will be sufficient for most applications in analysing hydraulic struc-
tures. In cases where the discharge varies spatially or velocity components in more
8 1 Basic Hydraulic Concepts

than one direction need to be considered, more complete forms are required. These
will be introduced where necessary.

1.3.2 Conservation of Energy

The law of energy conservation as applied in hydraulics states simply that the
energy of the flow at a particular location, or section, is equal to the energy at a
specified section upstream minus the losses between the two sections, i.e.

Hdownstream ¼ Hupstream  losses ð1:10Þ

where H represents the total energy of the flow and the subscripts represent the
spatial order of the sections. The total hydraulic energy includes both potential and
kinetic components. It is transformation between these components that determines
the variation of flow conditions induced by hydraulic structures, and these can be
described using the Bernoulli equation. This equation can be derived by integrating
Newton’s second law with respect to distance, subject to the assumptions of steady,
incompressible, irrotational flow, resulting in the form

p v2
þ zþ ¼ constant ¼ H ð1:11Þ
c 2g

in which p is pressure, c is the specific weight of the fluid (=qg), z is elevation and
v is local velocity. The terms p/c and z represent potential energy and V2/2g rep-
resents kinetic energy.
The limitation of irrotational flow means that the energy conservation Eq. (1.10)
with H defined by Eq. (1.11) cannot be applied across streamlines within a (rota-
tional) boundary layer. The common application to cross-section average flow
conditions (v = V) at successive locations in the flow direction is acceptable,
however, even with rotational flow. Note that energy is a scalar quantity and is
therefore unaffected by changes in flow direction.
The derivation of Eq. (1.11) included a division of the energy components by c,
meaning that the final terms represent the energy per unit weight of fluid; the total
energy therefore depends on velocity, but not on discharge. The consequent
dimensions of the terms in the Bernoulli equation are therefore length, and the terms
are commonly referred to as ‘heads’. Having length dimensions enable the terms to
be represented graphically, as shown in Fig. 1.6. The elevation and pressure head
are commonly referred to in combination as the piezometric head, indicating the
location of the hydraulic grade line (HGL).
Sketching energy lines is extremely useful for understanding and analysing
variations of flow velocity and pressure in hydraulic systems. For example, the
energy lines shown in Fig. 1.7 for flow through a pipe contraction enable the
variations of velocity and pressure to be visualized easily.
1.3 The Conservation Laws in Hydraulics 9

Total Energy Line, H


velocity
V2/2g head
Hydraulic Grade Line,
HGL
pressure
p/γ
head total
piezometric head
Pipe centre line head
z

Datum

Fig. 1.6 Energy terms and energy lines for flow in a pipe

Fig. 1.7 Energy lines for H


flow through a pipe V12 /2g
contraction V2 2 /2g

HGL
p1 /γ
p2 /γ

z1 z2

1 2

Datum

In most problems, energy conservation is applied between sections in the


direction of flow. It is then convenient to express the kinetic energy term of the
Bernoulli equation in terms of the average velocity over the section, V = Q/A.
Because of the nonuniformity of flow within boundary layers, this will give a result
different from that obtained by integration over the section, and will therefore be
inaccurate. To account for this, a correction factor called the Coriolis coefficient or
kinetic energy correction factor (a) is applied to the mean velocity head, which then
becomes aV2/2g. A value for this factor can easily be determined if the velocity
distribution is known from
R
v3 dA
a¼ ð1:12Þ
V 3A

For laminar flow, a has a value of about 2.0. Laminar flow is rare in practical
problems although it may occur in small-scale physical models of hydraulic
structures. For turbulent flow, the value averages about 1.06 and rarely exceeds
about 1.15, and is often assumed to have a value of 1.0 in practical calculations.
For open channel applications, the Bernoulli equation is usually modified
slightly. If the pressure distribution is hydrostatic, then the hydraulic grade line,
10 1 Basic Hydraulic Concepts

Fig. 1.8 Energy terms and


energy lines for flow in an
open channel αV 2/2g H
E
p/γ
HGL
.A y cos θ
y

z θ
z/

Datum

which represents the piezometric head (z + p/c) at any point on a vertical (such as A
in Fig. 1.8), coincides with the water surface. In this case, the Bernoulli equation
can also be written

V2
H ¼ z0 þ y cos h þ a ð1:13Þ
2g

where z0 is the height of the bed level above the datum, y is the flow depth measured
normal to the bed and h is the inclination of the bed to horizontal, as shown in
Fig. 1.8.
The specific energy, E, is defined as the total energy relative to the channel bed
as datum, i.e.

V2
E ¼ y cos h þ a ð1:14Þ
2g

This concept was introduced by Bakhmeteff in 1912. It rests heavily on the


assumption of a hydrostatic pressure distribution, and it is important to identify
conditions where this assumption might not hold.
The cos h term is usually ignored because bed slopes are small and it is then
close to 1.0. Chow (1959) points out that this correction amounts to less than 1%
until the slope exceeds 6°, or about 0.10. For greater slopes, cos h should be
included. Another difficulty associated with steep slopes is that velocities are
usually high, and flowing water entrains air when velocities exceed about 6 m/s.
This can be clearly seen on spillway faces, where the flow becomes aerated quite
suddenly some distance below the crest. When this happens, ‘bulking’ occurs and
the flow depth increases. Predicted pressures are then higher than observed. If air
entrainment is expected, the water density used in the calculations should be that of
the air–water mixture.
1.3 The Conservation Laws in Hydraulics 11

hydrostatic
p

Fig. 1.9 Pressure distribution at a free overfall

(a) (b)
hydrostatic
p
hydrostatic
p

Fig. 1.10 Pressure distributions in convex (a) and concave (b) flow

The pressure distribution will also not be hydrostatic if there is distinct curvature
of the flow. An important instance of non-hydrostatic pressure is in the drawdown
to critical flow at a free overfall, shown in Fig. 1.9. In addition to the strong vertical
curvature, aeration of the underside of the nappe causes the pressure to be atmo-
spheric at the brink.
The geometry of a structure can also induce curvilinear flow, for example, at the
base of a spillway. In this case, the flow would be concave, but convex flow can
also occur. These situations are illustrated in Fig. 1.10.
In concave flow, the centrifugal forces reinforce gravity and the pressures are
greater than hydrostatic. In convex flow, the centrifugal forces oppose gravity and
the pressures are less than hydrostatic.
The actual pressure can be determined by adding a centrifugal component to the
hydrostatic. The approximate centrifugal pressure can be computed as the product
of the mass of a column of water with unit area (cy/g) and the centrifugal accel-
eration (v2/r, where r is the radius of curvature). Dividing by c gives this pressure as
a head. The total pressure head is then

y v2
h ¼ hs  ð1:15Þ
g r

in which hs is the hydrostatic pressure head and the + and − signs correspond to
concave and convex flow, respectively.
12 1 Basic Hydraulic Concepts

For simplicity, the pressure head for curvilinear flow is often represented by a0
ycos h, where a0 is a pressure coefficient. It can be shown that a0 is given by
Z
1
a0 ¼ 1 þ c v dA ð1:16Þ
Qy

in which c is the centrifugal term in Eq. (1.15), i.e. c = y/g v2/r.


For complicated curved profiles, the pressure distribution should be determined
using ideal flow theory (flow nets), numerical modelling or model testing.

1.3.3 Conservation of Momentum

Integration of Newton’s second law with respect to time leads to a formulation of


momentum conservation. For a solid object, this shows that a force acting on the
object for a certain time duration will result in a change in its momentum. For
fluids, the continuum of fluid and continuous application of forces must be con-
sidered. A force applied to a fluid (F) causes a change in its momentum flux or the
rate at which momentum passes a section. For steady, incompressible flow, this can
be expressed as
X  
Fx ¼ q Q Vfx  Vix ð1:17Þ

which is known as the force–momentum flux equation. The subscript x refers to the
direction considered, Vxf and Vxi are the final and initial velocities in this direction.
The final and initial locations are the inflow and outflow sections of an identified
free body or control volume of fluid. The forces in the prescribed direction include
pressure forces at the initial and final locations, boundary forces between these
locations, and the weight of the free body if the flow is not horizontal.
The force–momentum flux equation applies to the whole flow under considera-
tion, not to a unit weight of fluid as does the Bernoulli equation, because it involves
discharge as well as velocity. It has the advantage that it deals with overall influences
on the free body of fluid defined by the bounding geometry and the inflow and
outflow sections—the distribution of forces between the sections is not of concern,
only the net effect. Note also that because the equation is in terms of vector quantities,
direction is an important consideration—energy does not change as flow goes around
a bend, but momentum flux does. Evaluating the total force between the inflow and
outflow sections may require application of the equation in mutually perpendicular
directions to establish the force components before their vector addition.
As in the energy equation, the use of an average velocity (V = Q/A) will
introduce an error if the flow is not uniformly distributed over the section.
A correction factor (b) to be applied to the velocities can be determined from the
velocity distribution as
1.3 The Conservation Laws in Hydraulics 13

R
v2 dA
b¼ ð1:18Þ
V2 A

For laminar flow, b has a value of about 1.33. For parallel turbulent flow in open
channels, the value is about 1.05. As this can be considered close to 1.0 in most
practical applications, it is often assumed to be equal to 1.0 and may not appear in
the equation.
For open channels, the sum of forces will include pressure forces at the sections
at the beginning and end of the free body to which Eq. (1.17) is applied. If these
sections are both located in flow regions where streamlines are straight and parallel,
then these forces may be calculated from the hydrostatic pressure distribution. If the
flow is curvilinear at either or both sections, then the pressure forces must be
corrected. Chow (1959) recommends multiplying the hydrostatic pressure force by
a force coefficient b0 , given by
Z
1
b0 ¼ 1 þ c dA ð1:19Þ
Az

in which z is the depth of the centroid of the flow section below the water surface
and c = y/g v2/r, as used for the correction factor a0 in Eq. (1.16).
In open channel applications, the net force in the force–momentum flux equation
will always include pressure forces at the end sections. If these are hydrostatic,
Eq. (1.17) can be expressed as

P
¼ M2  M1 ð1:20Þ
c

in which section 2 is downstream of section 1, P is the net force applied to the free
body between the sections in the direction of flow and M is the momentum function,
defined by

Q2
M¼ þAz ð1:21Þ
gA

in which z is the depth of the section centroid below the water surface. Note that
M includes the hydrostatic forces at the ends of the free body.
For flow over unit width M is defined as

q2 y2
M¼ þ ð1:22Þ
gy 2

in which q is the discharge per unit width. If the unit width definition is used in
Eq. (1.20), then P is the force per unit width.
The form of the momentum equation as expressed by Eqs. (1.20)–(1.22) has no
great practical advantage. It just means that the pressure forces at the sections do not
14 1 Basic Hydraulic Concepts

need to be accounted for separately, which makes some applications more efficient.
For nonprismatic channels, it is often easier to use the form of Eq. (1.17) directly.
Applications of the conservation laws for analysing the flow situations identified
in Fig. 1.2 are discussed in the following sections. Attention is limited to steady
flow, as most hydraulic structures can be analysed under this assumption.

1.4 Steady Uniform Flow and Flow Resistance

Steady uniform flow is the most fundamental flow condition in channels and, at
least conceptually, the easiest to analyse. It also provides a good opportunity for
developing resistance relationships that can also be used in nonuniform situations—
under steady uniform flow, the rate of energy dissipation through resistance mat-
ches the rate at which it is made available by the channel gradient, both of which are
constant. The energy loss term in Eq. (1.10) is then defined by the channel slope,
and this is related to the flow velocity by an empirically based resistance equation.
Steady uniform flow is the type that occurs for a steady discharge in a long
channel with constant geometry and roughness characteristics. The engineering
problem is to establish the relationship between discharge and flow depth. This is
necessary for designing a canal to convey a certain discharge, where the flow depth
would determine the size of channel required. The same type of relationship can be
used inversely to determine the flow capacity of a channel with known character-
istics. For this condition, the physical characteristics that determine the flow depth
are those that are associated with the surface resistance at the boundaries. The
relationships to be developed describe the influence of boundary surface resistance,
and will be used wherever flow resistance must be accounted for, not just for steady
uniform flow.
The relationship between discharge and flow depth is an expression of the
continuity equation (1.9), in which the cross-sectional flow area (A) and the
depth-averaged velocity (V) depend on the flow depth (y) and the channel char-
acteristics, i.e.

Q ¼ AðyÞ Vðy; channel characteristicsÞ ð1:23Þ

where A(y) is defined by the cross-sectional geometry of the channel and the
relationship V(y, channel characteristics) is determined by the resistance of the
channel to flow. As well as matching energy dissipation and channel gradient,
uniform flow can be interpreted as an equilibrium situation in which the down-
stream component of the weight of a flow element driving its motion is exactly
balanced by the shear stress on the boundary resisting its motion (Fig. 1.11). The
driving force depends on the weight of the water and hence the flow depth (yo, with
the subscript o indicating uniform flow) and the resisting force depends on the
velocity—otherwise, the element would not be in equilibrium at a particular
velocity. The required relationship V(y, channel characteristics) can be determined
1.4 Steady Uniform Flow and Flow Resistance 15

Sf P
V2/2g
L H
Q Sw
yo FW HGL
FH1
FR FH2 x
So
θ
W
1 2

Fig. 1.11 Motion of an element in uniform flow

by first evaluating the resisting boundary shear stress (so) in terms of the channel
characteristics and flow depth (so = f(yo, channel characteristics)), formulating an
independent relationship between the shear stress and velocity (so = f(V)), and then
combining these relationships by eliminating so.

1.4.1 The General Resistance Equation

The relationship between the resisting stress and channel characteristics, i.e. so = f
(yo, channel characteristics), can be obtained by analysing the motion of an element
isolated in uniform flow (Fig. 1.11).
Because the flow is uniform, the flow depth and velocity do not change in the
flow direction. Therefore, the water surface is parallel to the bed, i.e. Sw = So and,
because the energy line is a distance V2/2g above the water surface (at the hydraulic
grade line), the energy line is also parallel to the water surface, so Sf = Sw = So.
The motion of the free body can be analysed by applying Newton’s second law,

F ¼ ma

where F is the net force acting on the free body, m is its mass and a is its
acceleration in the direction of F.
There are four forces comprising F acting on this free body:
– The component of its weight (W) in the flow direction, given by

FW ¼ W sin h

W is the product of density (q), gravitational acceleration (g) and volume AL.
Because channel slopes are generally very small, sin h can be approximated by
tan h, which is the bed slope So. Therefore
16 1 Basic Hydraulic Concepts

FW ¼ W sin h ¼ q g A L So

– The resisting force arising from a shear stress (with average value so) over the
contact surface between the water and the channel is given by

FR ¼ so P L

– Hydrostatic pressure forces on the upstream and downstream faces of the free
body (FH1 and FH2). Because the flow is uniform and the flow depths are equal,
these forces act in opposite directions and are equal in magnitude and therefore
cancel.
Uniform flow implies that the flow depth and velocity are constant in the flow
direction, and there is therefore no acceleration. The net force on the element is
therefore zero, and so

FR þ FW ¼ 0

i.e.

so P L þ q g A L So ¼ 0

and so

so P L ¼ q g A L S o

The average shear stress on the boundary can therefore be expressed as

A
so ¼ q g So
P

The quantity A/P is the hydraulic radius, R. Therefore

s o ¼ q g R So

or, using c ¼ qg;

s o ¼ c R So ð1:24Þ

This is a particular solution for uniform flow. A similar relationship can be


derived for steady flow generally, i.e. including nonuniform flow where
So 6¼ Sw 6¼ Sf, where the hydrostatic forces on the upstream and downstream faces
of the free body do not cancel and the acceleration is not zero. This gives the
general result
1.4 Steady Uniform Flow and Flow Resistance 17

s o ¼ c R Sf ð1:25Þ

This general equation always applies. If the flow is known to be uniform, then it
is known that Sf is numerically equal to So and Eq. (1.24) applies. The resistance
equations to follow include slope, which is conventionally designated as S, without
subscript. This is always Sf and is also So if the flow is uniform. The same equations
are used to evaluate Sf in gradually varied nonuniform flow analysis. As well as
providing the basis for resistance equations, Eqs. (1.24) and (1.25) are useful for
evaluating shear stresses for assessing sediment erosion in rivers and for deter-
mining u* as required for describing vertical velocity profiles or classifying tur-
bulent flow through Re*.
Formulating the required relationship V(y, channel characteristics), also requires
an independent relationship between the shear stress and velocity (so = f(V)). This
was first done by Antoine Chézy in 1768.

1.4.2 The Chézy Equation

Chézy used dimensional analysis to relate the bed shear stress and velocity as

so ¼ a q V 2 ð1:26Þ

in which a is a dimensionless coefficient which could depend on channel


characteristics.
Equating the right-hand sides of Eqs. (1.25) and (1.26) gives

so ¼ a q V 2 ¼ c R S

from which
rffiffiffiffiffiffiffiffiffiffiffi
g
V¼ RS ð1:27Þ
a

This is usually written as


pffiffiffiffiffiffiffi
V ¼C RS ð1:28Þ

and is known as the Chézy equation, with C known as the Chézy resistance
coefficient. This enables calculation of the flow velocity from the channel slope,
cross-section dimensions and flow depth, provided a value for C is known. With
both the flow depth and velocity known, the discharge can be simply calculated
from the continuity equation.
18 1 Basic Hydraulic Concepts

1.4.3 The Darcy–Weisbach Equation

Through empirical work and dimensional considerations by Henri Darcy and Julius
Weisbach in the mid-nineteenth century, Eq. (1.29) was developed for relating the
friction loss (hf) in a pipe to its length (L), diameter (D) and roughness.

f L V2
hf ¼ ð1:29Þ
2gD

The friction factor, f, is related to the physical roughness of the boundary and the
turbulence characteristics of the flow. The friction factor is dimensionless, making
the equation more theoretically appealing than the Chézy equation.
Noting that hf /L = Sf and D = 4R, Eq. (1.29) can be reformulated to provide a
resistance equation for open channel applications, i.e.
sffiffiffiffiffiffi
8 g pffiffiffiffiffiffiffi
V¼ RS ð1:30Þ
f

Experimental work in the first half of the twentieth century produced ways for
estimating values for the friction factor, in particular, the Colebrook–White equa-
tion and the Moody diagram. The Moody diagram (Fig. 1.12) shows the relation-
ship between the friction factor, Reynolds number and the relative roughness,
ks/D. Although the diagram and the corresponding equations were developed for
pipe data, the effect of cross-section geometry has been found to be small and they
apply for channels as well, as long as the Reynolds number is defined as 4VR/m and
the relative roughness as ks/4R.
Comparing Eqs. (1.28) and (1.30) shows that the Chézy and Darcy–Weisbach
equations are equivalent, with
sffiffiffiffiffiffi
8g
C¼ ð1:31Þ
f

so the results for f can easily be used to calculate values of C.


The Moody diagram shows the variation of f in two major zones, for laminar
flow (Re < 2000) and turbulent flow (Re > 4000). The turbulent zone can be fur-
ther subdivided into zones of hydraulically smooth (Re* < 5), transitional
(5 < Re* < 70), and hydraulically rough (Re* > 70) flow. The variations in each of
the zones have been represented by equations, which are easier and more accurate
to apply in practice than the diagram.
In the laminar flow region, f is independent of relative roughness and inversely
proportional to Reynolds number and can be represented by
1.4 Steady Uniform Flow and Flow Resistance 19

laminar turbulent
0.10
0.08
5·10-2
0.06
rough
0.04 10-2
transitional
0.03
ks/4R
f laminar
0.02 10-3

10-4
0.01
10-5

103 104 105 106 107 108


Re = 4VR/ν

Fig. 1.12 Moody diagram for friction factor

K
f ¼ ð1:32Þ
Re

For pipes K = 64, but different values have been reported for free surface flow.
Free surface laminar flow is not common in engineering practice, and further details
are not provided here.
In the turbulent flow region, different equations describe the variation of f with
Re and relative roughness in the different zones.
For hydraulically smooth flow (Re* < 5), two different equations apply over dif-
ferent ranges of Re. The Blasius equation describes the variation for Re < 105, i.e.

0:316
f ¼ ð1:33Þ
Re0:25

For Re > 105, the variation is described by


pffiffiffi
1 Re f
pffiffiffi ¼ 2 log ð1:34Þ
f 2:51

Having these two equations means that if flow is hydraulically smooth


(Re* < 5), Re must still be determined to enable the better equation to be selected
and used.
20 1 Basic Hydraulic Concepts

Note that for hydraulically smooth flow f depends on Re but not on the surface
roughness. This means that provided the roughness elements are completely sub-
merged within the viscous sublayer, their size does not influence the resistance.
This does not apply for transitional flow. The shapes of the curves on the Moody
diagram indicate that f depends on both Re and ks. The corresponding equations
must reflect this.
For transitional flow (5 < Re* < 70), the Colebrook–White equation applies. For
pipes, this is
 
1 ks 2:51
pffiffiffi ¼ 2 log þ pffiffiffi ð1:35Þ
f 3:7 D Re f

Substituting D = 4R to make the equation applicable for channels gives


 
1 ks 2:51
pffiffiffi ¼ 2 log þ pffiffiffi ð1:36Þ
f 14:8 R Re f

Note that if Re is defined as VR/m, then the number in the last term will be
different from 2.51; only for Re = 4VR/m it is 2.51.
It has been found empirically that for channels, the coefficients should be slightly
modified, giving
 
1 ks 2:5
pffiffiffi ¼ 2 log þ pffiffiffi ð1:37Þ
f 12 R Re f

In this case, f depends on both Re and ks. Note that the second term in the
transitional equation corresponds to hydraulically smooth flow and the first term to
hydraulically rough flow (to follow below).
For hydraulically rough flow (Re* > 70), f depends only on the relative
roughness, and not at all on Re (as can be seen on the Moody diagram). The Re
term in the modified Colebrook–White equation therefore falls away, giving

1 12 R
pffiffiffi ¼ 2 log ð1:38Þ
f ks

The transitional equation can be used in this zone, but the Re term will become
negligibly small as flow becomes fully turbulent.
Equations (1.33) to (1.38), and the Moody diagram, show that f and hence
C depends on the shape and size of the conduit (through R) and on the flow
condition (through Re) as well as the surface roughness. Even for hydraulically
rough flow, f and C are not constant for the same channel as the discharge and
hence the relative roughness varies. ks, however, is a representation of the channel
roughness that is independent of flow conditions and is therefore a more reliable
characterization than direct estimates of f or C. Some representative values of ks are
given in Table 1.1.
1.4 Steady Uniform Flow and Flow Resistance 21

Table 1.1 Values of ks for Surface ks (mm)


common surfaces
PVC (plastic) 0.01–0.02
Steel pipe 0.045
Very smooth concrete 0.15
Smooth cement-plastered surfaces with flush 0.30
joints
Steel-trowelled concrete 1.5
Unfinished concrete 7
Shotcrete 14

Calculating f for smooth (using Eq. (1.34) and transitional flow (Eq. (1.37) requires
iterative solution because it appears on both sides of the equations as they stand. For
many problems, it is useful to make the equations explicit by substituting Eq. (1.30) for
V in the Reynolds number. The f terms then cancel, and the smooth term becomes

2:5 2:5 m
pffiffiffi ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffi ð1:39Þ
Re f 4 R 8 g R S

allowing an explicit solution. Iterative solution cannot be avoided when using


Eq. (1.33), but its applicable range of Reynolds number is unlikely to occur in
practical engineering applications.
Solutions for problems where the flow depth is known and the discharge is
required are straightforward. The flow regime can be determined directly by cal-
culating Re*, enabling the applicable equation for f to be identified. If the flow
regime is smooth, the Reynolds number should also be calculated to enable
selection between Eqs. (1.33) and (1.34). The flow velocity can then be calculated
using Eq. (1.30), and the discharge from the continuity equation.
Example 1.1
A long steel flume on a slope of 0.0030 has a rectangular cross section with a width of
0.50 m. Determine the discharge when the uniform flow depth is 0.30 m.
Solution
The discharge is given by the continuity equation
22 1 Basic Hydraulic Concepts

Q ¼ AV
A ¼ W yo ¼ 0:50  0:30 ¼ 0:15 m2
sffiffiffiffiffiffi
8 g pffiffiffiffiffiffiffiffiffi
V¼ R So
f
A A 0:15
R¼ ¼ ¼ ¼ 0:136 m
P W þ 2 yo 0:50 þ 2  0:30
The equation to use for f depends on Re:
u ks
Re ¼
m pffiffiffiffiffiffiffiffiffiffiffiffi
u ¼ g R So
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
¼ 9:8  0:136  0:0030 ¼ 0:063 m/s
ks for steel ¼ 0:045 mm ðTable 1:1Þ
m for water  1  106 m2 =s

Therefore

0:063  0:045  103


Re ¼ ¼ 2:8
1  106
Re* < 5, so flow is smooth. f is given by either Eqs. (1.33) or (1.34), depending on Re,
which is unknown as yet. Assume Eq. (1.34) applies, and check Re when V is known,
i.e.
pffiffiffi
1 Re f
pffiffiffi ¼ 2 log
f 2:5

and making the substitution from Eq. (1.39), i.e.

2:5 2:5 m
pffiffiffi ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffi gives
Re f 4 R 8 g R S
pffiffiffiffiffiffiffiffiffiffiffiffiffi
1 4R 8 g R S
pffiffiffi ¼ 2 log
f 2:5 v
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
4  0:136 8  9:8  0:136  0:0030
¼ 2 log ¼ 9:18
2:5  1  106
Therefore
pffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
V ¼ 9:18 8  9:8 0:136  0:0030 ¼ 1:64 m/s

Check Re:

4 V R 4  1:64  0:136
Re ¼ ¼ ¼ 8:9  105
m 1  106
Re > 105, so Eq. (1.34) is correct.
1.4 Steady Uniform Flow and Flow Resistance 23

Therefore
Q ¼ 0:15  1:64 ¼ 0:25 m3 =s
(Note that the Chézy C and Manning n can be calculated from f using Eqs. (1.31) and
(1.41.) So

1 pffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffi
C ¼ pffiffiffi 8 g ¼ 9:18  8  9:8 ¼ 81
f

and
R1=6 0:1361=6
n¼ ¼ ¼ 0:0089Þ
C 81

In problems where discharge is known and flow depth is required, it is not


possible to establish the flow regime until the flow depth is known. The equations
for f also require a value for R, which depends on the flow depth. An iterative
solution is therefore required, as follows:
– assume a value for yo,
– use this yo to calculate R and A from the channel cross-sectional geometry and
V from continuity,
– calculate Re = 4VR/m and Re* = (gRS)1/2ks/m,
– calculate f using the appropriate equation,
– calculate V using the Darcy–Weisbach equation and this f,
– calculate Q = AV and
– compare this Q with the known value, and adjust yo until they are equal.
Most practical engineering work involves channels which are large enough that
flow is invariably in the hydraulically rough turbulent flow regime. It is therefore
convenient to assume this condition, as its equation for f is the easiest to use, and
then to check the regime and revise the calculations if necessary. (For hydraulically
rough flow, the Manning equation (to follow) is even simpler and usually used.)
Example 1.2
A long drainage channel on a slope of 0.0010 has a 0.60 m wide rectangular cross
section lined with steel-trowelled concrete. Determine the flow depth when the discharge
is 0.12 m3/s.
Solution
The flow depth must be found by iteration. Only calculations for the final (correct) trial
value are shown.
Assume yo = 0.30 m. The discharge for this depth is calculated and compared with the
specified value.
The discharge is given by the continuity equation
24 1 Basic Hydraulic Concepts

Q ¼ AV
A ¼ W yo ¼ 0:60  0:30 ¼ 0:18 m2
sffiffiffiffiffiffi
8 g pffiffiffiffiffiffiffiffiffi
V¼ R So
f
A A 0:18
R¼ ¼ ¼ ¼ 0:15 m
P W þ 2 yo 0:60 þ 2  0:30
The equation to use for f depends on Re :
u  ks
Re ¼
m pffiffiffiffiffiffiffiffiffiffiffiffi
u ¼ g R So
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
¼ 9:8  0:15  0:0010 ¼ 0:038 m/s
ks for steel-trowelled concrete ¼ 1:5 mm
ðTable 1:1Þ
m for water  1  106 m2 =s

Therefore

0:038  1:5  103


Re ¼ ¼ 57
1  106
Re* is between 5 and 70 and so the flow is transitional. f is given by Eq. (1.37), i.e.
 
1 ks 2:5
pffiffiffi ¼ 2 log þ pffiffiffi
f 12 R Re f

and making the substitution from Eq. (1.39), i.e.

2:5 2:5 v
pffiffiffi ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffi gives
Re f 4R 8 g R S
 
1 ks 2:5 m
pffiffiffi ¼ 2 log þ pffiffiffiffiffiffiffiffiffiffiffiffiffi
f 12 R 4 R 8 g R S
 
0:0015 2:5  1  106
¼ 2 log þ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
12  0:15 4  0:15 8  9:8  0:0010
¼ 6:14

Therefore
pffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
V ¼ 6:14 8  9:8 0:15  0:0010 ¼ 0:67 m/s

and

Q ¼ 0:18  0:67 ¼ 0:12 m3 =s

The calculated discharge is correct. If it were not then a new value for yo would be
selected and the calculations repeated.
(Again the Chézy C and Manning n can be calculated from f using Eqs. (1.31) and
(1.41.) Then
1.4 Steady Uniform Flow and Flow Resistance 25

1 pffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffi
C ¼ pffiffiffi 8 g ¼ 6:14  8  9:8 ¼ 54
f

and
R1=6 0:151=6
n¼ ¼ ¼ 0:0135
C 54

which is slightly higher than the tabulated value for steel-trowelled concrete because the flow
is transitional and not hydraulically rough.

Use of the Moody diagram, or the corresponding equations, to determine


C makes the Chézy and Darcy–Weisbach equations exactly equivalent. The Chézy
equation is still widely used for historical reasons and because the range of flow
conditions encountered in channels allows some simplifications to be made,
allowing C to be defined more simply than through these equations. For the most
common condition of hydraulically rough turbulent flow
pffiffiffiffiffiffi 1
C ¼ 8 g pffiffiffi
f
 
pffiffiffiffiffiffi 12 R
¼ 8 g 2 log ð1:40Þ
ks
12 R
¼ 18 log
ks

It is important to remember when using this equation (and others) how it was
developed, so that it is used appropriately. Equation (1.40) should be used only if
the flow is turbulent and hydraulically rough, and if the flow resistance arises only
from shear at the boundary.

1.4.4 The Manning Equation

Before the development of the Darcy–Weisbach friction factor relationships, various


attempts were made to replace the Chézy resistance coefficient, C, with one that is
determined by the surface roughness only, independent of the flow condition. The
Manning equation was developed to provide this, based on the recognition of the
relationship between C and the size and shape of the flow cross section, expressed as

R1=6
C¼ ð1:41Þ
n

in which n is characteristic of the surface roughness only. This can be introduced


into the Chézy equation, which becomes

R1=6 pffiffiffiffiffiffiffi
V¼ RS
n
26 1 Basic Hydraulic Concepts

Table 1.2 Values of Surface n


Manning n for common
surfaces Glass, plastic, machined metal 0.010
Corrugated metal 0025
Cement plaster 0.011
Smooth concrete 0.012
Steel-trowelled concrete 0.013
Float-finished concrete 0.015
Unfinished concrete 0.017
Brick in cement mortar 0.015
Gunite 0.019
Rough asphalt 0.016

leading to
1 2=3 1=2
V¼ R S ð1:42Þ
n

in SI units. This is the Manning equation, with n being the Manning resistance
coefficient.
The Manning equation is not dimensionally homogeneous and the 1 should be
replaced by the conversion factor 1.49 if used with foot-second units.
For lined conduits, n can be related to C or f through Eqs. (1.31) and (1.41), i.e.
sffiffiffiffiffiffiffiffiffiffiffiffi
R1=6 R1=3 f
n¼ ¼ ð1:43Þ
C 8g

so the Moody diagram and associated equations can be used indirectly to quantify
n. Strickler in 1923 proposed a direct relationship between n and ks,
ks1=6
n¼ ð1:44Þ
6:7 g1=2

Because of the widespread use of the Manning equation, tables of values of n for
different common surfaces are found in most hydraulics textbooks. Table 1.2
provides some representative values. Values of n are approximately constant with
depth, unlike f and C.
All the tables of n values implicitly assume that resistance depends only on the
physical roughness. The known dependence on R is accounted for in the form of the
equation, but the dependence on Re is not. The implication of this is that the Manning
equation, using standard tabulated values of n, is only valid for hydraulically rough
turbulent flow, which is the usual case for engineering-scale channels. If this is not the
case, then n should be adjusted or the Darcy–Weisbach equation used with the friction
factor defined by the equations given before.
1.4 Steady Uniform Flow and Flow Resistance 27

Example 1.3
A long channel lined with steel-trowelled concrete has a slope of 0.00080 and a
trapezoidal cross section with a base width of 1.0 m and side slopes (Ss) of 1V:1H.
Determine the discharge when the uniform flow depth is 0.40 m.
Solution
The discharge is given by the continuity equation

Q ¼ AV
 
1 yo
A ¼ W yo þ 2 yo
2 Ss
 
1 0:40
¼ 1:0  0:40 þ 2 0:40 ¼ 0:56 m2
2 1:0
1
V ¼ R2=3 So1=2
n
A

P sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

yo 2
P ¼ W þ2 þ yo 2
Ss
s ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi
0:40 2
¼ 1:0 þ 2 þ 0:402 ¼ 2:13 m
1:0
0:56
R¼ ¼ 0:26 m
2:13
n ¼ 0:013 for steel-trowelled concrete ðTable 1:2Þ

So

1
V¼ 0:262=3 0:000801=2 ¼ 0:89 m/s
0:013

and

Q ¼ 0:56  0:89 ¼ 0:50 m3 =s

This is a fairly small channel and the flow may not be hydraulically rough turbulent,
which is a condition for use of the Manning equation with a fixed value of n. The flow
regime should therefore be checked by calculating Re*.

u ks
Re ¼
m pffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u ¼ g R So ¼ 9:8  0:26  0:00080 ¼ 0:045 m/s

ks can be estimated from Eq. (1.38), assuming that the flow is hydraulically rough, i.e.

1 12 R
pffiffiffi ¼ 2 log
f ks
28 1 Basic Hydraulic Concepts

So
12 R
ks ¼ pffi
101=2 f

with f calculated from n using Eq. (1.43), i.e.


sffiffiffiffiffiffiffiffiffiffiffiffi
R1=3 f

8g

or
8 g n2 8  9:8  0:0132
f ¼ ¼ ¼ 0:021
R1=3 0:261=3
So

12  0:26
ks ¼ pffiffiffiffiffiffiffiffi ¼ 0:0011 m
101=2 0:021

Therefore

0:045  0:0011
Re ¼ ¼ 50
1  106
Re* is less than 70, so the flow is transitional and Manning’s equation with a constant
value of n is strictly not applicable. The flow condition can be located on the Moody
diagram with

4 V R 4  0:89  0:26
Re ¼ ¼ ¼ 9:3  105
m 1  106
and
ks 0:0011
¼ ¼ 1:1  103
4 R 4  0:26
This point plots on Fig. 1.12 just within the transitional zone. However, the increase in
f (and hence n) within the transitional zone is gradual towards the rough zone and the
error in using the fixed value of n will be small and within the uncertainty in the estimate
of its value.

The Manning equation is the most widely used for natural channels, where
boundary shear is not the only source of flow resistance. Resources are available for
estimating Manning’s n for natural channels, but are beyond the scope of this text.

1.5 Steady Rapidly Varied Flow

1.5.1 Application of Energy and Momentum Conservation

Rapidly varied flow is the class of nonuniform flow in which changes in flow
conditions take place over short distances. In this context, ‘short’ is taken to mean
1.5 Steady Rapidly Varied Flow 29

distances over which friction losses are negligibly small. In the absence of abrupt
features presenting expansion losses or drag forces, the loss term in the energy
conservation equation is then usually ignored. In most applications, it can also be
assumed that the bed slope is small, so cos h  1.0, the flow is approximately
parallel, so the pressure distribution is hydrostatic, and the effect of the velocity
distribution through the depth is negligible, so a  1.0. The specific energy rela-
tionship (Eq. (1.14) can then be written as
V2
E ¼ yþ ð1:45Þ
2g

If there are departures from these assumptions, appropriate corrections should be


made.
It is usually convenient to incorporate the continuity equation into the specific
energy, to express it in terms of discharge, rather than velocity, i.e.

Q2
E ¼ yþ ð1:46Þ
2 g A2

For rectangular channels, this is often expressed in terms of the unit width
discharge, q (=Q/b, where b is the channel width), i.e.

q2
E ¼ yþ ð1:47Þ
2 g y2

Equation (1.47) leads to some useful results for rectangular channels and is also
a convenient form for explaining some general principles in free surface flow. The
equation is cubic in y, implying that three values may exist for a given discharge
and specific energy pair. One of these is negative and clearly of no practical interest.
The relationship between the two positive flow depths and specific energy for a
given value of q is shown in Fig. 1.13.
It is clear from Fig. 1.13 that for a given value of q, there is a wide range of
conditions for which two distinctly different flow depths (known as alternate
depths) would have the same value of specific energy. The upper limb of the curve
represents relatively deep, slow subcritical flow while the lower limb represents
relatively shallow, fast supercritical flow. The energy of subcritical is mainly
potential, while that of supercritical flow is mainly kinetic.
Figure 1.13 can be used to explain the response of a flowing water surface to a
change in bed elevation, such as shown in Fig. 1.14. Surface resistance energy loss
is negligible over the short distance between the sections, and if the step is smooth,
there will also be negligible expansion loss, so

H1 ¼ H2 ¼ H3
30 1 Basic Hydraulic Concepts

q
y y + V2/2g

alternate depths

y + V2/2g

Fig. 1.13 Specific energy relationship

From the definition of specific energy, this can be expressed as

E1 ¼ E2 þ Dz ¼ E3

If the flow conditions at one of the sections (say subcritical flow at section 3) is
known, then it is a straightforward matter to determine the value of E at that section
(from Eq. 1.47) and, from the above relationship, at sections 1 and 2. For small Dz,
solution of the energy relationship at section 2 will yield two possible values of
y. Considering that all flow conditions between sections 1 and 2 must lie on the given
curve (unless q is changed by a change in width), and that the decrease in E is limited
by the magnitude of Dz, the flow at the two sections must be represented by points on
the same limb of the E-y curve, i.e. flow cannot be changed from subcritical to
supercritical (or vice versa) by a simple upward step. For a small value of Dz, the flow
depth will decrease from section 1 to section 2, and then increase back to the original
value from section 2 to section 3, following the curve (Fig. 1.14a). A flow regime
change can only be induced by an upward step large enough to induce flow at the
turning point of the curve (critical flow), followed by a downward step. If the vertical
step is large enough, the subtraction of Dz from E1 will produce an infeasible value of
E2, i.e. a value for which there is no value of y for the specified discharge. In this case,
the flow will ‘choke’, and the flow at section 1 will back up until E1 is large enough for
Dz to be subtracted to give a feasible value of E2. The flow at section 2 will then be
critical, i.e. at the turning point of the E-y curve, and the subsequent downward step
will enable supercritical flow to occur at section 3 (provided that this is permitted by
conditions further downstream) (Fig. 1.14b).
If the undisturbed flow in the channel were supercritical and Dz relatively small,
the flow depth would increase from section 1 to section 2 and reduce again at
1.5 Steady Rapidly Varied Flow 31

(a) y q
H

E1 E2 E3
Δz

Δz
1 2 3 2 1 E
3

backup
y q
(b)
H
E2 = Ec
E1 E3 Δz

Δz
1 2 3 2 1 E
3

y q
(c)
H
E1 E2 E3
Δz

Δz
1 2 3 2 1 E
3

Fig. 1.14 Flow response to changes in bed level

section 3, following the E-y curve on the supercritical limb (Fig. 1.14c). If Dz were
large enough for its subtraction from E1 to produce an infeasible value, the flow
would again choke, and in this case, the flow at section 1 would change to sub-
critical and the profile would again be as shown in Fig. 1.14b. The reason for this
regime change at section 1 is clarified later.
The definition of specific energy implies that curves for higher values of q would
plot to the right of that shown in Fig. 1.13, and that curves for lower values would
plot to the left. The E-y diagram can therefore also be used to explain the response
of the water surface to a lateral channel contraction (Fig. 1.15). If the bed is level
then E1 = E2 = E3. For subcritical flow, a moderate contraction will cause a
reduction in flow depth as the flow condition moves from a lower to a higher value
32 1 Basic Hydraulic Concepts

1 2 3

(a) Plan y

H 1,3
q2
q1 2
E1 E2 E3

E
1 2 3

(b)
backup
y
q2
H 1
E2 = Ec
E1 E3 q1 2
3

E
1 2 3

(c)

Fig. 1.15 Flow response to a lateral contraction

of q, with the same value of E; an expansion to the original width will cause the
flow depth to increase to its original level (Fig. 1.15b). If the width at section 2 is
small enough, the curve for q2 will not extend to the original E value, so no flow
depth is feasible at section 2 for this value of E. The flow will choke, and the flow
depth at section 1 will back up, thereby increasing E until a feasible flow condition
at section 2 is attained; this will be for the minimum value of E on the q2 curve, i.e.
critical flow (Fig. 1.15c). As for the upward step, a lateral contraction cannot induce
a change from subcritical to supercritical flow unless the contraction is sufficient to
induce critical flow, and is followed by an expansion. Again, the diagram can be
used to trace the solution for supercritical flow and again choking would produce
critical flow at section 2 and subcritical flow at section 1 (Fig. 1.15c).
The critical flow condition clearly has great significance. Any feature which
requires greater energy to pass a given discharge then would otherwise be available,
and will induce backing up. If the flow maintains a free surface through the feature,
backing up will be just sufficient to induce critical flow, as this is the condition for
which the specific energy is the least possible for the given discharge. The critical
condition is represented by the turning point of the E-y curve, and the relationship
between discharge and flow depth at this point can be obtained by equating to zero
1.5 Steady Rapidly Varied Flow 33

the derivative of E with respect to y. This gives the following useful relationships
for rectangular channels:
sffiffiffiffiffi
3 q
2
yc ¼ ð1:48Þ
g

2
yc ¼ Ec ð1:49Þ
3

in which the subscript c denotes the critical condition.


For nonrectangular sections, q is not defined and Eqs. (1.48) and (1.49) cannot
be used. In general, the relationship between critical flow depth and discharge
should be obtained from the definition of the Froude number applied at the critical
condition, i.e.

Q2 B
Fr2 ¼ 1 ¼ ð1:50Þ
g A3

in which Fr is the Froude number, B is the surface width of the flow and A is the
cross-sectional area of the flow. The relationship between critical depth and specific
energy should be obtained from the more general form of Eq. (1.49), derived from
Eq. (1.46), giving

A
Ec ¼ yc þ ð1:51Þ
2B

Conservation of energy cannot be applied if there are significant energy losses


unless relationships (usually empirical) are available for quantifying them. Energy
considerations are also not useful for quantifying forces associated with changing
flow conditions. In these cases, application of momentum conservation is useful.
This can be done using the force–momentum flux equation (Eq. (1.17) or the
momentum function relationship (Eqs. (1.20) and (1.21)). As for energy applica-
tions, it is usually assumed that the flow is approximately parallel at the sections
considered, so the pressure distribution is hydrostatic, and the effect of the velocity
distribution across the flow section is negligible, so b  1.0. Again, if there are
departures from these assumptions, appropriate corrections should be made.
Although the use of the momentum function has no great practical advantage
over the force–momentum flux equation, its form can be used conceptually to
understand distinctions between subcritical and supercritical flow behaviour. The
form for rectangular channels (Eq. (1.22)) is a cubic equation in y, rather similar to
the specific energy relationship (Eq. (1.47)), and again there are two positive values
of y with the same value of momentum function for a given value of q, as shown in
Fig. 1.16. These are known as conjugate or sequent depths.
As for the specific energy diagram, the upper limb of the curve represents
subcritical flow and the lower limb represents supercritical flow. The momentum
34 1 Basic Hydraulic Concepts

y y2/2 + q2/gy
q

conjugate (sequent)
depths

y2/2 + q2/gy

Fig. 1.16 Momentum function relationship

characteristics of subcritical flow are dominated by the flow depth, representing the
hydrostatic pressure contribution, and those of supercritical flow arise mainly from
the dynamic component associated with the velocity.
The relationship between momentum function values at successive sections is
given by Eq. (1.20), i.e.

P
¼ M2  M1 ð1:20Þ
c

The horizontal distance between flow conditions plotted for adjacent sections in
Fig. 1.16 therefore represents the force applied to the flow between the sections
divided by c. In Fig. 1.17, for example, the force exerted by the sluice gate on the

P/γ

M
1 2 2 1

Fig. 1.17 Flow response to applied force


1.5 Steady Rapidly Varied Flow 35

flow (which is in the negative flow direction) can be indicated on the momentum
function diagram. The force exerted by the flow on the sluice gate is equal in
magnitude and opposite in direction.
For the case of a simple hydraulic jump, there is no force between the upstream
supercritical and downstream subcritical flows, and the flow depths are therefore a
conjugate pair. (Hydraulic jumps are discussed further in Chap. 6.)

1.5.2 The Control Concept

Most open channel flow problems are concerned with evaluating relationships
between the flow depth (y) and the discharge (Q). Typical problems would be to
calculate the flow depth at a particular location in a channel system for a given
discharge, or to calculate the discharge given the flow depth at a particular point.
For a given value of Q, the flow depth at any point is caused by some physical
characteristics in the channel. This cause needs to be identified its effect predicted.
In the examples of energy principle applications used so far, it has been assumed
that the flow conditions are completely known at one section (e.g. just upstream or
downstream of a step or contraction), i.e. values are known for Q (or q), y and E. In
real problems, all this information may not be known. For example, only Q or E or
y may be known and the others must be determined. (Note that the specific energy
relationship enables any one of these to be calculated if the other two are known.)
Once the flow conditions at one section are known, the conditions at the other
sections can be related to them through energy conservation. This process depends
on being able to define the conditions at the first section. This first section, where
the calculations begin, is the one where the flow conditions are directly related to
their cause. It is called a ‘control’ section.
A control is a feature that determines a unique relationship between flow depth
and discharge. This means that the flow depth for a given discharge (or vice versa)
can be calculated by analysing that feature in isolation—it is not necessary to relate
the flow conditions to those at any other section.
Two control features are already familiar:
– A uniform reach of channel determines the flow depth through boundary
resistance. The control effect is described by a resistance equation (such as
Manning’s) which gives the unique relationship between Q and y.
– A contraction induces critical flow. If it is known that flow is critical, the flow
depth can be related directly to discharge (e.g. by Eq. (1.48) for rectangular
channels).
In a real channel system, there will be many features which could possibly act as
controls. A particular feature may be a control at certain discharges and not at
others. The effect of one control may override the effect of another. A major dif-
ficulty in open channel flow analysis is identifying the active controls.
36 1 Basic Hydraulic Concepts

Q
yo (> yc)

So, n, geometry

Fig. 1.18 Basic channel with resistance control only

Q
yo (> yc)

Δz
So, n, geometry

Fig. 1.19 Basic channel with structure incorporated

As an illustrative example, consider a long channel with known discharge, slope,


roughness and cross-sectional geometry (Fig. 1.18).
The flow will be uniform and only one flow depth is possible—that is deter-
mined by the resistance characteristics of the channel. The effects of these char-
acteristics are described by Manning’s equation, so the flow depth can be
calculated. The channel itself is the control over the whole length. Suppose here that
the uniform flow depth is subcritical.
Now suppose a short, smooth hump is incorporated in the channel (Fig. 1.19).
Is this feature a control? It will certainly modify the flow conditions locally, but
it is a control only if it creates a location where the flow depth can be calculated
independently of the conditions anywhere else in the channel.
If Dz is small, the hump will just cause a local change in depth without affecting
upstream or downstream conditions, as discussed before (Fig. 1.20). The feature is
not a control because the depth change it induces is relative to the depth determined
by the channel, i.e. y2 can only be calculated from y1 or (more correctly) y3, each of

q
y
1,3
H
2

yo Δz

Δz
1 2 3 E

Fig. 1.20 Structure not acting as a control


1.5 Steady Rapidly Varied Flow 37

backup
y q
H 1
yo
Ho Δz
Δz
Eo yc
yo 2
3
Δz Eo E
1 2 3

Fig. 1.21 Structure acting as a control

yo
Q
yc
yo
Δz
So, n, geometry

Fig. 1.22 Extended influence of control structure

which can be calculated from Manning’s equation; y2 cannot be calculated


independently.
If Dz is large enough, the specific energy associated with the uniform flow will
be insufficient to enable the given discharge to negotiate the hump, and no value of
y will satisfy the energy equation, i.e. Eo – Dz is less than the minimum value of
E for the given q. Choking will then occur: the flow at section 1 will back up and
flow will go through critical at section 2 (Fig. 1.21).
The hump is now acting as a control. The flow conditions at sections 1 and 3 can
no longer be determined using Manning’s equation. However, knowing that flow at
section 2 is critical enables this flow depth to be calculated directly. The depths
immediately upstream and downstream can then be calculated relative to this
condition. The flow upstream and downstream may be affected for long distances
through gradually varied flow (Fig. 1.22).
So, the control section is where the flow depth can be calculated directly from
the discharge; if the flow depth calculation requires a flow depth somewhere else
then that section is not a control.
Whether the hump in this example is a control or not depends only on the
magnitude of Dz, but also on the discharge and the channel characteristics. If it is a
control then supercritical flow must be maintained immediately downstream. This
requires that the supercritical flow induced by the feature must have sufficient
momentum to keep the hydraulic jump from encroaching on the control section.
To determine the flow depths around a feature, it is essential to be able to
establish whether or not it is a control for a specified discharge. Here, the hump will
be a control if the uniform flow has insufficient specific energy for the flow to get
38 1 Basic Hydraulic Concepts

over the hump. This can be determined by initially assuming uniform conditions at
section 1; if there is no solution to

E2 ¼ Eo  Dz

then the hump is a control. More quickly, by knowing that the minimum energy is
the critical value, the hump is a control if

Eo  Dz \ Ec

and, for rectangular channels,


sffiffiffiffiffi
3 3 3 q
2
Ec ¼ yc ¼
2 2 g

The effect of a control depends significantly on whether the uniform flow


without it is subcritical or supercritical.
Consider again the channel with the hump, starting off with a small
Dz (Fig. 1.20), and what would happen if Dz were increased enough to become a
control.
As Dz increases, there is just a small change in flow depth at section 2 until Dz is
large enough to reduce E to the critical point. If Dz is then increased further q will
momentarily decrease, causing a slight, local increase in depth. This local distur-
bance to the water surface will propagate upstream as a surface wave, eventually
causing a region of flow with greater depth, lower energy dissipation rate, and hence
increased E. This can only occur if the disturbance can propagate upstream, which it
can do only if the surface wave celerity is less than the flow velocity, i.e. flow is
subcritical. If the channel flow is supercritical it must first change to subcritical
before the extra energy can be accumulated. So, if the channel was steep in the
above example, a value of Dz constituting a control would induce subcritical flow
and a hydraulic jump on its upstream side.
In supercritical flow, flow conditions cannot be affected by anything downstream
—the water ‘doesn’t know what happens downstream’. A control has been defined
as providing a unique relationship between discharge and flow depth. This can only
hold if the flow at the control section is unaffected by the downstream flow, and this
requires there to be supercritical flow immediately downstream of the control
section.
Therefore: Subcritical flow is controlled from downstream.
Supercritical flow is controlled from upstream.
It follows that if a feature is acting as a control, the flow upstream of it will be
subcritical and the flow downstream of it will be supercritical. For example, in
Fig. 1.23, the hump is a control only in the third case (c). In the first case (a), the
flow is subcritical throughout and therefore controlled by the uniform channel
1.5 Steady Rapidly Varied Flow 39

(a) control direction (b) control direction (c) control direction

yo yc
yo

1 2 3 1 2 3 1 2 3

Fig. 1.23 Occurrence and direction of control

characteristics downstream of the hump: y1 depends on y2 which depends on y3,


which is equal to yo. In the second case (b), the flow is supercritical throughout, and
therefore controlled by the uniform channel characteristics upstream: y3 depends on
y2 which depends on y1, which is equal to yo. In the third case (c), y2 = yc and both
y1 and y3 are controlled by and depend on this depth.
Example 1.4
Water flows in a long, 2.50 m wide rectangular channel at a uniform flow depth of
1.07 m and a uniform velocity of 1.68 m/s. Calculate the flow depths immediately
before, on top of, and immediately after a smooth 0.12 m hump on the channel bed.
Solution
H

Vo = 1.68 m/s
yo = 1.07 m
0.12 m
1 2 3

The effect of the hump will depend on whether the flow is subcritical or supercritical.
This is determined by calculating the Froude number:

V 1:68
Fr ¼ pffiffiffiffiffiffi ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ 0:52
gy 9:8  1:07

Fr < 1, so the uniform flow is subcritical.


The hump is ‘smooth’, so energy losses may be neglected. Then

H1 ¼ H2 ¼ H3

and
E1 ¼ E2 þ D z ¼ E3

Whether the hump is a control is determined by assuming y1 = yo and comparing the


consequent value of E2 with Ec, i.e.
40 1 Basic Hydraulic Concepts

E2 ¼ Eo  Dz
Vo2 1:682
Eo ¼ yo þ ¼ 1:07 þ ¼ 1:21 m
2g 2  9:8

So

E2 ¼ 1:21  0:12 ¼ 1:09 m

And

3
Ec ¼ yc
2

with
sffiffiffiffiffi
3 q
2
yc ¼
g

and

q ¼ Vo yo ¼ 1:68  1:07 ¼ 1:80 m3 =s/m

So
rffiffiffiffiffiffiffiffiffiffiffi
3 1:80
2
yc ¼ ¼ 0:69 m
9:8

and

3
Ec ¼ 0:69 ¼ 1:04 m
2
E2 > Ec, so choking does not occur. Note that the maximum hump height for no choking
is

Dzmax ¼ Eo  Ec ¼ 1:21  1:04 ¼ 0:17 m

The flow depth at section 2 is given by the solution of

q2
E2 ¼ 1:09 ¼ y2 þ
2 g y22

i.e.

1:802
1:09 ¼ y2 þ
2  9:8 y22

The two positive solutions to this equation are y2 = 0.88 m (subcritical) and y2 = 0.55 m
(supercritical). Because the flow at section 1 is subcritical and there is only a rise in bed
level between sections 1 and 2, the flow at section 2 must also be subcritical. The flow
depth at section 2 is therefore 0.88 m.
1.5 Steady Rapidly Varied Flow 41

The flow depth at section 3 could be obtained from the second part of the energy
relationship, i.e.

E2 þ D z ¼ E3

and reversing the above calculations. However, the energy relationship also gives E3 = E1
and the flow must be subcritical throughout, so y3 = y1 = yo = 1.07 m.

Example 1.5
The hump height in Example 1.4 is changed to 0.75 m.

a. Calculate the flow depths immediately before, on top of, and immediately after the hump.
b. Calculate the force exerted by the flow on the hump.

Solution

1.73 m 0.69 m 0.34 m yo = 1.07 m


P
0.75 m
1 2 3

a. The hump is now higher, so the possibility of it being a control must be checked.
Following the same procedure as before:
E2 ¼ Eo  Dz ¼ 1:21  0:75 ¼ 0:46 m

which is less than Ec = 1.04 m. The hump is therefore a control, y2 = yc and y1 is no


longer equal to yo but determined by the control. It is found from

E1 ¼ E2 þ D z

with

E2 ¼ Ec ¼ 1:04 m

Therefore

1:802
E1 ¼ 1:04 þ 0:75 ¼ 1:79 ¼ y1 þ
2  9:8 y21

giving the subcritical solution y1 = 1.73 m.


Again, E3 = E1 and because the hump is a control, y3 will be the supercritical
solution of the same equation, i.e. y3 = 0.34 m.

b. The force on the hump is equal to the force by the hump on the flow, which is given by
Eq. (1.20), i.e.
42 1 Basic Hydraulic Concepts

P
 ¼ M3  M1 ðP is in the  ve flow direction)
c

So

P
¼ M1  M3
c

with
q2 y2
M1 ¼ þ 1
g y1 2
1:802 1:732
¼ þ ¼ 1:69 m2
9:8  1:73 2

and
q2 y2
M3 ¼ þ 3
g y3 2
1:802 0:342
¼ þ ¼ 1:03 m2
9:8  0:34 2
Therefore

P
¼ 1:69  1:03 ¼ 0:66 m2
c

So

P ¼ 0:66  9:8  103 ¼ 6:5  103 N/m width

or 6.5 kN/m, and the total force on the hump is

Ptotal ¼ P W ¼ 6:5  2:50 ¼ 16:3 kN

Calculations of rapidly varied flow conditions should strictly proceed in the


direction in which the control is being exercised, i.e. upstream from the control for
subcritical flow and downstream from the control for supercritical flow. For testing
whether a feature is a control or not, it is best to consider the conditions of the
approaching flow without the feature being there. In the illustrative examples used
up to now, the starting condition has been assumed to be uniform flow, but this is
not necessarily so—it could be a condition induced by some other control. For
example, if a second hump is located downstream from another which acts as a
control (Fig. 1.24), whether the second hump is a control or not is determined by
the energy of the flow approaching it, as controlled by the first hump.
So the second hump will be a control if E4 – Dz < Ec. (If the humps are far apart,
a gradually varied profile will occur between them and then y4 > y3 and E4 < E3.) If
the two humps have the same Dz, the second is less likely to be a control than the
1.5 Steady Rapidly Varied Flow 43

Δz

1 2 3 4

Fig. 1.24 Interaction between control features

first one, because the control effect of the first one will have increased the available
energy; the second hump might be a control without the first one in place and not a
control with it in place. If the second hump is larger than the first, its upstream
subcritical flow may submerge the first, which would then no longer be a control.
This would need to be checked using the momentum function. If the subcritical M4
with the second hump as a control is greater than the supercritical M3 with the first
hump as a control, then the subcritical flow will override the supercritical flow at
section 3. If y3 is then greater than yc + Dz, the first hump will not be a control. If
the supercritical M3 is greater than the subcritical M4, then the supercritical flow
will continue over the second hump and it will not be a control. To summarize,
whether a feature is a control or not depends not only on its own potential effect, but
also on the presence of other features upstream or downstream.
Example 1.6
A long, concrete-lined (n = 0.013) channel on a slope of 0.010 has a rectangular cross
section with a width of 4.0 m. The features shown below are installed in the channel a
short distance apart. Determine the flow depths at sections 1 to 6 if the discharge is
40.0 m3/s.

0.80 m 0.90 m

1 2 3 4 5 6

Solution
Check if the first hump is a control, ignoring possible influence of the second hump. It is
not a control if

Eo Dz1 [ Ec
q2
Eo ¼ yo þ
2 g y2o
Q 40:0
q¼ ¼ ¼ 10:0 m3 =s/m
W 4:0
The uniform flow depth can be determined from the Manning equation together with
continuity, i.e.
44 1 Basic Hydraulic Concepts

A 2=3 1=2
Q¼ R S
n
A W yo 4:0 yo
R¼ ¼ ¼
P W þ 2 yo 4:0 þ 2 yo

i.e.
 2=3
4:0 yo 4:0 yo
40:0 ¼ 0:0101=2
0:013 ð4:0 þ 2 yo Þ

From which, by trial, yo = 1.46 m


Therefore

10:02
Eo ¼ 1:46 þ ¼ 3:85 m
2  9:8  1:462

and
sffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffiffi
3 3 3 q
2 3 3 10:02
Ec ¼ yc ¼ ¼ ¼ 3:25 m
2 2 g 2 9:8

Therefore

Eo  Dz1 ¼ 3:85  0:80 ¼ 3:05 m\Ec

The first hump is therefore a control unless overridden by the second.


Check if the second hump is a control with the approach flow determined by the first. It
is not a control if

E4  Dz2 [ Ec

The distance between sections 2 and 4 is short, so

E4 ¼ E3 ¼ Ec þ Dz1 ¼ 3:25 þ 0:80 ¼ 4:05 m

Therefore

E4  Dz2 ¼ 4:05  0:90 ¼ 3:15\Ec

The second hump is therefore a control.


y6 is determined from energy conservation between sections 5 and 6, i.e.

E6 ¼ E5 þ Dz2 ¼ 3:25 þ 0:90 ¼ 4:15 m

Therefore

10:02
E6 ¼ 4:15 ¼ y6 þ
2  9:8y26

From which, taking the supercritical solution, by trial, y6 = 1.35 m.


1.5 Steady Rapidly Varied Flow 45

Check if the control effect of the second hump overrides the control of the first. This will
be the case if

M4 [ M3
q2 y2
M4 ¼ þ 4
g y4 2

y4 is determined from energy conservation between sections 4 and 5, i.e.

E4 ¼ E5 þ Dz2 ¼ 3:25 þ 0:90 ¼ 4:15 m

Therefore

10:02
E4 ¼ 4:15 ¼ y4 þ
2  9:8y24

From which, taking the subcritical solution, by trial,


y4 = 4.14 m
So

10:02 4:142
M4 ¼ þ ¼ 11:0 m2
9:8  4:14 2
Similarly

q2 y2
M3 ¼ þ 3
g y3 2

with

E3 ¼ E2 þ Dz1 ¼ 3:25 þ 0:80 ¼ 4:05 m

Therefore

10:02
E3 ¼ 4:05 ¼ y3 þ
2  9:8y23

From which, by trial, y3 = 1.38 m.


So

10:02 1:382
M3 ¼ þ ¼ 8:3 m2
9:8  1:38 2
Therefore M4 [ M3 and y3 is controlled by the second hump, i.e. y3 = y4 = 4.14 m.
y3 is greater than yc + Dz1 = 2.17 + 0.80 = 2.97 m, so critical flow no longer occurs at
section 2. y2 is now found by energy conservation between sections 2 and 3, i.e.

E2 ¼ E3  Dz1
46 1 Basic Hydraulic Concepts

with

10:02 10:02
E3 ¼ y3 þ ¼ 4:14 þ ¼ 4:44 m
2  9:8y3
2 2  9:8  4:142

Therefore

E2 ¼ 4:44  0:80 ¼ 3:64 m

Therefore
10:02
E2 ¼ 3:64 ¼ y2 þ
2  9:8y22

From which, by trial, y2 = 3.11 m


And since E1 = E3 and both are subcritical, y1 = y3 = 4.14 m.
The water surface profile over the humps is therefore as shown below:

4.14 m 3.11 m 4.14 m 4.14 m 3.25 m


1.35 m

1 2 3 4 5 6

In these applications, it has been assumed that Q is known and the flow depths
are required. Measuring structures act as controls to enable Q to be determined from
a measurement of flow depth (Chap. 7).
An important problem involving the identification of control features and using
control relationships is that of determining the discharge from a reservoir into a
long, uniform channel (Fig. 1.25).
The discharge depends on what is controlling the flow. The control is what limits
the amount of water leaving the reservoir.
Two cases need to be considered:
1. If the channel is steep enough for the flow in the channel to be supercritical, then
the entrance is just like a very large upward–downward step (Fig. 1.26). Then
flow will be critical at the entrance, i.e. the entrance is the control section, and
the critical flow relationships can be used to calculate the discharge. Note that as
long as critical flow occurs at this section, the discharge is determined only by
the characteristics of this section. It does not depend at all on the characteristics
of the channel, because supercritical flow cannot be influenced from

Fig. 1.25 Flow from a


reservoir into a channel H

Q?
So, n, geometry
1.5 Steady Rapidly Varied Flow 47

Fig. 1.26 Flow from a


reservoir into a steep channel H yc

Fig. 1.27 Flow from a


reservoir into a mild channel H yc
Eo
yo H

downstream. Therefore, the discharge cannot be changed by changing the


channel characteristics, such as the slope, size or roughness.
2. Under some conditions, however, the channel characteristics do determine the
discharge. If the channel is mild, i.e. the uniform flow is subcritical, then yo will
extend right back to the entrance, because it is controlled from downstream and
its control is the channel itself (Fig. 1.27). Critical flow at the entrance section is
then submerged, yo > yc and the discharge is reduced. The entrance is no longer
a control section because it is not followed by supercritical flow.
Because subcritical flow extends right up to the reservoir, it is the channel
characteristics which determine the discharge, and a relationship other than that
describing critical flow is required to calculate the discharge. This is provided by
Manning’s equation (or another similar resistance equation) combined with the
continuity equation, i.e.

A 2= 1=
Q¼ R 3S 2 ð1:52Þ
n

This equation has two unknowns (Q and yo), however, so some other rela-
tionship involving them is required to get a solution.

The discharge must also be influenced by the levels of the water and channel
bed at the entrance—this defines the amount of energy available, and the control
relationship (Eq. (1.52)) relates this to discharge. At the entrance, the specific
energy is equal to the height of the water surface above the channel bed
(Fig. 1.27). Because the subcritical flow extends right back to the reservoir, this
height must also be the uniform specific energy, so

Q2
H ¼ Eo ¼ yo þ ð1:53Þ
2 g A2o
48 1 Basic Hydraulic Concepts

Simultaneous solution of Eqs. (1.52) and (1.53) gives the discharge and the
uniform flow depth.
Note that the solution approaches for these two cases cannot be interchanged. It
would be wrong to use the critical flow equations if the channel has subcritical
uniform flow because critical flow does not occur. It would also be wrong to use
Manning’s equation and the specific energy relationship if the uniform flow in the
channel is supercritical because the flow is nonuniform in the channel for some
distance and so the specific energy at the entrance is not the same as for uniform
flow, i.e. H 6¼ Eo. It is usually not obvious if the uniform channel flow is subcritical
or supercritical, and an assumption must be made and subsequently checked. It is
usually easier to assume a supercritical channel first.
Example 1.7
Water flows from a lake into a 2.0 m wide rectangular channel with a Manning n of
0.013 and a slope of 0.0025. Calculate the discharge in the channel when the water level
in the lake is 1.8 m above the channel bed at the outfall. Assume zero entrance loss.
Solution
It is not obvious whether the channel is mild or steep. Assume it is steep and then check
and correct if necessary.
For a rectangular steep channel

Q ¼ qW

with
qffiffiffiffiffiffiffiffi
q¼ g y3c

and

2 2
yc ¼ H ¼ 1:80 ¼ 1:20 m
3 3

So
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
q¼ 9:8  1:203 ¼ 4:12 m3 =s/m

and

Q ¼ 4:12  2:00 ¼ 8:24 m3 =s

Check if yo with this discharge is subcritical or supercritical. yo can be determined from the
Manning equation together with continuity (Eq. (1.52)), i.e.
1.5 Steady Rapidly Varied Flow 49

A 2=3 1=2
Q¼ R S
n

A W yo 2:0 yo
R¼ ¼ ¼
P W þ 2 yo 2:0 þ 2 yo
 2=3
2:0 yo 2:0yo
i.e. 8:24 ¼ 0:00251=2
0:013 ð2:0 þ 2 yo Þ

From which, by trial, yo = 1.50 m.

yo > yc so the channel is mild. The discharge is therefore incorrect and must be recalculated
by simultaneous solution of Eqs. (1.52) and (1.53). This is done by trial, assuming a value for
yo, calculating Q from Eq. (1.52), using this to calculate H from Eq. (1.53) and then com-
paring this with the specified value. Only the final, correct, trial is shown here.
Try yo = 1.42 m. Then from Eq. (1.52)

A 2=3 1=2
Q¼ R S
n
 
2:0  1:42 2:0  1:42 2=3
¼ 0:00251=2 ¼ 7:71 m3 =s
0:013 2:0 þ 1:42

Inserting yo = 1.42 m and Q = 7.71 m3/s in Eq. (1.53) gives

Q2
H ¼ E o ¼ yo þ
2 g A2o
7:712
¼ 1:42 þ ¼ 1:80 m
2  9 : 8 ð2:0  1:42Þ2

which is the specified value. Therefore Q = 7.71 m3/s.

1.6 Steady Gradually Varied Flow

Flow in a channel will always tend to be uniform, and will depart from this
condition if it is disrupted by structural features or changes in slope, roughness or
cross-sectional geometry. Such disruptions can induce gradually varied nonuniform
flow conditions both upstream and downstream. If a disrupting feature acts as a
control, the induced conditions will be subcritical upstream and supercritical
downstream. Analysis and design of many hydraulic structures must include
determination of the associated gradually varied profiles.
Gradually varied profiles (unlike rapidly varied profiles) depend on flow resis-
tance, and result from the existence of a difference between the energy gradient (Sf)
and the slope of the bed (So). Upstream of a control, the flow depth is greater than
uniform, and therefore Sf < So and specific energy increases in the downstream
direction. Downstream of the control, the excess energy is dissipated by bed shear,
50 1 Basic Hydraulic Concepts

and through the gradually varied profile Sf > So. The bed shear is described by the
general resistance equation in terms of Sf,

s o ¼ c R Sf ð1:25Þ

For uniform flow Sf = Sw (the water surface slope) = So, and the value of So can
be used for Sf in the resistance equations (Manning, Chézy and Darcy–Weisbach).
This assumption cannot be made for nonuniform flow, and Sf must be used, which
varies along the profile. It is assumed, however, that the value of Sf at a section in
gradually varied flow is the same as it would be for uniform flow with the same
velocity and hydraulic radius. This enables Sf to be evaluated at sections in grad-
ually varied flow through application of the conventional resistance equations,
using the same resistance coefficients as used for uniform flow.

1.6.1 The Gradually Varied Flow Equation

The problem in gradually varied flow analysis is to describe the form of the water
surface profile, i.e. to quantify the variation of flow depth (y) with distance along the
channel (x). An equation for this can be derived by considering the change in
energy in the direction of flow. It is assumed that the slopes to be considered are
small enough that cos h  1.0 (the error would be less than 1% for a slope of 8°). It
is also assumed that a = 1.0, implying a uniform velocity distribution across the
section. (The error in the velocity head would be around 6% for turbulent flow, and
so the error in total energy would be less, considerably so for subcritical flow. If
more accurate calculations are required, a value of 1.06 for a may be used for
turbulent flow.) Equation (1.13) for the total energy at a section can then be written
as

V2
H ¼ z0 þ y þ ð1:54Þ
2g

The variation of the different components in the flow direction is obtained by


differentiating this equation with respect to the flow direction, x, resulting in

dE
¼ So  Sf ð1:55Þ
dx

in which So is the downstream change in bed level, dz′/dx, and Sf is the downstream
change in the total energy level through friction, dH/dx, both assumed positive
when decreasing. This indicates that specific energy will change along a channel if
the energy and bed gradients are different. If Sf = So, then dE/dx = 0 and flow is
uniform. Backing up behind a control causes Sf < So, so that dE/dx > 0 and specific
energy increases to provide the energy required by the control.
1.6 Steady Gradually Varied Flow 51

Rather than the change in energy along the channel, it is the change in flow
depth that is usually important. Equation (1.55) can be modified accordingly by
applying the chain rule of differentiation and noting that dE/dy can be shown to be
represented as

dE
¼ 1  Fr2 ð1:56Þ
dy

Equation (1.55) then becomes

dy So  Sf
¼ ð1:57Þ
dx 1  Fr2

Equation (1.57), to be referred to as the gradually varied flow equation, shows


that the rate of change of flow depth with distance depends on how far from uniform
the flow is (S0 – Sf) and how far from critical it is (1 – Fr2). The closer the flow is to
critical, the more rapid will be the change in flow depth; the closer to uniform it is
the more gradual the change. If Sf = So, then dy/dx = 0 and flow is uniform;
otherwise, it is nonuniform.

1.6.2 Classification of Gradually Varied Profiles

The computation of gradually varied profiles requires prior knowledge of the range
of water levels expected. The general forms of gradually varied water surface
profiles under different conditions can be described by qualitative application of
Eq. (1.57). The forms depend significantly on whether uniform flow in the channel
is subcritical or supercritical, and slopes are defined in these terms.
A mild slope is one on which uniform flow is subcritical. This can be defined
by the relationship between the uniform and critical flow depths, i.e.

yo [ yc
A steep slope is one on which uniform flow is supercritical, i.e.

yo \yc
A critical slope is one on which uniform flow is critical, i.e.

yo ¼ yc

Note that these definitions are hydraulic rather than geometric—it is generally
not possible to specify a value of channel gradient as steep or mild, and a particular
channel may be mild for some discharges and steep for others.
52 1 Basic Hydraulic Concepts

Whether the flow depth increases or decreases in the flow direction is determined
by the signs of the numerator and denominators of Eq. (1.57). The sign of the
numerator is determined by the value of the flow depth, y, in relation to the uniform
flow depth, yo. The Manning equation can be arranged in terms of Sf as

V 2 n2
Sf ¼ ð1:58Þ
R4=3

Clearly, if y = yo then Sf = So. It follows from Eq. (1.58) that if y < yo, then R is
less than the uniform value and V is greater than the uniform value, and hence
Sf > So and the numerator is negative. Similarly, if y > yo, then Sf < So and the
numerator is positive.
The sign of the denominator is determined by the value of the flow depth in
relation to the critical flow depth, yc. If y < yc, then the flow is supercritical, Fr > 1
and the denominator is negative. If y > yc, then the flow is subcritical, Fr < 1 and
the denominator is positive.
Consideration of these inequalities in the regions of flow depth defined by the
channel bed, the uniform flow depth and the critical flow depth allows the general
forms of the water surface profiles to be described and classified. The possible
profiles are shown in Figs. 1.28, 1.29 and 1.30 for mild, steep and horizontal slopes.
Notice in these profiles that when the water surface approaches the uniform flow
depth it does so asymptotically. When it approaches the bed and the critical flow
depth, it does so steeply.

1.6.3 Gradually Varied Flow Computation

Quantitative description of gradually varied profiles requires solution of the grad-


ually varied flow equation in one of its forms, i.e.

dH
¼ Sf ð1:59Þ
dx

Fig. 1.28 Gradually varied


flow profiles on a mild slope M1

M2 yo

M3 yc

Mild
1.6 Steady Gradually Varied Flow 53

Fig. 1.29 Gradually varied S1


flow profiles on a steep slope

yc
S2

S3 yo

Steep

Fig. 1.30 Gradually varied


flow profiles on a horizontal H1
slope

yc
H2

Horizontal

dE
¼ So  Sf ð1:55Þ
dx

or

dy So  Sf
¼ ð1:57Þ
dx 1  Fr2

Three approaches to solution of these equations have been used, viz. direct
integration, graphical integration and numerical integration. Different circumstances
may dictate the use of different methods. Graphical integration is rarely used
because of the ready availability of efficient numerical solution software. Direct
integration techniques have been developed which are accurate, easy and quick to
apply, but are limited to simple situations with prismatic channels with regular
geometric cross sections and require tables of integration values. Numerical inte-
gration is now the most widely used approach and several different methods have
been developed. Some methods are simpler to apply than others, but these are
54 1 Basic Hydraulic Concepts

generally more restrictive in terms of applicable conditions, e.g. the simplest


methods can only be used for prismatic channels.
Whatever method is used, the computation must begin at a control section—or at
least a position where the depth is known—and proceed in the direction in which
the control is exercised. The computation should proceed in the downstream
direction for supercritical flow and in the upstream direction for subcritical flow.
This requires that before the computations can be done, the control feature must be
analysed to determine the control depth on the profile, and the type of profile must
be identified so that the range of possible flow depths is known.
Computation of Distance from Flow Depth
The Direct Step Method is a simple method for computing profiles in prismatic
channels, using Eq. (1.55) (or alternatively Eq. (1.57)) in finite difference form.
A series of flow depths known to occur along the profile are specified (which is why
the profile type must be known), and the distances between them are calculated,
starting at the control depth (determined by the control analysis).
For example, for an M1 profile upstream of a sluice gate (Fig. 1.31), the sluice gate
is first analysed to determine the control depth at the downstream end of the profile.
Then a number of flow depths (yi) between this depth and the uniform flow depth (yo)
are nominated. The distance from the control depth to the next smaller depth (Dx1) is
then calculated, and then the distance from this to the next smaller (Dx2), and so on
until a sufficient length of the profile has been defined. An M1 profile approaches
uniform flow asymptotically, so its extent is infinite, and a flow depth satisfactorily
close to uniform must be chosen as an approximation for calculation purposes.
The distances (Dxi) between adjacent nominated depths are calculated from
Eq. (1.55) in finite difference form and expressed in terms of Dx, i.e.

DE
Dx ¼ ð1:60Þ
So  Sf

DE is the difference between the specific energy values at the adjacent sections
under consideration. This can be calculated directly from the nominated flow

Q M1
y7 control depth
y6
y5
y4
y3
y2 yo
y1
Δx7
Δx6
Δx5
Δx4
Δx3
Δx2
Δx1

Fig. 1.31 Step computation procedure


1.6 Steady Gradually Varied Flow 55

depths, the discharge and the channel cross-sectional geometry. Sf is the energy
gradient between the sections, and is approximated as the average of the values at
the sections. The values of Sf at a section can be calculated from the flow depth
using one of the resistance equations, for example, using Manning’s equation,

V 2 n2
Sf ¼ ð1:58Þ
R4=3

Computation is direct, and no iteration is required. This approach is generally


satisfactory for regular, prismatic channels where an extended water surface profile
must be defined. It is less satisfactory in cases where the flow depth at a particular
location must be determined, for example, at the bottom of a chute or drop structure
so that the Froude number can be determined for stilling basin design, or at the inlet
into a channel from a lake for calculating the discharge if the flow is subcritical and
nonuniform. In these cases, the calculations would have to extend past the location
required and the flow depth found by interpolation, or the last calculation step could
be repeated iteratively until the required distance results. Such problems can still be
solved by the Direct Step Method, but the procedure is cumbersome. The following
method allows the nomination of cross-section locations at which the flow depths
are calculated, i.e. x is nominated and y is calculated.

Example 1.8
The channel in Example 1.5 has a slope of 0.0010 and a Manning’s n of 0.013.
Determine the distance from the hump to the location of a hydraulic jump.
Solution
As a control, the hump induces supercritical flow on its downstream side. The channel is
mild, so the flow assumes an M3 profile which ends in a hydraulic jump to the sub-
critical uniform flow controlled from downstream.

M3 yo = 1.07 m

0.75 m 0.34 m 0.42 m


1 2

The distance required, x, is from y1 as controlled by the hump to y2 which is the


conjugate of the uniform flow depth. This is given by

q2 y2
M2 ¼ Mo ¼ þ o
g yo 2
1:802 1:072
¼ þ ¼ 0:88 m2
9:8  1:07 2
56 1 Basic Hydraulic Concepts

Therefore

1:802 y2
0:88 ¼ þ 2
9:8 y2 2

From which, by trial, y2 = 0.42 m.


The problem is to calculate a distance from flow depths, so the Direct Step Method is
applicable. Calculations are shown below, using equal depth increments of 0.010 m. In
the table

y flow depth,
A flow area = W  y,
P wetted perimeter = W + 2  y,
R hydraulic radius = A/P,
V velocity = Q/A,
E specific energy = y + V 2/2g,
Sf energy gradient = V 2 n2/R4/3,
Sf average energy gradient over Dx.

y A (m2) P (m) R (m) V V2/ E (m) Sf Sfm So − Sf ΔE Δx (m) x (m)


(m) (m/s) 2 g (m)
0.34 0.850 3.180 0.267 5.294 1.430 1.770 0.02751 0
0.026348 −0.02535 −0.071 2.78
0.35 0.875 3.200 0.273 5.143 1.349 1.699 0.025186 2.78
0.024153 −0.02315 −0.064 2.76
0.36 0.900 3.220 0.280 5.000 1.276 1.636 0.023119 5.54
0.022198 −0.0212 −0.058 2.74
0.37 0.925 3.240 0.285 4.865 1.207 1.577 0.021276 8.28
0.020452 −0.01945 −0.053 2.71
0.38 0.950 3.260 0.291 4.737 1.145 1.525 0.019627 10.99
0.018887 −0.01789 −0.048 2.68
0.39 0.975 3.280 0.297 4.615 1.087 1.477 0.018147 13.67
0.01748 −0.01648 −0.044 2.65
0.40 1.000 3.300 0.303 4.500 1.033 1.433 0.016814 16.32
0.016212 −0.01521 −0.040 2.62
0.41 1.025 3.320 0.309 4.390 0.983 1.393 0.01561 18.94
0.015066 −0.01407 −0.036 2.58
0.42 1.050 3.340 0.314 4.286 0.937 1.357 0.014521 21.52

The distance to the hydraulic jump is therefore approximately 22 m

Computation of Flow Depth from Distance


In situations where it is required to calculate flow depths at specified locations,
Eq. (1.57) should be used in finite difference form, i.e.
 
So  Sf
Dy ¼ Dx ð1:61Þ
1  Fr2
1.6 Steady Gradually Varied Flow 57

As before, calculations begin at a control section where the flow depth is known.
A series of distance increments (Dx) along the channel are specified and changes of
flow depth through them (Dy) are calculated sequentially (Fig. 1.31) using
Eq. (1.61). Because the average values of Sf and Fr through Dx depend on the
unknown depth at the end section, conventional practice is to assume a value for
this depth and recalculate it iteratively. For regular channels, such as in most
hydraulic structure applications, it is acceptable to use the characteristics of the
known section to represent the entire reach up to the unknown section, especially if
Dx is kept small. (This just assumes the values of Sf and Fr to represent the reaches
between the midpoints of adjacent sections rather than the reaches between the
sections themselves.) Iteration is then unnecessary and the change in flow depth can
be calculated directly from Eq. (1.61).
For irregular channels, and especially natural rivers, iterative solution is necessary.
The Standard Step Method, which applies Eq. (1.61) iteratively and provides for
checking assumed flow depths, is described in basic open channel hydraulics texts.
For analysis of some hydraulic structures, it is necessary to extend this theory to
allow for a varying discharge along the profile, such as for side weirs where water
leaves the channel or side-channel spillways where water enters the channel. The
necessary adjustments and extensions are made where these structures are presented
in Chap. 4.

Problems
1:1 Water flows at a uniform depth of 0.50 m in a 1.0 m wide rectangular
channel constructed on a slope of 0.0010. Determine the discharge
a. if the channel is lined with concrete with ks = 0.50 mm and
b. if the channel is made of steel with ks = 0.080 mm.
1:2 A rectangular channel is 2.50 m wide and has a slope of 0.0020 and
n = 0.015.
a. What is the discharge when the flow depth is 1.20 m?
b. What is the flow depth when the discharge is 2.50 m3/s?
c. What are the values of C and f for this channel?
d. Suggest a representative value of ks for this channel.
1:3 A rectangular concrete-lined channel is 3.0 m wide and conveys a discharge
of 7.5 m3/s. What are the alternate depths if the specific energy is 2.4 m?
1:4 The floor elevation of a 3.6 m wide rectangular channel rises 0.22 m. If the
mean velocity upstream of the rise is 1.2 m/s and the flow depth is 1.2 m,
what is the change in water surface elevation through the transition?
1:5 Water flows at 2.0 m3/s in a 1.0 m wide rectangular channel. There is a
smooth downward step of 0.20 m in the channel and the flow depth
downstream of the step is 1.2 m. What is the flow depth upstream of the
step?
58 1 Basic Hydraulic Concepts

1:6 Find the critical flow depth, critical specific energy and critical slope for the
following concrete-lined channels when the discharge is 1.0 m3/s.
a. Rectangular, 1.0 m wide.
b. Trapezoidal with base width = 0.50 m and side slopes of 45°.
c. Triangular with side slopes of 2H:1V.
1:7 Show that the general definitions or expressions for Fr, yc, Vc and E reduce to
the simpler, particular expressions applicable for rectangular channels.
1:8 Water flows at a discharge of 9.0 m3/s and a depth of 1.5 m in a long,
rectangular, concrete-lined channel with a width of 3.0 m. A short, smooth
hump is installed in the channel. Determine the flow depths on top of the
hump and immediately upstream and downstream if the height of the hump is
(a) 0.20 m and (b) 0.30 m.
1:9 A long channel has a rectangular section 3.0 m wide and incorporates a
short, smooth contraction to a width of 2.60 m, followed by an expansion to
the original width. If the uniform flow has a velocity of 3.0 m/s and a depth
of 3.0 m, determine
a. the flow depth in the contraction,
b. the minimum width in the contraction for the upstream flow to be uniform
and
c. the flow depths in the contraction and immediately upstream and
downstream if the width in the contraction is 2.20 m.
1:10 Water is flowing at 2.0 m3/s in a 1.0 m wide rectangular channel. There is an
abrupt downward step of 0.25 m in the channel and the flow depth down-
stream of the step is 1.20 m.
a. Using the specific energy and momentum function principles, calculate
the energy loss caused by the step. Sketch the energy line and water
surface profile, indicating values upstream and downstream of the step.
b. What minimum step height would cause the upstream flow depth to be
critical?
1:11 A long, wide, rectangular channel has a Manning’s n of 0.014 and incor-
porates a short, smooth, upward step of 0.20 m. If the discharge in the
channel is 2.0 m3/s/m wide, determine the flow depths immediately upstream
and downstream of the step
a. if the channel slope is 0.00010,
b. if the channel slope is 0.0010 and
c. if the channel slope is 0.010.
1:12 A long, concrete-lined (n = 0.013) channel on a slope of 0.0010 has a
rectangular cross section with a width of 4.00 m. The channel incorporates a
smooth contraction to a width of 2.80 m followed immediately by a smooth
expansion back to the original width. A short distance downstream is a
1.6 Steady Gradually Varied Flow 59

similar contraction to a width of 2.60 m. Determine the flow depths before,


after and within each contraction for a discharge of 20.0 m3/s.
1:13 A long, concrete-lined (n = 0.013) channel with a rectangular section 4.0 m
wide and a slope of 0.0045 incorporates the features shown below. Both
humps are 0.10 m high and the contraction reduces the channel width locally
to 3.0 m. All the features are ‘smooth’ and do not induce local energy loss,
and are close enough to each other that gradually varied flow changes
between them are insignificant. If the discharge in the channel is 40 m3/s,
what are the flow depths immediately before and after each feature?

flow Longitudinal
section

hump contraction hump

flow
Plan

1:14 Water flows from a lake into a long channel with a Manning’s n of 0.014.
The water level in the lake is 2.7 m above the channel bed at the outfall.
a. Calculate the discharge in the channel if it has a 3.0 m wide rectangular
cross section and a slope of (i) 0.0070 and (ii) 0.0020.
b. Calculate the discharge if the slope is 0.0070 and the channel is trape-
zoidal in section with a bottom width of 4.5 m and side slopes of 2H:1V.
Assume zero entrance loss.

Reference

Chow, V. T. (1959). Open-channel hydraulics. McGraw-Hill.

Further Reading

Chadwick, A., Morfett, J., & Borthwick, M. (2013). Hydraulics in civil and environmental
engineering (5th ed.). CRC Press.
Chanson, H. (1999). The hydraulics of open channel flow: An introduction. Elsevier.
Chaudhry, M. H. (1993). Open channel flow. Prentice-Hall.
French, R. H. (1985). Open-channel hydraulics. McGraw-Hill.
Henderson, F. M. (1966). Open channel flow. Macmillan.
Chapter 2
Underflow Gates

2.1 Introduction

Gates are frequently used to control discharge in canals or on spillway crests. The
vertical sluice gate is particularly common, but there are also a variety of other
types as well, such as the radial (or Tainter) gate and the drum gate (Fig. 2.1).
The different types have relative advantages in different situations. The vertical
gate may require costly roller and track assemblies, while the radial gate may
require more structural expense. Radial gates are common on dam crests and outlets
(Fig. 2.2) and have the advantage of being able to be lifted clear of the water
surface to allow debris to pass through. They are not often used in small canals.
(The use of gates on spillway crests is discussed further in Chap. 4.)
As for other flow regulation structures, gates can operate as true controls, in
which case the outflow is unsubmerged and flow downstream is supercritical. If the
flow downstream is subcritical and sufficiently deep, the outflow may be submerged
and the structure will not be a true control. In this case, it is still necessary to be able
to analyse the flow to determine discharge (such as from a reservoir into a mild
channel) or the change in flow depth from downstream to upstream.

2.2 Unsubmerged Analysis

The unsubmerged analysis is straightforward, and based on energy conservation.


The gate is a control; the flow upstream is subcritical and the flow downstream is
supercritical (Fig. 2.3). Energy losses need not be considered because there is no
flow expansion and friction is negligible over short distances, so

H1 ¼ H2 ð2:1Þ

© Springer Nature Switzerland AG 2020 61


C S James, Hydraulic Structures,
https://doi.org/10.1007/978-3-030-34086-5_2
62 2 Underflow Gates

(a) Vertical (b) Radial (c) Drum

Fig. 2.1 Types of underflow gate

Fig. 2.2 Radial gates at a reservoir outlet

and, if the bed is horizontal between the sections

E1 ¼ E2 ð2:2Þ

i.e.

q2 q2
y1 þ ¼ y 2 þ ð2:3Þ
2gy21 2gy22
2.2 Unsubmerged Analysis 63

y1
vena contracta

a
y2 q

1 2

Fig. 2.3 Unsubmerged flow through vertical sluice gate

Under free flow conditions, the flow depth at the vena contracta is defined by the
geometry of the opening, and can be estimated as

y2 ¼ Cc a ð2:4Þ

For a vertical gate, the value of the contraction coefficient Cc is typically 0.61 for
a/y1 less than about 0.7. The value changes with the inclination of the lip of the
gate. It is therefore greater than 0.61 for a radial gate, and varies with the gate
opening (Fig. 2.4); for a drum gate, it is close to 1.0. Equations are available for
relating Cc to the angle h, such as that proposed by Henderson (1966) for pre-
liminary estimates, i.e.

Cc ¼ 1  0:75 h þ 0:36 h2 ð2:5Þ

Fig. 2.4 Contraction at an


inclined gate

θ
64 2 Underflow Gates

vena contracta
yo
yo conjugate

Fig. 2.5 Flow profile for unsubmerged flow through vertical sluice gate

in which h is in units of 90°. This equation gives values of Cc to within ±5% for
h  1.
If required, the specific energy relationship can be developed to give an equation
for Q as a function of the upstream water level. However, most problems can be
solved from first principles.
For the unsubmerged case, on a long mild slope, the downstream flow follows a
supercritical gradually varied profile before a hydraulic jump to the subcritical
uniform flow (Fig. 2.5). The position of the jump is where the flow depth on the
supercritical profile is the conjugate of the uniform flow depth, i.e. where the
M values of the supercritical and subcritical flows are equal. With deeper uniform
flow (greater M) the hydraulic jump will be nearer to the gate, and if it is deep
enough (Mo > Mvena contracta), subcritical flow will submerge the exiting flow
(Fig. 2.6). The gate will then not be a control. (Note that the downstream subcritical
flow might be determined by another control feature rather than being uniform.)

2.3 Submerged Analysis

If the gate opening is submerged, the upstream water level cannot be determined by
analysing the gate alone, and depends on the flow conditions downstream as well as
the gate opening. The above analysis no longer applies, and requires some exten-
sion. Flow will still issue as a jet under the gate, but there will be a turbulent mass
of water above it. This mass has no net motion in the downstream direction and
therefore does not contribute to the discharge but, by virtue of its weight, it has a
strong influence on the pressure in the jet.
The common analysis problem is to determine the upstream flow depth, y1, for a
known discharge, q. The flow is subcritical throughout, and control is therefore
from downstream. The analysis must therefore begin by establishing the down-
stream flow depth, y3, and working upstream. The depth y3 will be the uniform flow
depth if the channel is regular over a long distance, or it will depend on some other
downstream control which would have to be analysed first. It is not possible to
relate the flow conditions between sections 1 and 3 directly. Energy conservation
cannot be applied because the flow expansion and associated turbulent mixing
2.3 Submerged Analysis 65

y1

y y3

y2

1 2 3

Fig. 2.6 Submerged flow through vertical sluice gate

between sections 2 and 3 cause significant, but unquantifiable energy loss.


Momentum conservation can also not be applied because the force imposed on the
flow by the sluice gate cannot be easily quantified (it is not hydrostatic because the
water levels on either side are different). The problem can be solved by including
section 2 and relating the conditions here to those at sections 1 and 3. Section 2 is
located at the vena contracta of the jet issuing under the sluice gate. The major
energy loss occurs in the expansion after the vena contracta so energy conservation
can be applied between sections 1 and 2, with energy loss considered negligible.
No forces are applied to the flow between sections 2 and 3 (apart from the boundary
shear force, which is negligible over short distances) and so momentum conser-
vation can be applied between these two sections. This approach of using energy
and momentum conservation together is also useful in analysing submerged flow
conditions for many other structures, such as transitions and culverts.
Beginning with a known flow condition at section 3, the conditions at section 2
are determined through momentum conservation, i.e.

M2 ¼ M3 ð2:6Þ

The expression for M3 is straightforward, but the flow at section 2 is concentrated


within the jet (Fig. 2.7) and the formulation must be modified to account for this.
The first term of the momentum function (q2/gy) represents the dynamic con-
tribution and so the correct depth to use at section 2 is that associated with the
velocity, through continuity, i.e. y2 = Cca. The second (y2/2) term represents the
hydrostatic pressure force contribution, and so the appropriate flow depth is that
associated with the hydrostatic pressure, i.e. y as shown in Fig. 2.7. Knowing the
gate opening, a, therefore enables the actual flow depth at section 2 (y) to be
determined by evaluating Eq. (2.6) as
66 2 Underflow Gates

y
p/γ
v

Fig. 2.7 Pressure and velocity distributions downstream of submerged gate

q2 y2 q2 y2
þ ¼ þ 3 ð2:7Þ
gð C c aÞ 2 gy3 2

Once the conditions at section 2 are known, these can be related to those at
section 1 through energy conservation, i.e.

E1 ¼ E2 ð2:8Þ

E1 can be expressed in terms of y1 and q in the usual way but, again, the jet-flow
conditions at section 2 require adjustment of the usual specific energy formulation,
taking cognizance of the meanings of the individual terms. The first term in the
specific energy function (y) represents the piezometric contribution to energy, i.e.
the sum of the elevation and pressure energy terms. Because the pressure at
section 2 is determined by the distance below the water surface, the full flow depth
(y as shown in Figs. 2.6 and 2.7) must be used for this term. The second term in the
specific energy function (q2/2gy2) is the velocity head, and the depth here represents
the flow area (through continuity). It must therefore be the depth associated with the
jet, i.e. Cc a. The equation to solve for y1 is therefore

q2 q2
y1 þ 2
¼ yþ ð2:9Þ
2gy1 2gðCc aÞ2

If q is known and y1 is required, the solution described above is straightforward.


If q is to be determined from known values of y1 and y3 (a typical discharge
measurement problem) then an iterative solution is required because the momentum
function and specific energy equations both contain two unknowns (q and y). The
two equations must be set up and solved simultaneously.
2.3 Submerged Analysis 67

The simplifications made in this analysis are not entirely realistic. There is an
energy loss between sections 1 and 2; the velocity distribution in the jet at the vena
contracta is not uniform; the roller above the vena contracta does contribute
momentum flux to the jet; the kinetic energy correction factor (a) and momentum
correction factor (b) should not be assumed to equal 1.0 at all sections; the contraction
coefficient value is not always equal to 0.61. Castro-Orgaz et al. (2013) proposed
refinements to the analysis to account for the various assumptions but observed that,
owing to compensating inaccuracies, the standard approach (as outlined above)
would provide results with similar accuracy. The approach becomes questionable,
however, if a > 1.64yc when the gate acts as a subcritical flow transition.
It may not be known beforehand whether the gate is submerged or not. The best
way to find out is first to assume that it is not submerged, and then to check if M2
> M3. If it is, then there will be a hydraulic jump some distance downstream and the
unsubmerged analysis will be correct. If M3 > M2, then the outflow will be sub-
merged and the analysis must be redone as described above.
Example 2.1
Water is released from a large reservoir through a vertical sluice gate into a long,
rectangular, 4.0 m wide, concrete (n = 0.013) channel on a slope of 0.00050. Plot the
relationship between the water level in the reservoir and the discharge when the gate is at
its maximum opening of 0.80 m.
Solution

submerged
H

a
unsubmerged

1 2 So = 0.00050

Calculation of depth from discharge is more direct than calculation of discharge from
depth, so a range of discharge values is nominated and the water level in the reservoir
(H) calculated for each. For each discharge, it must be determined whether the gate is
unsubmerged or submerged, and the appropriate analysis is applied.
For Q = 2.0 m3/s:
y1 = Cc  a = 0.61  0.80 = 0.488 m,
y2 = yo, from Manning equation

A 2=3 1=2
Q ¼ 2:0 ¼ R So
n
 2=3
4:0yo 4:0yo
¼ 0:000501=2
0:013 ð4:0 þ 2yo Þ

From which, by trial, y2 = yo = 0.523 m.


68 2 Underflow Gates

Check submergence:

q2 y2
M1 ¼ þ 1
gy1 2

Q 2:0
q¼ ¼ ¼ 0:50 m3 =s=m
W 4:0

0:502 0:4882
¼ þ ¼ 0:224 m2
9:8  0:488 2

And

0:502 0:5232
M2 ¼ þ ¼ 0:185 m2
9:8  0:523 2
M1 > M2, so gate is unsubmerged.
Assuming negligible velocity in the reservoir and no energy loss through the gate,

q2
H ¼ E1 ¼ y1 þ
2gy21
0:502
¼ 0:488 þ ¼ 0:542 m
2  9:8  0:4882
For Q = 3.5 m3/s:
y1 = Cc  a = 0.61  0.80 = 0.488 m,
y2 = yo, from Manning equation

A 2=3 1=2
Q ¼ 3:5 ¼ R So
n
 2=3
4:0yo 4:0yo
¼ 0:000501=2
0:013 ð4:0 þ 2yo Þ

From which, by trial, y2 = yo = 0.758 m.


Check submergence:

q2 y2
M1 ¼ þ 1
gy1 2

Q 3:5
q¼ ¼ ¼ 0:875 m3 =s=m
W 4:0

0:8752 0:4882
¼ þ ¼ 0:302 m2
9:8  0:488 2

And
0:8752 0:7582
M2 ¼ þ ¼ 0:390 m2
9:8  0:758 2
M1 < M2, so gate is submerged
Calculate y1 by momentum conservation:
2.3 Submerged Analysis 69

q2 y2
M1 ¼ þ 1 ¼ M2
gðCc aÞ 2

Therefore
  1=2
q2
y1 ¼ 2 M2 
gCc a
i.e.
  1=2
0:3752
y1 ¼ 2 0:390  ¼ 0:679 m
9:8  0:61  0:8
Assuming negligible velocity in the reservoir and no energy loss through the gate,

q2
H ¼ E1 ¼ y1 þ
2gðCc aÞ2
0:3752
¼ 0:679 þ ¼ 0:843 m
2  9:8  0:4882
Similar calculations for other discharges produce the rating curve below. Note the
distinct change in the rate of increase of H with Q once the gate is submerged.

1.2

1.0

0.8
H (m)

0.6

0.4

0.2

0.0
0 1 2 3 4 5 6
Q (m3/s)

2.4 Hysteretic Behaviour

Whether a sluice gate in a steep channel acts as a control or simply allows


undisturbed supercritical flow to pass through it depends on both the gate opening
and the preceding flow conditions. Figure 2.8 shows the possible water surface
profiles, assuming a long channel and hence uniform undisturbed flow. Figure 2.8a
shows a gate opening greater than the undisturbed water depth. If the gate opening
is reduced until it just touches the surface, the gate will immediately become a
control—the water will back up as subcritical flow on the upstream side and a
hydraulic jump will occur a long distance upstream before an S1 profile (Fig. 2.8b).
If the gate is closed further, the upstream water level will increase and the hydraulic
jump will advance upstream. If the gate is then opened progressively, the upstream
level will decrease and the hydraulic jump will move downstream. The gate will
remain a control, with the associated S1 profile and hydraulic jump, with openings
greater than the original undisturbed water depth until the hydraulic jump has
70 2 Underflow Gates

Fig. 2.8 Hysteretic


behaviour of a vertical sluice
gate

yo (< yc )

(a) Undisturbed flow

S1

yo S3

(b) Gate-controlled flow

yo conj.

yo

(c) Limit of gate-controlled flow

yo

(d) Undisturbed flow


2.4 Hysteretic Behaviour 71

reached the gate (Fig. 2.8c). Any further opening will restore the undisturbed
uniform flow (Fig. 2.8d). There is therefore a range of gate openings greater than
the undisturbed depth for which the flow depth upstream is controlled either by the
gate, for which the unsubmerged analysis above applies, or by the channel char-
acteristics through a resistance equation.
This behaviour is described in more detail by Defina and Susin (2003). They also
observe that similar hysteretic behaviour can occur in a mild channel, provided that
the Froude number of the approach flow exceeds about 0.8.
Example 2.2
A vertical sluice gate is located in a long, rectangular, concrete channel with a width of
3.0 m and a slope of 0.015. Plot the relationship between the flow depth immediately
upstream of the gate and the gate opening if the gate opening reduces from 1.70 m to
0.60 m and then increases back to 1.70 m when the discharge is 15.0 m3/s.
Solution

yo a

So = 0.0150
1 2
n = 0.013

The flow depth will not be affected if the gate opening is initially greater than the
uniform flow depth. Calculate this using the Manning and continuity equations:

A 2=3 1=2
Q ¼ 15:0 ¼ R So
n
 2=3
3:0yo 3:0yo
¼ 0:0151=2
0:013 3:0 þ 2yo

From which, by trial, yo = 0.813 m.


Therefore, as a reduces, y1 will be constant at yo = 0.813 m until a = 0.813 m (dashed
profile above).
As soon as the gate touches the water surface, backing up will occur (solid profile
above). Then

q2 q2
E1 ¼ y1 þ ¼ E2 ¼ y2 þ
2gy21 2gy22

Q 15:0
q¼ ¼ ¼ 5:0 m3 =s=m
W 3:0
72 2 Underflow Gates

y2 ¼ Cc a ¼ 0:61  0:813 ¼ 0:496 m

Therefore

5:02 5:02
y1 þ ¼ 0:496 þ ¼ 5:68 m
2  9:8y1
2 2  9:8  0:4962

From which, by trial, y1 = 5.64 m.


Similar calculations give y1 values for a reducing to 0.60 m, and then increasing until
y1 is the conjugate of yo. This is found from

q2 y2o conj q2 y2
Mo conj ¼ þ ¼ Mo ¼ þ o
g yo conj 2 g yo 2

Therefore

5:02 y2o conj 5:02 0:8132


þ ¼ þ ¼ 3:47 m2
9:8yo conj 2 9:8  0:813 2

From which, by trial, yo conj = 2.13 m.


From energy calculations for increasing a values, y1 is found to equal 2.13 m for
a = 1.52 m. Any further increase in a would result in undisturbed flow through the
gate once more, i.e. y1 = yo. Therefore, y1 = yo = 0.813 m for a increasing from
1.52 m to 1.70 m, as for the initial condition.
The relationship between y1 and a is therefore as shown below:

12

a decreasing
10
a increasing

8
y1 (m)

0
0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8
a (m)
2.4 Hysteretic Behaviour 73

Problems
2:1 A concrete-lined channel has a width of 5.0 m, a slope of 0.0060 and
Manning’s n of 0.013. A vertical sluice gate in the channel is set at an opening
of 0.80 m. If the discharge is 20.0 m3/s, locate the resulting hydraulic jump
relative to the sluice gate using a numerical step procedure with three
approximately equal steps.
2:2 A vertical sluice gate is installed in a long, 2.0 m wide, concrete-lined, rect-
angular channel on a slope of 0.00050.
a. If the water depth immediately upstream of the gate is 3.0 m and the gate
opening is 0.50 m, what is the discharge in the channel?
b. For the flow conditions defined in (a), determine the average boundary
shear stress immediately downstream of the sluice gate.
c. What is the water depth immediately upstream of the gate if the gate
opening is increased to 1.00 m?

References

Castro-Orgaz, O., Mateos, L., & Dey, S. (2013). Revisiting the energy-momentum method for
rating vertical sluice gates under submerged flow conditions. Journal of Irrigation and
Drainage Engineering, ASCE, 139(4), 325–335.
Defina, A., & Susin, F. M. (2003). Hysteretic behavior of the flow under a vertical sluice gate.
Physics of Fluids, 15(9), 2541–2548.
Henderson, F. M. (1966). Open Channel Flow, Macmillan.
Chapter 3
Open Channel Transitions

3.1 Introduction

A transition is a structure associated with a change in flow conditions. The most


common transitions in practice provide either expansions or contractions in con-
veyance structures, often incorporating changes in cross-sectional shape. The fol-
lowing cases are common:
• A change in the size and/or shape of an artificial canal in association with a
change in gradient or required capacity.
• Connection of a structure such as a diversion, reservoir outlet or a siphon to a canal
(Figs. 3.1 and 3.2). This could be a canal–canal or a pipe–canal connection.
• A change from a lined to an unlined canal.
• Entry to a chute or drop structure in a canal or from a spillway.
• Entry to and exit from a flume through a tunnel or over a crossing.
Transitions may be controls, in which case they can affect the flow for great
distances both upstream and downstream. If they do not act as controls they may
still affect the upstream water level under subcritical flow conditions to accom-
modate the energy changes and losses they induce; these can influence conditions
for a long distance upstream, but obviously have no effect downstream. Such
extensive influences require analysis by gradually varied flow methods and will not
be considered here, although the analysis would obviously have to include estab-
lishment of control depths which are determined by the local effects of the structure.
The design of a transition structure requires the specification of its geometry to
ensure satisfactory performance in terms of flow characteristics. Depending on
particular requirements, it may be necessary to predict the variation of flow depth
through the structure, as this will determine the height required for side walls and
define initial depths for analysing gradually varied flow upstream and downstream.
If a transition is not a control it may still be necessary to evaluate the energy loss it
induces; it is commonly required to produce a design that minimizes energy loss. It

© Springer Nature Switzerland AG 2020 75


C S James, Hydraulic Structures,
https://doi.org/10.1007/978-3-030-34086-5_3
76 3 Open Channel Transitions

Fig. 3.1 A transition from a reservoir outlet to a canal

Fig. 3.2 A transition from a tunnel exit to a canal


3.1 Introduction 77

may also be necessary to predict the distribution of flow velocity through the
structure to check for distortions and possible separation zones and, especially for
supercritical flow, the occurrence and pattern of induced surface waves. The design
is usually done by trial, i.e. a geometry is configured and then analysed to establish
its performance, and modified if necessary.
All transitions considered here are relatively short and the flow through them
will be rapidly varied, associated with changes in direction, slope or cross-section
size and/or shape. Simple expansions and contractions have already been discussed
in the development of specific energy and momentum function concepts in Chap. 1.
The situation in practical structures requires application of the same principles, but
with some additional considerations. The effect of a transition structure on flow
characteristics depends greatly on whether the flow is subcritical or supercritical,
with the latter usually presenting greater design difficulties. These situations are
treated separately.

3.2 Subcritical Flow Transitions

If flow is subcritical through a transition the analysis is fairly straightforward, using


energy and momentum principles. Problems which cannot be adequately addressed
theoretically are the estimation of energy loss for some geometries and undesirable
flow conditions, such as waves and distorted velocity profiles, which require some
empirical input.
The main purpose of the analysis is to determine the flow depths upstream and
downstream for a specified discharge, Q. The approach is similar for expansions
and contractions. For illustration purposes, an abrupt flat bed expansion is shown in
Fig. 3.3. Variations of this geometry include straight or curved gradually tapered
transitions, and warped transitions where changes in cross-sectional shape occur.
Because flow is subcritical, y3 must be known from an analysis of its control
further downstream; it is commonly the uniform flow depth. The upstream flow
depth, y1, can then be related to this using energy conservation, i.e.

E1 ¼ E3 þ loss ð3:1Þ

Chow (1959) expanded this relationship to give the change in water surface as
 
V32 V12
y1  y3 ¼ ð 1 þ ci Þ  ð3:2Þ
2g 2g

for a contraction, where the water level drops through the transition, and
 
V12 V32
y3  y1 ¼ ð 1  co Þ  ð3:3Þ
2g 2g
78 3 Open Channel Transitions

Fig. 3.3 An abrupt channel


expansion, shown in plan

b1 b3

1 2 3

Table 3.1 Energy loss Transition type ci co


coefficients for flat-bedded
contractions and expansions Warped 0.10 0.20
(Chow 1959) Cylinder quadrant 0.15 0.25
Simplified straight line 0.20 0.30
Straight line 0.30 0.50
Square-ended 0.30+ 0.75

for an expansion, where the water level rises through the transition. For most
geometries, the loss coefficients ci and co must be determined empirically. Chow
(1959) provides values for different transition geometries, as listed in Table 3.1.
Note that energy losses through expansions are considerably greater than through
contractions. Formica (1955) expressed losses in terms of the downstream velocity
head and found values of up to 0.23 V23/2 g for square-edged contractions in
rectangular channels and 0.11 V23/2 g for rounded contractions; he also found the
coefficient values to vary with the ratio y3/b2. Yarnell’s (1934) results for bridge
piers would also give indications of suitable coefficient values.
The upstream and downstream depths can also be related using momentum
conservation in terms of the momentum function, i.e.

P
M1 ¼ M3  ð3:4Þ
c

in which P is the total force exerted in the flow direction by the transition surfaces,
or using the force–momentum flux equation, i.e.
X
F ¼ q QðV3  V1 Þ ð3:5Þ
3.2 Subcritical Flow Transitions 79
P
in which F includes the forces from the transition surfaces as well as the
hydrostatic forces at sections 1 and 3. Application of momentum conservation
avoids having to estimate energy losses, but requires estimating the forces acting on
the free body between the upstream and downstream sections. This presents the
difficulty that the forces from the transition surfaces depend on the flow depth at
section 2 (equal to the depth at section 1), which is the unknown that is sought, and
the solutions to Eqs. (3.4) and (3.5) therefore require iteration. The calculation is
nevertheless fairly straightforward for abrupt expansions where the transition forces
depend only on the depth at the upstream section. It is more complicated if the
transition is not abrupt, where the hydrostatic pressure varies through the transition
under an unknown water surface profile. Najafi-Nejad-Nasser and Li (2015) pre-
sented an analytical solution for a tapered expansion incorporating a hump, but this
depends on an empirical description of the water surface variation.
For an abrupt flat bed expansion, as shown in Fig. 3.3, Henderson (1966)
derived an expression for the energy loss between sections 1 and 3 by inserting
section 2 and using M2 = M3 and E1 = E2. This yields the expression
 
V12 2 1 2s
DE ¼ E1  E3 ¼ 1þ 2  2 2  2 ð3:6Þ
2g Fr1 r s Fr1

in which r = b2/b1 and s = y3/y1. Henderson (1966) showed that if Fr1 is small
enough for Fr41 to be considered negligible, then this can be expressed as
  !
V2 1 2 Fr21 ðr  1Þ2
DE ¼ E1  E3 ¼ 1 1 þ ð3:7Þ
2g r r4

(Note that the result given as Eq. (7.1) in Henderson (1966) is incorrect, and should
read as Eq. (3.7) above.)
Using one of these expressions for the loss term in Eq. (3.1) will enable the flow
conditions at section 1 to be determined from the conditions at section 3. The loss
term depends on the conditions at section 1, however, so an iterative solution is
again required. According to Henderson, the last term in the brackets in Eq. (3.7) is
not very significant unless the Fr1 exceeds 0.5 or the ratio b2/b1 is less than 1.5 (this
assumption would allow easier solution for preliminary design). The first condition
is not common, however, and the second would imply a very small head loss
anyway. If this term is neglected and it is assumed that y1= y2= y3, then a much
simpler equation for head loss results, i.e.

ðV1  V3 Þ2
DE ¼ ð3:8Þ
2g

Comparison of the predictions of this equation with experimental results


obtained by Formica (1955) indicates that it overpredicts energy losses by about
80 3 Open Channel Transitions

10%. None of these expressions for DE enable Eq. (3.1) to be solved without
iteration, and direct use of either Eq. (3.4) or (3.5) would provide the solution
anyway. They do, however, provide a rational form for developing empirical
quantification of the loss. For example, Eqs. (3.6) and (3.7) show that the loss can
be expressed as the upstream velocity head multiplied by a factor that depends on
flow conditions and the transition geometry. Equation (3.8) suggests a dependence
on both upstream and downstream velocities, while Eqs. (3.2) and (3.3) give the
loss in terms of the difference between upstream and downstream velocity heads.
Formica (1955) expressed the loss in terms of the downstream velocity head. All
forms have been used in the definition of empirical loss coefficients.
For some transition designs, it is advantageous to minimize the energy loss
(although whether this is a real advantage in all cases should be carefully consid-
ered). The energy loss for an expansion can be reduced by tapering the side walls
and curving them smoothly to eliminate sharp corners. The normally recommended
taper is 1:4 (1 unit width increment for every 4 units of length) for subcritical flows;
a more gradual taper is considered not to make further reductions in losses com-
mensurate with the associated increase in cost. A tapered expansion is more difficult
to analyse theoretically through momentum conservation, because the variation of
flow depth through the expansion makes the boundary force difficult to estimate.
The most practical approach is to apply energy conservation, through Eq. (3.1) with
the loss term accounted for empirically. It is reasonable to assume that the form of
the expansion loss equation would be similar to Eq. (3.8), and that the effect of
tapering can then accounted for by introducing an empirical coefficient. The energy
loss for a 1:4 straight taper, for example, is given by Henderson (1966) as

ðV1  V3 Þ2
DE ¼ 0:3 ð3:9Þ
2g

Smith and Yu (1966) investigated the energy losses in abrupt and tapered
transitions from a rectangular lined channel to a trapezoidal unlined channel. They
expressed the loss in terms of the upstream velocity head only (cf. Eqs. (3.6) and
(3.7)), with a term to represent the geometry and an empirical coefficient (CL) to
account for other effects, i.e.

V12
DEc ¼ CL ð1  Ar Þ2 ð3:10Þ
2g

in which Ar = by1/By3, b is the upstream channel width and B is the average width
of the downstream channel. They found that for gradual transitions (a taper of about
1:10) a straight wall flare produces smaller losses than a curved one (CL = 0.25
compared with 0.40). For a rapid transition with a taper of 1:4, CL is 0.50.
Skogerboe et al. (1971) investigated this type of expansion further. They found that
CL is not constant, but varies with the Froude number at the inlet and the expansion
ratio B/b. This is not surprising considering appearance of these terms in the
complete rational energy loss Eqs. (3.6) and (3.7) for the abrupt expansion.
3.2 Subcritical Flow Transitions 81

Short transitions can also include changes in bed level, with or without lateral
expansions or contractions. Empirical loss formulations are not widely documented
for these, but abrupt steps can be easily treated using momentum conservation,
similarly to lateral expansions and contractions.
The shape of a transition between the inlet and outlet sections has been identified
as important for reducing energy losses and ensuring satisfactory flow behaviour,
especially for avoiding flow separation and distorted velocity profiles. Vittal and
Chiranjeevi (1983) have reviewed and improved methods for shaping rectangular–
trapezoidal transitions and describing the flow conditions through them.
Smith and Yu (1966) found for their rectangular–trapezoidal transitions that
separation from one side occurred unless the taper is long (about 1:10). With their
shorter transition, separation occurred and the flow concentrated along one side,
resulting in scour in the downstream, unlined channel. They proposed evening out
the velocity profile by installing baffles extending within the transition and pro-
viding riprap protection at the beginning of the unlined channel. These increased
the value of CL to 0.8, so some energy loss is incurred in order to even out the
velocity profile. The structure is much shorter, however, than it would have to have
been without baffles to prevent separation.
Najafi-Nejad-Nasser and Li (2015) showed that flow separation and energy loss
in tapered expansions can be reduced by placing a low, gradual hump in the
transition. They investigated a 1:4 expansion with the bed rising from the entrance
to the expansion up to a crest with a height of 5–9% of the approach flow depth at
the end of the expansion and then dropping back over the same distance in the
wider channel. The hump has the effect of accelerating the flow and thereby neu-
tralizing the deceleration associated with the lateral expansion; this can reduce the
head loss by up to a factor of 4.
The wider setting of a transition should be considered in its design. A transition
will induce departure from uniform flow that can extend for a considerable distance
upstream. A contraction causes an afflux, which is exacerbated by energy loss,
requiring raised side walls for some distance. The afflux can be reduced or elimi-
nated by including a downward step in the transition, but this requires lowering of
the downstream channel which may be difficult to accommodate. It is therefore
advantageous to minimize the energy loss in a contraction.
For an expansion, it may not be desirable to reduce the energy loss. An
expansion causes drawdown in the upstream channel, resulting in the specific
energy at the entrance being lower than for the uniform flow upstream. The dif-
ference in specific energies between the upstream and downstream uniform con-
ditions is all dissipated along the gradually varied profile and through the transition.
Reducing the loss through the transition will increase the water level drawdown
immediately upstream and this will extend the length of the gradually varied profile
to increase the energy dissipated along it. This will also increase the boundary shear
stress upstream, which may cause scour problems in unlined channels. Allowing a
82 3 Open Channel Transitions

loss through the transition will reduce the drawdown and hence the extent of the
gradually varied flow profile upstream. The drawdown can also be reduced or
eliminated by incorporating an upward step in the transition.
For expansion transitions, subcritical flow will occur through the transition only
under certain conditions, even if the flow downstream would be expected to be
subcritical. If b3/b1 is too large, the drawdown of the upstream flow depth will
approach critical and flow after the expansion will be supercritical, with a hydraulic
jump to the downstream subcritical flow. Whether this happens depends on b3/b1,
the discharge and the downstream flow depth, and can be determined using specific
energy principles. This situation can be avoided by incorporating an upward step in
the transition.
The formation of waves can affect the flow through transitions, even for sub-
critical flow, especially if the flow is close to critical, but this is of more concern for
transitions with supercritical flow.
Example 3.1
A long, concrete-lined (n = 0.013) channel with a rectangular cross section and a gra-
dient of 0.00050 conveys a discharge of 35 m3/s. The channel width increases abruptly
in the direction of flow from 4.0 to 6.0 m.
Determine the flow depths immediately upstream and downstream of the expansion and
the energy loss through the expansion

a. using energy conservation, with losses determined using Henderson’s (1966) equation
and
b. using momentum conservation.
Sketch the water surface profile between the uniform conditions upstream and
downstream.
Solution

Q = 35 m3/s
4.0 m 6.0 m So = 0.00050
n = 0.013

Plan

ΔE
H

1 2
Long section
3.2 Subcritical Flow Transitions 83

Flow is uniform far upstream and downstream from the transition. Calculate flow depths
using the Manning equation with continuity, i.e.

A 2=3 1=2
Q¼ R So
n
Downstream:
 2=3
6:0 yo 6:0 yo
35 ¼ 0:000501=2
0:013 6:0 þ 2  yo

from which, by trial, yo = 2.69 m.


Upstream
 2=3
4:0 yo 4:0 yo
35 ¼ 0:000501=2
0:013 4:0 þ 2  yo

from which, by trial, yo = 4.16 m

a. y2 = yo = 2.69 m.

E1 ¼ E2 þ DE
q2
E2 ¼ y2 þ
2 g y22
ð35=6:0Þ2
¼ 2:69 þ ¼ 2:93 m
2  9:8  2:692
Energy loss by Henderson (1966):
 2 !
V12 1 Fr21 ðr  1Þ2
DE ¼ 1 þ
2g r r4

with
b2 6:0
r¼ ¼ ¼ 1:5
b1 4:0

Q 35
V1 ¼ ¼
A 4:0 y1

V1 35=4:0 y1
Fr1 ¼ pffiffiffiffiffiffiffiffi ¼ pffiffiffiffiffiffiffiffiffiffiffi
g y1 9:8 y1

Therefore
 2 !
 35 2 35    
4:0 y1 1 2 35=4:0 y1 2 ð1:5  1Þ2
E1 ¼ y1 þ 4:0
¼ 2:93 þ 1 þ pffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
2  9:8 y21 2  9:8 1:5 9:8  y1 1:54

from which, by trial, y1 = 2.3 m.


84 3 Open Channel Transitions

b. y2 = yo = 2.69 m
P
F
M1 þ ¼ M2
c
Q2
M2 ¼ þAy
gA
352 2:69
¼ þ 6:0  2:69 
9:8  6:0  2:69 2
¼ 29:45 m3
1
F ¼ c y21 ð6:0  4:0Þ
2
F
¼ y21
c

Therefore
Q2
þ A y þ y21 ¼ 29:45
gA
i.e.

352 y1
þ 4:0  y1 þ y21 ¼ 29:45
9:8  4:0  y1 2
from which, by trial, y1 = 2.3 m.
The water surface will approach the upstream uniform flow through an M2 profile, as
shown below. Energy is lost by friction through the gradually varied profile and by
expansion through the transition (DE).

ΔE
yo = 4.16 m H
y1 = 2.30 m y2 = yo yo = 2.69 m

1 2

3.3 Supercritical Flow Transitions

Supercritical flow transitions are common in chutes for spillways, diversions and
drop structures. The major problem experienced with them is the formation of
surface waves which raise the water level locally and require higher confining
walls. If the transition is not carefully designed a pattern of standing waves may be
set up which extends a long distance downstream. This section (drawn largely from
Henderson (1966)) contributes to understanding and predicting the locations and
heights of waves in transitions.
3.3 Supercritical Flow Transitions 85

Surface waves occur because any obstacle in a flow path generates a disturbance
which moves across the water surface. In supercritical flow, the disturbance cannot
be propagated upstream, and an oblique standing wave is formed. The formation of
such waves can be visualized by considering a series of instantaneous disturbances
on the surface of a body of still water. As shown in Fig. 3.4a, an instantaneous
disturbance will cause a circular wavefront centred at the point of disturbance (A1)
to travel outwards at a celerity, c, which depends on the water depth and the
wavelength.
If the disturbance is repeated at regular time intervals, a series of concentric
circular wavefronts will result (Fig. 3.4b). If the point of disturbance moves as well
(to points A2, A3…, etc.) (or the water moves relative to a stationary disturbance
point), then the pattern of circular wavefronts will no longer be concentric. If the
velocity of the point of disturbance is less than c, then the pattern shown in
Fig. 3.4c will result, with the wavefronts closer together on the upstream side of the
point than on the downstream side, but still moving away from the point. If the
velocity of the point of disturbance is just equal to c then the waves generated by
the disturbance accumulate ahead of the disturbance point, forming a shock front
which advances at V = c (see Fig. 3.4d). If the velocity of the point of disturbance
is greater than c then each new disturbance starts beyond the influence of the
previous one (Fig. 3.4e). By the time the point of disturbance has reached point An
there will be a series of wavefronts which can be enveloped by the common tangent
AnP1. If the disturbances are continuous rather than discrete, then a continuous
wavefront will form along this tangent.

c
(a) (b) c (c) c
c c
c V<c
c

. . V ...
A1 A1,2,... A3 A2 A1

P1 c
(d) c (e)
c V>c
c

. . . . β . . .
V V
A3 A2 A1 An A3 A2 A1

Fig. 3.4 Propagation of a small surface disturbance (plan view)


86 3 Open Channel Transitions

The direction of the tangent AnP1 is a defining characteristic of the wave pattern
and obviously depends on the relative values of the wave celerity and the velocity
of the disturbance; the greater V is relative to c, the smaller will be the angle b.
A surface wave will travel from A1 to P1 in the same time that the point of
disturbance travels from A1 to An. Therefore

A1 P1
sin b ¼
A1 An
c Dt ð3:11Þ
¼
V Dt
c
¼
V

The celerity and the velocity of the point of disturbance in this argument are
relative to the water, and therefore it makes no difference whether the point of
disturbance is moving in stationary water, or water is moving past a stationary point
of disturbance. So a slight irregularity on the side of a channel will set up a shock
wave under supercritical flow as shown in Fig. 3.5.
The orientation of this standing wave can be found from Eq. (3.11). The flow
velocity, V, can be found from an appropriate control relationship. The celerity for
oscillatory waves is given by
 
gL 2py
c2 ¼ tanh ð3:12Þ
2p L

in which L is the wavelength. If L is large compared with y, then 2py/L is small and
tanh(2py/L)  2py/L. It then follows that for long waves of low amplitude

c2 ¼ g y ð3:13Þ

Substituting this expression for c in Eq. (3.11) gives

1
sin b ¼ ð3:14Þ
Fr

In practice, disturbance waves are neither of long wavelength nor low amplitude.
Analysis of surges in channels gives results which show clearly that they travel at

Fig. 3.5 Oblique standing


β
wave in a uniform channel
(plan view)
V
3.3 Supercritical Flow Transitions 87

velocities greater than (gy)1/2. Equation (3.14) must therefore be refined for prac-
tical purposes. The treatment is slightly different for straight and curvilinear tran-
sitions, and these are treated separately.

3.3.1 Straight Transitions

The effect of disturbances being large can be examined by considering the wave-
front set up in a channel by the deflection of a vertical wall through an angle Dh, as
shown in Fig. 3.6. For design purposes, it is necessary to determine the direction of
the shock front (i.e. the angle b) and the height of the wave (i.e. the value of y2 in
excess of the given y1, which is the depth controlled from upstream for supercritical
flow).
This wavefront will cause a finite change in depth and Eq. (3.14) will not predict
b reliably. The situation can be analysed by treating the wavefront as a hydraulic
jump with a superimposed velocity component parallel to the front.
The momentum equation can now be applied in the direction normal to the
wavefront.

M1 ¼ M2

i.e.

q2 y2 q2 y2
þ 1¼ þ 2
g y1 2 g y2 2

Δθ
β

> V2 sin (β-Δθ)


V1
V1 V2 cos (β-Δθ)
>

V1 cos β V2 V2
V1 sin β (β-Δθ)
>

shock front

Fig. 3.6 Standing wave set up by wall deflection (plan view showing streamlines)
88 3 Open Channel Transitions

and so
 
q2 1 1 1 
 ¼ y22  y21
g y1 y2 2

Substituting q = V1 sin b y1 and rearranging gives


 
V12 sin2 b 1 y2 y2
¼ þ1
g y1 2 y1 y1

which is the same as for an ordinary hydraulic jump, except that V1 is replaced by
V1sinb. On the left-hand side of this equation, V21/gy1 can be replaced by Fr21. The
equation can then be rearranged to give an equation for b which is applicable for
waves with finite amplitude, i.e.
  1=2
1 1 y2 y2
sin b ¼ þ1 ð3:15Þ
Fr1 2 y1 y1

Note that if the amplitude is small then y2/y1 tends to unity and Eq. (3.15)
reduces to Eq. (3.14), which applies to small amplitude long waves.
To solve Eq. (3.15) for finite amplitude waves, it is also necessary to know
something about the ratio of the flow depths, y2/y1. A simple relationship between
y1 and y2 can be derived by applying continuity across the wavefront and relating
the velocities parallel to it. Applying continuity normal to the wavefront gives

V1 y1 sin b ¼ V2 y2 sinðb  DhÞ ð3:16Þ

The velocity along the front must be the same on both sides because there is no
force in this direction associated with the jump. Therefore

V1 cos b ¼ V2 cosðb  DhÞ ð3:17Þ

Eliminating V1/V2 from Eqs. (3.16) and (3.17) leads to

y2 tan b
¼ ð3:18Þ
y1 tanðb  DhÞ

The height of the disturbance wave and the direction of the shock front can be
determined by simultaneous solution of Eqs. (3.15) and (3.18). This can be done by
substituting Eq. (3.18) for y2/y1 in Eq. (3.15), solving for b, and then using this
value to determine y2/y1 from Eq. (3.18).
3.3 Supercritical Flow Transitions 89

Example 3.2
The width of a rectangular, concrete-lined channel is reduced in width from 8.0 to 5.0 m
through an 18.0 m long straight wall transition. At the design discharge of 50 m3/s, the
flow depth in the wider channel approaching the transition is 0.55 m. Determine

a. the minimum height required for the side walls through the transition and
b. the maximum flow depth within the transition and its location.

Solution
18.0 m
Δθ
β
l
8.0 m 5.0 m

1 2 (downstream of shock front)

a. The minimum side wall height is the flow depth after the shock front (y2), given by
simultaneous solution of Eqs. (3.15) and (3.18), i.e.

  1=2
1 1 y2 y2
sin b ¼ þ1 ð3:15Þ
Fr1 2 y1 y1

and

y2 tan b
¼ ð3:18Þ
y1 tanðb  DhÞ

Substituting (3.18) in (3.15):


  1=2
1 1 tan b tan b
sin b ¼ þ1
Fr1 2 tan ðb  DhÞ tan ðb  DhÞ

1:5
Dh ¼ arctan ¼ 4:76
18

V1
Fr1 ¼ pffiffiffiffiffiffiffiffi
g y1

Q 50
V1 ¼ ¼ ¼ 11:4 m/s
A1 8:0  0:55
11:4
¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ 4:91
9:8  0:55
  1=2
1 1 tan b tan b
¼ þ 1
4:91 2 tanðb  4:76 Þ tan ðb  4:76 Þ

from which, by trial, b = 15.7°


90 3 Open Channel Transitions

Substitute b = 15.7° in Eq. (3.18) to get y2:

tan b
y2 ¼ y1
tanðb  DhÞ
tan 15:7
¼ 0:55 ¼ 0:80 m
tanð15:7  4:76 Þ

Therefore, the minimum side wall height is 0.80 m.

b. The maximum flow depth within the transition occurs where the shock front s from the
two sides meet on the centre line.

Height of combined shock wave ¼ y1 þ 2ðy2  y1 Þ


¼ 0:55 þ 2ð0:80  0:55Þ ¼ 1:05 m

Location from beginning of transition,


8:0=2

tan b
4:0
¼ ¼ 14:2 m
tan 15:7

3.3.2 Curvilinear Transitions

The previous analysis applies to straight wall transitions where Dh is finite and
constant. For curvilinear transitions, Dh varies continuously through the transition
and there is no abrupt shock front, but rather a gradual variation through a smooth
wave. The change in flow depth can be analysed by applying Eq. (3.18) to very
small changes.
According to Henderson (1966), setting y2 = y1 + Dy and letting Dh tend to zero
leads to

dy y
¼ ð3:19Þ
dh sin b cos b

Because very small increments of y are now being considered, Eq. (3.11) is valid
and sinb can be replaced by c/V. Also, for small amplitudes c = (gy)1/2. Therefore

dy y sin b
¼
dh sin b cos b sin b
y
¼ 2 tan b
sin b
y V2
¼ 2 tan b
c
y V2
¼ tan b
gy
3.3 Supercritical Flow Transitions 91

and so

dy V 2
¼ tan b ð3:20Þ
dh g

This equation can be used to describe the variation of flow depth along the wall
of a curved transition. At any angle h, the flow depth y can be calculated and this
will also be the depth along a line radiating from the wall at a tangential angle b, as
defined by Eq. (3.14). This is illustrated in Fig. 3.7.
Ippen (1950) integrated Eq. (3.20) under the assumption that there is no energy
loss along such a continuous change in depth and that the specific energy is
therefore constant. The resulting relationship between y and h (in degrees) along the
wall is
pffiffiffi
pffiffiffi 1 3 1
h ¼ 3 tan pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  tan1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  h1 ð3:21Þ
Fr  1
2
Fr2  1

The constant of integration h1 can be evaluated by inserting the boundary


condition h = 0° at y = y1. If the specific energy, E, is calculated for the initial flow
depth and this is assumed to remain constant, then Fr can be calculated for any
specified value of y from the specific energy relationship expressed as
 
Fr2
E ¼ y 1þ ð3:22Þ
2

The angular position of the specified flow depth can then be calculated from
Eq. (3.21).
The procedure for defining the water surface profile along a side wall is then first
to determine the flow depth at the entrance to the transition (where h = 0) from
upstream conditions, calculate the corresponding value of Fr and hence the value of

θ
θmax
β ymax

Flow

θ increasing θ decreasing
y increasing y decreasing
converging +ve waves diverging -ve waves

Fig. 3.7 Wave pattern along a curved boundary (plan view; the oblique lines are water surface
contours)
92 3 Open Channel Transitions

h1 from Eq. (3.21). The value of E through the contraction is assumed to be


constant, and can therefore also be calculated from the entry condition. A range of
flow depths anticipated to occur through the transition is then specified, and for
each the corresponding value of Fr can be calculated from Eq. (3.22) and then its
angular position h from Eq. (3.21). Alternatively, values of h can be specified and
the corresponding values of Fr obtained (by trial) from Eq. (3.21) and hence y from
Eq. (3.22). The maximum flow depth at the maximum wall angle, hmax, is partic-
ularly important for design. The transition geometry will define the relationship
between angular and linear position. Ippen (1950) also produced a graphical
solution to Eq. (3.20) (which is reproduced in Henderson (1966)).
Henderson (1966) reports good agreement of Eq. (3.21) with experimental
results, except that the actual point of maximum depth occurred slightly beyond the
inflection point rather than right at it as predicted.
Ippen (1950) also integrated Eq. (3.20) under the assumption that the velocity
(rather than E) would be constant through the deflection. The resulting equation is
simpler and gives slightly lower values than Eq. (3.21). This equation is
 
y2 h
¼ Fr21 sin2 b1 þ ð3:23Þ
y1 2

with b1 = sin−1 (1/Fr1).


A contraction of supercritical flow presents the possibility of choking if the
downstream channel is too narrow for the available upstream energy to accom-
modate. This can be checked for using the specific energy concepts presented in
Chap. 1. Henderson (1966) identifies a further range of conditions where choking is
not assured, but is possible.
Although the theory for supercritical flow through transitions has been presented
for the contraction case, it applies for expansions as well.
Example 3.3
The width of a rectangular, concrete-lined channel is reduced in width from 8.0 to 5.0 m
through an 18.0 m long curvilinear transition. In plan, the side walls follow a simple
reverse curve formed by two equal circular arcs. At the design discharge of 50 m3/s, the
flow depth in the wider channel approaching the transition is 0.55 m. Determine

a. the minimum height required for the side walls through the transition and
b. the maximum flow depth within the transition and its location.

Solution

a. The minimum side wall height is the maximum flow depth, which occurs at the location
of hmax. (The curvature shown in the plan sketch is exaggerated for clarity.)
3.3 Supercritical Flow Transitions 93

18.0 m

θmax
90o-α
1.5 m
90o-α
α

1.5
α = tan −1 = 4.76o
18.0
θ max = α +α
o
= 4.76 + 4.76 = 9.52

Assuming no energy loss through the transition, the variation of y is described by


Ippen’s (1950) relationship (Eq. 3.21):
pffiffiffi
pffiffiffi 1 3 1
h ¼ 3 tan p ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  tan1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  h1
Fr2  1 Fr2  1

The constant h1 is obtained by defining h = 0 at the beginning of the transition,


where Fr1 = 4.91 (from Example 3.2).
Therefore
pffiffiffi
pffiffiffi 3 1
h1 ¼ 3 tan1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  tan1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ 22:6
4:912  1 4:912  1

ymax occurs where h = hmax, i.e.


pffiffiffi
pffiffiffi 3 1
hmax ¼ 9:52 ¼ 3 tan1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  tan1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  22:6
Fr  1
2
Fr2  1

from which, by trial, Fr = 3.32.


For no energy loss through the transition, E = constant = E1, and (with
V1 = 11.4 m/s from Example 3.2)

V12
E1 ¼ y1 þ
2g
11:42
¼ 0:55 þ ¼ 7:18 m
2  9:8
Therefore, at hmax
 
Fr2
E ¼ 7:18 ¼ ymax 1 þ ð3:22Þ
2

i.e.
7:18
ymax ¼ 2 ¼ 1:10 m
1 þ 3:32
2
94 3 Open Channel Transitions

Therefore, the minimum side wall height is 1.10 m.

b. The maximum flow depth occurs where the wave height corresponding to hmax from
either side meets on the centre line.

θmax
4.0 m
3.25 m β
2.5 m

9.0 m l

Combined wave height ¼ y1 þ 2ðymax  y1 Þ


¼ 0:55 þ 2ð1:10  0:55Þ ¼ 1:65 m

Location:

1
sin b ¼
Fr
1
¼ ¼ 0:301 ) b ¼ 17:5
3:32
Then

l ¼ 3:25 tanð90  ðhmax þ bÞÞ


¼ 3:25 tanð90  ð9:5 þ 17:5 ÞÞ ¼ 6:38 m

3.3.3 Suppression of Standing Wave Propagation

At points where disturbance lines from opposite sides of the channel intersect, the
resulting change in depth is the algebraic sum of the depth changes on the inter-
secting lines. The point of intersection acts as a new disturbance, producing dis-
turbance lines directed back towards the channel sides. These are, in turn, reflected
once more to produce disturbance lines which again intersect in the centre of the
channel. This process of alternate disturbance reflection against the walls and at the
channel centre produces the commonly observed diamond-shaped pattern of obli-
que standing waves, which can persist for a considerable distance downstream. This
is an undesirable flow condition and can be minimized by using a contraction with
straight side walls and careful choice of the flare angle. The idea behind this design
is shown in Fig. 3.8.
The points of contraction at the entrance to the transition create wavefronts from
each side of the channel at angles b1 to the original wall directions. These
3.3 Supercritical Flow Transitions 95

wavefronts intersect and reflect at the channel centre, directing wavefronts back to
the side walls. If they were to reach the side walls before or after the end of the
transition, they would be reflected back towards the channel centre. At the end of
the transition, the change in direction of the walls again creates a disturbance, but
because the change represents an expansion rather than a contraction, the wave-
fronts emanating from these points are negative. The idea for suppressing the wave
propagation is simply to select the transition length, L, (and hence the angle h) so
that the wavefronts reflected at the channel centre meet the side walls exactly at the
end of the transition, so that the positive wave is counteracted by the negative wave.
If the discharge varies, it is difficult to produce a design that will be correct for the
whole range. However, if the approach is controlled by friction, as for a long chute,
then the variation of velocity with discharge is not great and the solution will not be
very sensitive to discharge variations.
Sturm (1985) has provided a solution method for designing a straight wall
contraction to minimize standing wave propagation. He shows that the correct
solution will be obtained by applying Eqs. (3.15)–(3.18) across the two wavefronts
simultaneously with an equation for continuity through the transition. This conti-
nuity equation can easily be expressed in terms of the transition contraction ratio
(r) as
 
1 b1 Fr3 y3 3=2
¼ ¼ ð3:24Þ
r b3 Fr1 y1

The solution process is laborious because of the simultaneous application over


two wavefronts and Sturm provided a graphical solution, which is reproduced in
Fig. 3.9. In using this diagram, the upstream flow conditions (Fr1) and the ratio of
channel widths to be connected (r) will be known. The lower part of the diagram
provides the best value of h, from which the transition length, L, can be determined,
and the upper part provides the ratio of upstream and downstream flow depths for

β1
β2
b1 b3
V3
V2 β2-θ
V1

Fig. 3.8 Standing wave patterns in a straight-walled contraction designed to suppress wave
propagation (plan view)
96 3 Open Channel Transitions

this design. This part shows clearly how the wave height increases with Fr1, and
just how large the waves can be.
Supercritical contractions will choke if the contraction ratio is too small for the
design flow conditions. The critical contraction ratio can be determined by applying
specific energy principles, as discussed in Chap. 1. This analysis implies the
occurrence of a hydraulic jump upstream of the contraction and critical flow at the
end; the associated choking conditions are represented by line A in Fig. 3.9.
A second possible choking mechanism has been recognized (Henderson 1966;
Sturm 1985) where the gradually rising water level through the contraction reaches
the critical depth. This condition can be predicted through Sturm’s (1985) analysis
for Fr3 = 1, with allowance for energy loss, and is indicated by line B in Fig. 3.9.
Sturm suggests that choking is possible for design conditions between lines A and
B, but certain for designs to the right of line B. The near vertical orientation of line
A suggests a limit of about 4o (1:14) for h to ensure no choking. More detailed

12

10
20

y3/y1 8

10 6

5 Fr1
4
3

0
A

B
r = b3/b1 2
Fr1
2.5
0.5
3
4
5
6
8 10
0
10 20 30
θ (degrees)

Fig. 3.9 Design graph for straight wall transitions (adapted from Sturm 1985)
3.3 Supercritical Flow Transitions 97

analysis of the choking condition in a linear contraction is presented by Defina and


Viero (2010).
In addition to being designed to eliminate the reflection of wavefronts, straight
wall transitions cause smaller increase in flow depth than curved ones.
Equation (3.21) shows that the maximum depth is determined by the maximum
deflection angle. For curved transitions, whatever their form, the maximum angle is
greater than the constant angle in a straight transition, and hence the increase in
depth must be greater. It is also worth noting that the curve radius does not feature
in the equation and therefore has no effect on the maximum depth; it will only affect
the relative positions of different depths through the profile.
More advanced methods have moved away from analytical and experimental
results and towards computational modelling of flow (e.g. Abdo et al. 2019;
Bhallamudi and Chaudhry 1992; Jiminez and Chaudhry 1988; Mazumder and
Hager 1993).
Example 3.4
Determine the best length of the straight wall transition in Example 3.2 for suppression
of standing wave propagation.
Solution
The length of the transition can be expressed in terms of the difference in upstream and
downstream channel widths, and the angle Dh, i.e.

ð8:0  5:0Þ=2

tan Dh
For suppression of standing wave propagation, Dh is obtained from the graphical
solution of Sturm (1985) (Fig. 3.9) in terms of the width ratio b2/b1 and Fr1.
From Example 3.2, Fr1 = 4.91
and

b3

b1
5:0
¼ ¼ 0:63
8:0
Hence, from Fig. 3.9, Dh = 3.5o and so

ð8:0  5:0Þ=2
L¼ ¼ 24:5 m
tan 3:5

3.4 Dual Stable States and Hysteresis

For transitions involving changes from supercritical flow upstream to subcritical


flow downstream, situations occur where different stable states of flow are possible,
and the one that actually occurs depends on prior conditions (similar to the hys-
teretic behaviour described for underflow gates).
98 3 Open Channel Transitions

The change from supercritical to subcritical flow takes place through a hydraulic
jump, which occurs either upstream or downstream of the transition structure,
depending on the relative magnitudes of the momentum function immediately
before or after the structure. If the approaching supercritical flow has sufficient
energy to pass the structure and the resulting supercritical flow after the structure
has greater momentum than the downstream subcritical flow, then the jump will
occur further downstream. Similarly, if the downstream subcritical flow can persist
through the structure and the resulting subcritical flow before the structure has
greater momentum than the approaching supercritical flow, then the jump will occur
further upstream. Analysis of the transition can be related to the control conditions
either upstream or downstream, i.e. beginning either with the supercritical flow
upstream or the subcritical flow downstream. However, for certain flow conditions
and transition geometries, contradictory conclusions arise from the two approaches,
with the jump location being predicted as either downstream or upstream (see
Example 3.5). Both are stable states and the actual condition depends on prior flow
conditions, i.e. the behaviour is hysteretic.
For an upward step in a channel (Fig. 3.10), for example, the conditions for the
hydraulic jump to occur further downstream (solid water surface profile) are
therefore

E1  Dz  DE1 [ Ec ð3:25Þ

and
M2 [ Md ð3:26Þ

where DE1 is the energy loss associated with supercritical flow over the
step. Equation (3.26) means that y2 must be less than the conjugate of yd.
The conditions for the hydraulic jump to occur further upstream (dashed water
surface profile) are

E1 ¼ Ed þ Dz þ DE2 ð3:27Þ

and

M1 [ Mu ð3:28Þ

where DE2 is the energy loss associated with subcritical flow over the step. For a
smooth step, the energy losses are negligible and the hydraulic jump will be far
upstream or downstream from the step after a gradually varied profile. If the step is
steep, the energy loss becomes appreciable, and the hydraulic jumps move closer to
the step and the range of conditions for dual states decreases. For an abrupt step
(such as in a stilling basin), the range becomes negligibly small, and only one state
is likely; this has been confirmed by data obtained by Muskatirovic and Batinic
(1977), as presented by Viero and Defina (2019).
3.4 Dual Stable States and Hysteresis 99

Fig. 3.10 Supercritical–


subcritical transitions across yd
an upward step
yu Δz

1 2

Viero and Defina (2017) and Defina and Susin (2006) have generalized these
concepts for application to different transition types.
Example 3.5
The slope of a long, wide, concrete-lined channel changes from 0.020 to 0.00020 over a
short distance. Determine if the resulting hydraulic jump is upstream or downstream of
the transition when the discharge is 12 m3/s/m

a. if the two slopes meet at the same level and


b. if a short, smooth, upward step of 1.0 m is included in the transition.

Solution
First, establish the uniform flow conditions on the two slopes.
For So = 0.020,
Calculate yo from Manning equation with continuity (for a wide channel R  y)
(subscript u indicates upstream value):

y 2=3 1=2
q¼ y S
n  
Therefore
!3=5
qn
yo u ¼ 1=2
S
 
12  0:013 3=5
¼ ¼ 1:06 m
0:0201=2
sffiffiffiffiffi rffiffiffiffiffiffiffi
3 q
2 3 12
2
Critical flow depth ¼ yc ¼ ¼ ¼ 2:45 m
g 9:8

you < yc, so first slope is Steep


Also
q2 y2
Mo u ¼ þ ou
g yo u 2
122 1:062
¼ þ ¼ 14:4 m2
9:8  1:06 2
100 3 Open Channel Transitions

For So = 0.00020,
yod (subscript d indicates downstream):
!3=5
qn
yo d ¼ 1=2
So
 3=5
12  0:013
¼ ¼ 4:22 m
0:000201=2

yod > yc, so second slope is Mild


Also
q2 y2
Mo d ¼ þ od
g yo d 2
122 4:222
¼ þ ¼ 12:4 m2
9:8  4:22 2

a. Compare the M values for each slope at the slope transition:

Mo u [ Mo d ð14:4 m2 [ 12:4 m2 Þ

and therefore the supercritical flow advances on to the Mild slope and the hydraulic jump
is downstream of the transition.

b. Assume upstream control, i.e. you persists up to the step, and so y1 = you.

you
Δz = 1.0 m

1 2

The step is smooth, so there are no significant energy losses between 1 and 2. y2 can
then be calculated by energy conservation, i.e.

E2 ¼ E1  Dz

or
q2 q2
y2 þ ¼ y1 þ  Dz
2 g y22 2 g y21
122 122
y2 þ ¼ 1:06 þ  1:00 ¼ 6:60 m
2  9:8 y2
2 2  9:81  1:062

from which, by trial, y2 = 1.16 m and the supercritical flow has sufficient energy to
pass the step.
3.4 Dual Stable States and Hysteresis 101

Then
q2 y2
M2 ¼ þ 2
g y2 2
122 1:162
¼ þ ¼ 13:3 m2
9:8  1:16 2
and M2 [ Mo d so the hydraulic jump is further downstream.
Now assume downstream control, i.e. yod persists up to the step, and so y2 = yud.

yod

Δz = 1.0 m

1 2

Then

E1 ¼ E2 þ Dz

or
q2 q2
y1 þ 2
¼ y2 þ þ Dz
2 g y1 2 g y22
122 122
y1 þ ¼ 4:22 þ þ 1:00 ¼ 5:63 m
2  9:8 y1
2 2  9:81  4:222

from which, by trial, y1 = 5.38 m.


Then
q2 y2
M1 ¼ þ 1
g y1 2
122 5:382
¼ þ ¼ 17:2 m2
9:8  5:38 2
and M1 [ Mo u so hydraulic jump is further upstream.
The two assumptions regarding the direction of control both yield valid, but con-
tradictory solutions, indicating hysteretic behaviour. Similar calculations for a range
of discharges show that the two assumptions will both predict the hydraulic jump
upstream for q < * 9 m3/s/m and downstream for q > * 22 m3/s/m. Contradictory
predictions occur for discharges between these values.

Problems
3:1. A long, rectangular, concrete (n = 0.013) channel has a gradient of 0.00050
and is required to change in width from 3.5 to 5.0 m. The design discharge is
25 m3/s. It is required to determine the flow depths immediately upstream and
downstream of the transition and the energy loss through the transition.
102 3 Open Channel Transitions

a. The first design to be considered is for a flat bed abrupt transition. Analyse
the flow through the transition using the following approaches:
i. Energy conservation, assuming no losses.
ii. Energy conservation, assuming losses as described by Henderson’s
(1966) equation (Eq. (3.7)).
iii. Energy conservation, assuming losses according to Eq. (3.8).
iv. Momentum conservation, using the momentum function.
v. Momentum conservation, using the force–momentum flux equation.
Compare the results obtained from the different analyses, and
comment on the differences.
vi. Sketch the water surface profile upstream, through and downstream
of the transition.
vii. Explain how the transition affects the bed shear stress in the
upstream approach flow.
viii. How would the upstream bed shear stress be changed if an abrupt
upward step of 0.50 m was included in the transition? What would
be the energy loss through the transition?
ix. What step height would maintain uniform flow from upstream right
to the transition? What would be the energy loss? By how much
would the water level change over the transition?
b. The second design to be considered is a flat bed transition with a taper of
1:4. Analyse the flow through this transition using the following
approaches:
i. Energy conservation, accounting for losses appropriately.
ii. Momentum conservation, assuming the flow depth to vary linearly
through the transition.
Compare these results and comment. How does the taper influence the
upstream flow depth and energy loss in comparison with the abrupt
expansion?
iii. How does this transition affect the bed shear stress upstream?
Recommend a design for this transition.
3:2. Repeat problem 3.1.a.iii but for a channel on a slope of 0.0010. Explain the
flow condition and sketch the water surface profile. What would be a practical
design configuration for this situation?
3:3. A long, rectangular, concrete (n = 0.013) channel has a gradient of 0.00050 and
is required to change in width from 5.0 to 3.5 m. The design discharge is 25 m3/
s. Determine the flow depths immediately upstream and downstream of the
transition using Chow’s (1959) method and
a. assuming no losses,
b. assuming a square-ended transition and
c. assuming a cylinder quadrant transition.
3.4 Dual Stable States and Hysteresis 103

3:4. Water is to be released from a reservoir into a 5.0 m wide rectangular channel
at a flow depth of 1.0 m and a velocity of 12.0 m/s. The channel is to be
connected immediately to a 4.0 m wide chute through a 10.0 m long transi-
tion. Two transition shapes are being considered. The first is curvilinear, with
each side wall following the shape of a simple reverse curve consisting of two
equal circular arcs. The second has straight walls. For each transition shape,
estimate
a. the maximum flow depth along the side walls (using both of Ippen’s (1950)
solutions for the curvilinear transition) and
b. the maximum flow depth along the centre line of the transition and its
location.
3:5. Water is to be released from a reservoir into a 6.0 m wide rectangular channel
at a discharge of 98.8 m3/s with a specific energy of 10.8 m. A short distance
downstream the channel must be contracted to a width of 3.6 m. Determine
the length of straight wall transition to minimize surface wave transmission
using the graphical solution proposed by Sturm (1985).

References

Abdo, K., Riahi-Nezhad, C. K., & Imran, J. (2019). Steady supercritical flow in a straight-wall
open-channel transition. Journal of Hydraulic Research, 57(5), 647–661.
Bhallamudi, S. M., & Chaudhry, M. H. (1992). Computation of flows in open channel transitions.
Journal of Hydraulic Research, 30(1), 77–93.
Chow, V. T. (1959). Open-channel hydraulics. McGraw-Hill.
Defina, A., & Susin, F. M. (2006). Multiple states in open channel flow. In M. Brocchini & F
Trivellato (Eds.), Vorticity and turbulence effects in fluid structure interactions—Advances in
fluid mechanics (pp. 105–130). Wessex Institute of Technology Press.
Defina, A., & Viero, D. P. (2010). Open channel flow through a linear contraction. Physics of
Fluids, 22(5), 056601.
Formica, G. (1955). Esperienze preliminari sulle perdite di carico nei canali dovute a cambiamenti
di sezione (Preliminary tests on head losses in channels due to cross-sectional changes),
L’Energia elletrica, Milan, 32(7) (cited in Henderson (1966)).
Henderson, F. M. (1966). Open channel flow. Macmillan.
Ippen, A. T. (1950). Channel transitions and controls. In H. Rouse H (Ed.), Engineering hydraulics
(Ch. VIII). Wiley.
Jiminez, O. F., & Chaudhry, M. H. (1988). Computation of supercritical free-surface flows.
Journal of Hydraulic Engineering, 114(4), 377–395.
Mazumder, S. K., & Hager, W. H. (1993). Supercritical expansion flow in Rouse modified and
reversed transitions. Journal of Hydraulic Engineering, 119(2), 201–219.
Muskatirovic, D., & Batinic, D. (1977). The influence of abrupt changes of channel geometry on
hydraulic regime characteristics. In Proceedings 17th IAHR Congress (pp. 397–404). Baden
Baden.
Najafi-Nejad-Nasser, A., & Li, S. S. (2015). Reduction of flow separation and energy head losses
in expansions using a hump. Journal of Irrigation and Drainage Engineering, 141(3),
04014057.
104 3 Open Channel Transitions

Skogerboe, G. V., Austin, L. H., & Bennett, R. S. (1971). Energy loss analysis for open channel
expansions. Journal of the Hydraulics Division, ASCE, 97(HY10), 1719–1736.
Smith, C. D., & Yu, J. N. G. (1966). Use of baffles in open channel expansions. Journal of the
Hydraulics Division, ASCE, 92(HY2), 1–7.
Sturm, T. W. (1985). Simplified design of contractions in supercritical flow. Journal of Hydraulic
Engineering, 111(5), 871–875.
Viero, D. P., & Defina, A. (2019). Multiple states in flow through a sluice gate. Journal of
Hydraulic Research, 57(1), 39–50.
Vittal, N., & Chiranjeevi, V. V. (1983). Open channel transitions: Rational method of design.
Journal of Hydraulic Engineering, 109(1), 99–115.
Yarnell, D. L. (1934). Bridge piers as channel obstructions. Technical Bulletin, 442 (US
Department of Agriculture).
Viero, D. P., & Defina, A. (2017). Extended theory of hydraulic hysteresis in open-channel flow.
Journal of Hydraulic Engineering, 143(9), 06017014.
Chapter 4
Spillways

4.1 Introduction to Conveyance Structures

Conveyance structures include spillways, canals, transitions, culverts, diversions,


drops, intakes and siphons. The fundamental design problem for such structures is
the specification of the size and geometry of the structure to accommodate the
required discharge. This is commonly achieved through a process of iterative
analysis, i.e. a certain configuration is assumed and then analysed to determine its
discharge capacity. The initial configuration is then modified and reanalysed until
the required discharge capacity is obtained.
While the discharge through a structure is influenced by various geometric char-
acteristics, it is controlled at one position only. Estimation of the discharge through
the whole structure requires identification of the effective control feature, determi-
nation of the energy level at its location and application of its control relationship.
A hydraulic structure may incorporate a number of features which could act as
controls. For any particular condition, one of these will determine the discharge
through the structure. At other conditions, however, different features may be the
effective controls. It is important, therefore, to be able to predict which feature will
be the effective control for the condition under consideration.
These concepts can be illustrated by considering a closed conduit leading from a
reservoir and incorporating three adjustable gates, as shown in Fig. 4.1. Associated
with each gate is a control relationship between the energy level immediately
upstream (H) and the discharge (Q), i.e. Q = fi(Hi), in which the subscript i denotes
the particular control. Suppose gate 1 is set to be able to pass a very much smaller
discharge than either gate 2 or gate 3 (Fig. 4.2a). The discharge through the whole
structure will then be determined by the control relationship for gate 1, with the
energy level defined by the water level in the reservoir (HW), i.e. Q = f1(HW). Any
reasonably small adjustments to the settings of gates 2 and 3 will have no influence
on the discharge. Clearly, gate 1 is the effective control. Flow downstream of it is

© Springer Nature Switzerland AG 2020 105


C S James, Hydraulic Structures,
https://doi.org/10.1007/978-3-030-34086-5_4
106 4 Spillways

Q1 = f1(HW) Q2 = f2(H2) Q3 = f3(H3)


1 2 3

HW

Fig. 4.1 Hypothetical conveyance structure

supercritical and no features or characteristics of the structure further downstream


can have any influence on the discharge.
If gate 2 is set to be able to pass a very much smaller discharge than the other
two (Fig. 4.2b), then it will be the effective control and the discharge will be given
by Q = f2(H2). The local energy level, H2, must be calculated by subtracting from
HW all the losses up to section 2, i.e.

H2 ¼ HW  hL1  hf 12 ð4:1Þ

in which hL1 is the loss associated with flow through gate 1 and hf1-2 is the friction
loss between gates 1 and 2. The conduit will flow full upstream of the control and
flow downstream will be supercritical. Small adjustments to control 3 will have no
effect on the discharge through the structure. Adjustments to gate 1 will affect the
discharge, however, because hL1 depends on the gate setting.
Similarly, if gate 3 is set to be able to pass a very much smaller discharge than
the others (Fig. 4.2c), then it will be the effective control and the discharge will be
given by Q = f3(H3), with H3 given by

H3 ¼ HW  hL1  hf 12  hL2  hf 23 ð4:2Þ

in which hL2 is the loss through gate 2 and hf2-3 is the friction loss between gates 2
and 3.
Determining the discharge through the structure therefore requires identifying
the effective control and applying the appropriate control relationship using the
local energy level, which accounts for all losses upstream of the control section.
Nothing downstream of the control affects the discharge, so no downstream char-
acteristics need to be accounted for.
The major difficulty is identifying the effective control. In the example above,
this was made quite obvious, but it is not always so (it would not be, for example, if
all the gates in the hypothetical structure were set with equal openings). In many
real structures, the potential control features are often quite different, such as
overflow sections, contractions or lengths of closed conduit. The effective control
4.1 Introduction to Conveyance Structures 107

(a) Q = f1(HW)

HW

(b)
Q = f2(H2)
hL1
hf 1-2

HW
H2

(c) Q = f3(H3)
hL1
hf 1-2
hL2 hf 2-3
HW
H3
Q

Fig. 4.2 Different positions of effective control in hypothetical structure

should be identified by applying the principle that it is the feature that has the
highest energy requirement to pass a given discharge. For a given HW, therefore,
the discharge can be calculated assuming each potential control to be the effective
one. The lowest discharge will be the correct one, and indicates the effective control
for that value of HW. Sometimes the discharge is specified and HW is required. By
the same principle, this can be determined by calculating it assuming each potential
control to be the effective one and selecting the highest value.
108 4 Spillways

4.2 Spillway Structures

The most common use of spillways is in dam structures, where they are installed to
discharge excess or flood water which cannot be absorbed into storage. They are
also used at diversion structures to bypass flows in excess of diversion require-
ments, and as control structures to induce required flow depths.
The spillway is often the most expensive component of a dam and, from a safety
point of view, one of the most critical design aspects. Many dam failures in the past
can be attributed to inadequate spillway design.
An essential preliminary to the hydraulic design of a spillway is a hydrological
analysis to determine the design flood. This will begin by the adoption of an
acceptable level of risk of failure. Usually expressed as a flood frequency or return
period, this may be defined by legislation or policy, or may be based on economic
considerations if there is no danger to life or infrastructure downstream. A design
flood hydrograph for the site is then produced by an appropriate hydrological
model. If the reservoir is large, the attenuation effects should also be accounted for
by reservoir routing. If there are no constraints on flood discharge downstream, the
size of the spillway may be optimized together with the flood storage capacity to
obtain the most economical design.
Various types and configurations of spillways are commonly used. Selection
depends on factors such as the type of dam, the capacity required, and the geology
and topography of the site.
A spillway structure will generally comprise the following components:
– A control structure, which defines the relationship between the water level in
the reservoir and the discharge, such as a simple overflow crest, for example.
– A discharge channel, to convey the flow away from the control structure, such
as a spillway face, chute or closed conduit.
– A terminal structure, usually in the form of an energy dissipator (to be dis-
cussed in Chap. 6).
– Entrance or outlet channels may be required to convey the flow to the control
structure, as in a col spillway, or from the terminal structure to the river
downstream.
This chapter concentrates on the hydraulics of the different elements that may be
incorporated in a design and some of the problems commonly encountered in their
design or operation.

4.2.1 The Overflow Spillway

Also known as an “ogee” spillway because of the reverse-curved shape of the entire
structure from the crest to the terminal structure, this is the most common type of
spillway and is widely used in gravity, arch and buttress dams, as well as for small
4.2 Spillway Structures 109

Fig. 4.3 A low-head overflow spillway

weirs and control structures in conveyance channels (Fig. 4.3). Embankment dams
often incorporate concrete gravity sections to accommodate this type of spillway.
Hydraulically, it is a very simple structure, having only one possible control
location, which is at the crest. Design requires (i) specification of the crest shape,
(ii) determination of the crest level and length and (iii) determination of the flow
conditions over the crest, on the spillway face, and at the entry into a terminal
structure or stilling basin.
The shape of the crest is usually assumed to be ideally the same as the profile of
the underside of the nappe of a sharp-crested weir, as shown in Fig. 4.4. This
ensures approximately zero (atmospheric) pressure on the surface.
The shape of the profile depends on the upstream head, the inclination of the
upstream face and the height of the crest above the bed of the approach channel.
The desired shape for a particular set of conditions could be determined by con-
structing a flow net or performing a numerical potential flow analysis, but for
practical purposes various procedures for defining an appropriate shape have been
published. The curves downstream and upstream of the crest are usually described
separately. Both the United States Bureau of Reclamation (1973) and the U S Army
Corps of Engineers (1995) describe the downstream profile using the power
function

xn ¼ KHdn1 y ð4:3Þ

in which x and y are the coordinates (in feet) as indicated in Fig. 4.5, Hd is the
design head (in feet), n is a variable usually assumed to be 1.85 and K is a
110 4 Spillways

Fig. 4.4 Overflow spillway energy


crest profile
H/ H nappe

0.11H/

Fig. 4.5 Definition sketch reservoir level


for overflow spillway
geometry Hd Eqn (4.4)

Eqn (4.3)
P
y

coefficient which depends on the approach depth below the crest, P and Hd. For
P/Hd > about 1.5, K has a constant value of 2.0; values for low-head structures with
smaller P/Hd may be obtained from the U S Army Corps of Engineers (1995).
For the upstream segment of the profile, the United States Bureau of
Reclamation (1973) proposes a single or multiple circular arcs to connect tangen-
tially to the downstream profile at the crest. The Bureau also provides a standard
profile comprising multiple circular arcs with different radii for the entire profile
including the upstream and downstream segments; this is recommended for small
spillways where P exceeds one half of the maximum anticipated head—otherwise,
Eq. (4.3) should be used for the downstream segment.
The U S Army Corps of Engineers (1995) have adopted a procedure proposed by
Murphy (1973) for the upstream segment. This uses an elliptical shape described by
the dimensionless relationship

x2 ðB  yÞ2
þ ¼1 ð4:4Þ
A2 B2

for −x < −A and y < B, in which A and B are the half ellipse axes in the horizontal
and vertical directions, respectively. For P/Hd > about 2.0, A = 0.28Hd and
B = 0.164Hd; values for smaller P/Hd can be obtained from U S Army Corps of
Engineers (1995). Both the upstream and downstream segments are joined
4.2 Spillway Structures 111

tangentially to the respective face surfaces to conform with the dam wall geometry.
Although the hydraulic behaviour is quite sensitive to the upstream segment
geometry, Murphy (1973) found that little accuracy is lost with any slope on the
upstream face, so long as it joins the curved segment tangentially; the upstream face
is commonly vertical for concrete gravity dams.
For slender dam sections, such as in arch dams, the required crest profile may not
fit within the section as designed structurally. The profile can then be extended to
form a corbel on the upstream face, as shown in Fig. 4.6a, or the profile can be
made to bulge on the downstream face, as shown in Fig. 4.6b, although this
increases the possibility of cavitation. Another alternative would be to incorporate
the profile in the design of a splitter-type energy dissipator, as shown in Fig. 4.6c.
As indicated by Eqs. (4.3) and (4.4) in Fig. 4.5, the spillway profile depends on
the design head. The profile designed will therefore result in atmospheric surface
pressures only for the design head specified. Lesser heads will result in positive
pressures while greater heads will cause negative pressures.
Determination of the level and length of the spillway crest requires application of
a relationship between the discharge over the crest and the upstream water level.
The design discharge will be a known value, having been previously established
through a hydrological analysis, and the spillway size needs to be designed to
accommodate this.
At the design head, the nappe will be the same shape as for flow over the
corresponding sharp-crested weir and the pressures on the spillway surface will be
close to atmospheric. The equation for discharge (Q) is therefore similar to that for a
sharp-crested weir (a derivation is provided in Chap. 7) and usually applied in the
form

Q ¼ CLH 3=2 ð4:5Þ

in which C is an empirical discharge coefficient, L is the length of the crest and H is


the total head above the crest which can usually be assumed to coincide with the
reservoir water level. Application of Eq. (4.5) requires a value for C, which
depends on
– the approach flow depth,
– the difference between the actual head and the value for which the crest shape
was designed,

(a) (b) (c)

Fig. 4.6 Modified crest profiles


112 4 Spillways

2.2

2.1

2.0

C
1.9
(m1/2/s)

1.8

1.7

1.6
0.0 0.5 1.0 1.5 2.0 2.5 3.0
P/H

Fig. 4.7 Discharge coefficients for vertical-faced ogee spillways [adapted from United States
Bureau of Reclamation (1973)]

– the slope of the upstream face of the structure and


– the downstream flow conditions if the spillway is not high.
Values of C and their variations with the above influences have been determined
empirically and presented by the agencies that have provided specifications for the
crest profile. The value used should be consistent with the profile shape for which
they were determined, but as the different profile formulations all approximate the
same shape of the underside of the nappe over a sharp-crested weir, values from any
of the different sources can be accepted for practical purposes. The effect of
approach depth on coefficient values for vertical-faced ogee crests is shown in
Fig. 4.7.
The value of C depends on the head on the spillway. If the head is less than the
design head, then positive pressures develop on the spillway and the value is
reduced, decreasing the efficiency of the structure. If the head is greater than the
design head, then pressures are negative and C is greater. It is therefore sensible to
design the shape of the spillway for a head which is less than anticipated, to take
advantage of the enhanced discharge coefficient. Information provided by the
United States Bureau of Reclamation (1973) indicates that C is increased by about
6% if the actual head exceeds the design head by 50% and decreased by about 4% if
the actual head is 50% less than the design head.
There is a danger in operating a spillway at higher than design heads, however,
as separation may occur if the pressure reduction is too great. This can happen
intermittently leading to vibrations which can cause noise and structural damage. In
extreme cases, cavitation can occur which can result in severe pitting and erosion of
the surface. The occurrence and prevention of cavitation are discussed further in
Sect. 4.3. Chadwick and Morfett (1986) report research results indicating that
separation will not occur until the actual head is about 3 times the design head,
although Novak et al. (2001) recommend a maximum of about 1.65 times the
4.2 Spillway Structures 113

design head. The U S Army Corps of Engineers (1995) suggest designing the
spillway to keep the average pressure on the crest above −4.6 m, and provide
cavitation safety curves relating actual and design heads to crest pressures for crests
with and without piers. Cassidy (1970) presented a rational method for ‘underde-
signing’ a spillway crest, which considers the minimum pressure occurring on the
crest during possible flow at heads higher than the design value.
The United States Bureau of Reclamation (1973) also presents additional
information to account for the effects on C of upstream face slopes and downstream
influences.
Reese and Maynord (1987) tested the profile defined by Eqs. (4.3) and (4.4) and
provided further quantifications of C for varying relative heads, upstream slopes
and P/H  2.0. Further details are provided by U S Corps of Engineers (1995),
including values for milder upstream face slopes and low approach depths.
It should be noted that C is not dimensionless and the values given by the United
States Bureau of Reclamation and U S Army Corps of Engineers publications are
generally for use with foot-second units. They must be converted when using SI
units.
It is common to regulate the flow over a spillway by installing gates on the crest
(Fig. 4.8). Gates are also installed to increase reservoir capacity. Various types of
gates are used, including radial (tainter), drum, vertical lift and flap types. Radial
and vertical gates (Fig. 4.9) are the most commonly used, and are described in some
detail by U S Army Corps of Engineers (1995).
Crest gates change the flow conditions over the spillway, which becomes more
an orifice type of flow than a free surface flow. This affects the trajectory of the
issuing jet, and hence the ideal spillway surface profile, and also the discharge–head
relationship. The trajectory for a vertical orifice can be expressed (in foot-second or
SI units) by the parabolic equation (United States Bureau of Reclamation 1973)

Fig. 4.8 Radial gates on a side-channel overflow crest


114 4 Spillways

H1
H H
β H2
Go Go

x/Hd x/Hd

(a) Radial Gate (b) Vertical Lift Gate

Fig. 4.9 Common gate types for controlled spillway crests

x2
y ¼ ð4:6Þ
4H

where H is the head on the centre of the opening, and for an orifice inclined at h to
the vertical can be described by

x2
y ¼ x tan h þ ð4:7Þ
4H cos2 h

A profile designed according to Eq. (4.7) will be wider than the usual one, such
as described by Eq. (4.3). Gates are often installed on existing structures and the
negative pressures that would occur on the downstream segment of the usual profile
can be reduced by locating the sills of the gates downstream of the crest (by x/
Hd = 0.2, as recommended by Novak et al. (2001) and shown in Fig. 4.9).
Different gate configurations obviously define different control relationships. For
the radial gate, the U S Army Corps of Engineers (1995) suggest using the high
head orifice equation

Q ¼ CWb Go ð2gH Þ1=2 ð4:8Þ

in which Wb is the gate width, Go is the gate opening (the minimum distance from
the gate lip to the spillway surface) and H is the distance from the water surface to
the centre of the gate opening (see Fig. 4.9). The discharge coefficient, C, depends
on the angle, b, between the tangents to the gate lip and the spillway surface at the
point defining Go. The U S Army Corps of Engineers (1995) provide a graph for C,
giving values for x/Hd between 0.1 and 0.3 varying from about 0.67 for b = 50° to
about 0.73 for b = 100°.
4.2 Spillway Structures 115

For a vertical lift gate, the U S Army Corps of Engineers (1995) propose the
relationship

3=2 3=2
H2  H1
Q ¼ Qf ð4:9Þ
H 3=2

in which Qf is the free flow (ungated) discharge [according to Eq. (4.5) and H1 and
H2 are the depths below the water surface of the bottom of the gate and its seating
level on the spillway surface (see Fig. 4.9)].
The length, L, in Eq. (4.5) does not account for the presence of piers (for gate
mountings, for example) and contractions of flow at the abutments. These may be
accounted for by reducing the net crest length according to (United States Bureau of
Reclamation 1973)
=  
L ¼ L  2 NKp þ Ka H ð4:10Þ

in which L/ is the effective length of the crest, N is the number of piers, Kp is a pier
contraction coefficient and Ka is an abutment contraction coefficient. The contrac-
tion coefficients depend on the geometry and H. Kp varies from 0 for pointed nose
piers to 0.020 for square-nosed piers with rounded corners. Ka varies from 0 for
angled, rounded abutments to 0.20 for square abutments. Further details are given
by the United States Bureau of Reclamation (1973), the U S Army Corps of
Engineers (1995) and Hager (1988).
Application of the control relationships described above enables an appropriate
spillway length and crest level to be determined. Detailed design also requires
prediction of flow characteristics over the crest, down the spillway face and at the
entry to the terminal structure at the end of the spillway face. The U S Army Corps
of Engineers (1995) provide guidance on the conditions favourable for cavitation on
the crest surface and for defining the upper nappe profile of the flow over the crest.
Knowledge of this profile is valuable for designing the geometry of the side walls
adjacent to the crest and gate mounting structures.
The flow conditions at the entry to the terminal structure are determined by the
flow down the spillway face. This surface is usually very steep (about 60°), making
the flow too complex for the usual open channel theory and gradually varied flow
techniques to be used reliably. Over the crest, the flow is rapidly varied and friction
has only a small influence until the boundary layer has developed fully (Fig. 4.10).
Once the boundary layer is fully developed, air entrainment is likely and the flow
cannot be analysed in the usual way. (Some aspects of aerated flow are discussed in
Sect. 4.3.) Bradley and Peterka (1957) presented the results of investigations of this
problem undertaken by the United States Bureau of Reclamation. Figure 4.11 (also
presented in Henderson 1966) was developed from their prototype tests on the
Shasta and Grand Coulee Dams.
Figure 4.11 gives the correction to be applied to the “theoretical” velocity (Vth)
to obtain the actual velocity (Va) at the foot of the spillway as a function of the
116 4 Spillways

Fig. 4.10 Boundary layer


development on spillway H
limit of boundary layer

Fig. 4.11 Velocity at the foot 180


of a spillway (adapted from H (m)
Bradley and Peterka 1957) 0.76 1.5 2.3 3.0 4.6 6.1 7.6 9.1 12.2
160

140

120

100
h
(m) 80

60

40 H

h
20

0 0.2 0.4 0.6 0.8 1.0


Va / Vth

spillway height and the head on the crest. This theoretical value is calculated by
equating the energy values at the crest and toe, assuming no losses, to give
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 
H
Vth ¼ 2g h  ð4:11Þ
2

in which g is gravitational acceleration and h is the height of the reservoir level


above the base of the spillway. The correction given by Fig. 4.11 is considered by
the authors to be sufficiently accurate for preliminary design for spillway slopes in
the range 0.8H–0.6H:1V. Note that for a high head on a low spillway, the
4.2 Spillway Structures 117

theoretical velocity is quite accurate. In this case, the boundary layer is not fully
developed and friction is not controlling the flow. For a low head on a high
spillway, friction will control the flow and substantial correction is required. Model
testing is not reliable for this situation because the effects of aeration are not reliably
reproduced.

4.2.2 Labyrinth and Piano Key Weirs

The discharge capacity of an overflow spillway is limited by the allowable upstream


water level and the length of the crest. The labyrinth design provides an increased
crest length by folding the crest in plan to form a series of cells (Fig. 4.12). This
form provides similar discharge to a straight crest, but at a lower head, allowing
greater reservoir storage capacity without compromising flood safety. Labyrinth
weirs are often used for increasing reservoir storage of existing dams, in preference
to installing control gates, and simultaneously increasing flood discharge capacity.
Geometries vary widely, with the plan shape of cells commonly triangular, trape-
zoidal (a > 0) or rectangular (a = 0).
The discharge is usually related to the upstream energy level by the standard
weir equation (see Chap. 7),

2 pffiffiffiffiffi 3=2
Q ¼ Cd L 2gH ð4:12Þ
3

in which Cd is the (dimensionless) discharge coefficient and L is the total crest


length. The value of Cd depends on H and various factors describing the weir
geometry, including the crest height (P), the width and shape of the crest, and the
plan geometry defined by the side wall angle (a) and the apex and exit widths.
Various formulations for Cd in terms of these geometric parameters have been

Flow
A
flow path 2

α H

P
flow path 1

A
1 2 1 2
(a) Plan (b) Section A-A

Fig. 4.12 Labyrinth weir geometry


118 4 Spillways

proposed (e.g. by Khode et al. 2012 and Crookston and Tullis 2013a). The value
peaks at a low head and then declines rapidly as interference of the nappes from
opposite sides and submergence increases, as shown, for example, (not for design)
in Fig. 4.13. The value of Cd for a straight weir peaks at around 0.80 for
H/P = 0.30–0.50 and declines to around 0.70 for H/P = 1.0. Cd for a labyrinth weir
is therefore always less than for a straight weir, but the reduction in efficiency is
usually more than compensated for by the increased crest length.
The complex flow patterns produced in labyrinth weirs can cause undesirable
conditions, such as nappe vibration with associated noise, surging flow and fluc-
tuating pressure forces on the walls. Crookston and Tullis (2013b) have described
these phenomena, identified the hydraulic conditions under which they occur and
proposed some measures for countering their effects.
The piano key weir is a modification of a rectangular labyrinth weir, with the cells
having sloping floors and extending beyond the base on which they are founded
(Figs. 4.14 and 4.15). This geometry allows an even longer crest than a labyrinth
layout with the same ‘footprint’; this makes it particularly useful for raising the
storage level on existing dam structures. The overhanging structure reduces the
construction material considerably, and improves the hydraulic efficiency slightly.
The discharge can be described by the same equation as for labyrinth weirs
(Eq. (4.12)), with Cd now depending also on the overhang geometry and bed slope.
As for labyrinth weirs, the hydraulic efficiency is greatest at low heads, and declines
significantly as the head increases (Fig. 4.13). The hydraulic performance of piano
key weirs has been described by Anderson and Tullis (2012), Leite Ribeiro et al.
(2012), Machiels et al. (2014), amongst others.

Fig. 4.13 Typical Cd 0.55


variation for labyrinth and
piano key weirs (adapted from
Anderson and Tullis 2012) 0.50
Piano Key
0.45
Labyrinth

0.40
Cd
0.35

0.30

0.25

0.20
0.0 0.2 0.4 0.6 0.8 1.0
H/P
4.2 Spillway Structures 119

Fig. 4.14 Piano key weir (photograph by F. Denys)

Flow flow path 2


A H

flow path 1

A
1 2 1 2
(a) Plan (b) Section A-A

Fig. 4.15 Piano key weir geometry

4.2.3 The Side-Channel Spillway

A side channel spillway is commonly used where there is insufficient space to


accommodate the required spillway length in line with the dam axis, such as where
the abutments are steep. The spillway crest is usually placed along one abutment
and perpendicular to the dam axis, as shown in Fig. 4.16.
The crest is usually a conventional overflow spillway, and in some cases, inflow
is from both sides of the channel. An L-shaped configuration, with water entering
the top end of the side channel in the main flow direction is also commonly used
120 4 Spillways

(b) Side-channel cross section

(a) Plan
(c) Side-channel long section

Fig. 4.16 Side-channel spillway

(Fig. 4.17); this assists in changing the direction of the lateral flow and reducing
surface waves. Water flows over the control section into a channel or trough at right
angles to the main inflow, usually leading to a chute. The side channel will often
have a control structure (such as a weir) at its downstream end to ensure subcritical
flow along its length. This enhances energy dissipation through plunging of the

Fig. 4.17 An L-shaped side-channel spillway


4.2 Spillway Structures 121

entry flow and prevents transverse wave action. It also facilitates entry into a chute
or tunnel by allowing sharper changes in flow direction.
The channel should be designed so that the inflow is unsubmerged at the design
discharge, or the spillway discharge coefficient will be reduced and the analysis will
be much more complex. The main design problem is therefore to select the depth,
width and slope of the side channel to ensure this condition. The size of the side
channel can be kept small by making it steep enough to induce supercritical flow
and not including a control structure at the downstream end. Supercritical flow is
undesirable, however, because it has insufficient depth to absorb and damp the
incoming flow and the side-channel flow will be swept towards the opposite wall
resulting in violent wave action. It is good practice, therefore, to include a control
structure to force deep, subcritical flow along the length of the side channel. Sizing
the channel requires an analysis of the water surface profile.
The flow along the side channel is gradually varied, but the conventional
analysis techniques are inadequate because they do not allow for a varying dis-
charge along the profile. Also, the standard gradually varied flow equation is based
on the energy equation and assumes that the friction gradient accounts for all energy
losses. In the side channel, there is an appreciable energy loss associated with the
turbulent mixing of the water plunging in from the side. This additional loss cannot
easily be estimated, and without its inclusion, the energy equation cannot be used.
The incoming flow has no momentum in the flow direction, however, and therefore
does not affect the total momentum in the side channel. The flow can therefore be
analysed on the basis of momentum conservation.
If the bed is horizontal and the effect of friction can be considered small com-
pared with the effect of the varying discharge, then no forces act on the water in the
flow direction and the momentum function is constant. The water profile can then
be computed simply by establishing the momentum function at the downstream
control and solving for the flow depth at any position with the appropriate
discharge.
If the channel has a slope (which introduces a weight component in the flow
direction) and friction is significant, then a more complete solution is required. This
can be obtained by considering the momentum of an element in the profile with a
length Dx, as shown in Fig. 4.18.
The forces acting on this element in the flow direction are
– the component of its weight in the x direction,

cADx sin h

in which A is the cross-sectional area and h is the angle of the bed to horizontal,
– the shear force on the boundary,

so PDx
122 4 Spillways

ΔQ

Δx

F1

x
F2
1 θ

γ A Δx 2

Fig. 4.18 Side channel with lateral inflow

in which so is the boundary shear stress = cRSf, R is the hydraulic radius (=A/P),
P is the wetted perimeter and Sf is the energy gradient and
– pressure forces (F) at sections 1 and 2.
The change in momentum between sections 1 and 2 is equal to the sum of forces
acting on the element in the flow direction. In terms of the momentum function
(which incorporates the pressure forces), therefore

cADx sin h cRSf PDx


DM ¼ 
c c

For small slopes, sin h  tan h ¼ So , and R = A/P. Dividing through by Dx and
replacing finite differences by differentials give

dM  
¼ A So  Sf ð4:13Þ
dx

From the definition of the momentum function

Q2
M¼ þ Az
gA

in which z is the depth of the cross-section centroid below the water surface. M can
be differentiated with respect to x to obtain

dM 2Q dQ Q2 dA dy
¼  þA ð4:14Þ
dx gA dx gA2 dx dx
4.2 Spillway Structures 123

The right-hand sides of Eqs. (4.13) and (4.14) can be equated to eliminate dM/
dx. The term dA/dx can be expressed as (dA/dy)(dy/dx) and dA/dy = B, the surface
width, because incremental changes at the surface are being considered. Making
this substitution and rearranging terms give

dy So  Sf  gA2 dx
2Q dQ
¼ ð4:15Þ
dx 1  Fr2

in which Fr is the Froude number. Note that if there is no lateral inflow, then dQ/
dx = 0 and Eq. (4.15) reduces to the familiar ordinary gradually varied flow
equation. For a side-channel spillway, the inflow over the spillway section is
usually uniformly distributed and so the lateral inflow term dQ/dx has a constant
value. The total discharge in the side channel, Q, however, varies with the distance,
x. (Some sources include a in the third term of the numerator and in Fr.).
Equation (4.15) can be solved in finite difference form by specifying sections
along the side channel, separated by known Dx distances. Starting at the control
section, Dy can then be calculated progressively from section to section. The flow
depth is required for calculating A, Sf and Fr, which should strictly be calculated for
each Dx as the mean of the values at the bounding sections. This would require an
iterative solution using an initial estimate for the flow depth at the new section.
However, if Dx is kept small, the known flow depth can be assumed to apply
through Dx, enabling direct solution. The procedure is then to calculate dy/dx from
Eq. (4.15) using the current section values yi and Qi. At the next section,
yi+1 = yi + dy/dx  Dx and Qi+1 = Qi + dQ/dx  Dx, and these then become the
current values for the next calculation.
If the side channel has a smooth transition to a chute, with no control structure, at
its end critical flow must occur at some position somewhere along its length. The
full water surface profile must then be computed from the critical section in the
upstream direction for the subcritical portion and downstream for the supercritical
portion. Henderson (1966) provides the following formula for the distance, x, of the
critical section from the head of the side channel:

8Q2x
x¼  3 ð4:16Þ
gB2 So  CgP2 B

in which Qx is the inflowing discharge per unit length of channel (i.e. dQ/dx) and
C is the Chézy resistance coefficient (which can be substituted by the Manning
coefficient, n, if desired, with C = R1/6/n). Equation (4.16) must be solved itera-
tively, i.e. an initial value of x is assumed, which is used to calculate the discharge
at that section (=Qx  x); the critical depth can then be calculated from Fr2 = 1 and
used to calculate P, B and R; x is then calculated from Eq. (4.16) and compared
with the initially assumed value; the initial estimate can then be updated and the
procedure repeated until satisfactory convergence is obtained. If the calculated
124 4 Spillways

value of x is greater than the length of the channel, then subcritical flow will persist
up to the entrance to the chute. If there is a control structure at the end of the
channel, it may override the critical section and produce subcritical flow over the
whole channel length.
As mentioned before, flow conditions in the side channel can be improved by
introducing a longitudinal flow component to the upstream end of the side channel
by realigning the crest to an L-shape, with a portion normal to the channel flow
direction and a portion parallel to it. The water surface profile analyses outlined
above need modification for this situation and an appropriate design procedure has
been presented by Knight (1989).
Example 4.1
A 75 m long ogee spillway discharges to an 8.0 m wide, rectangular side channel with a
slope of 0.15 and Manning n of 0.013. The channel leads through a smooth transition to
a steep chute at the end of the spillway crest. Compute the water surface profile along the
channel when the discharge over the spillway is 3.5 m3/s/m.
Solution
Determine if a critical flow section occurs along the channel using Eq. (4.16), i.e.

8Q2x
x¼  3
g B2 So  CgP
2B

with the substitution C = R1/6/n,

8Q2x
x¼  3
2
g B2 So  RgPn
1=3 B

Solving by trial, for x = 48.7 m: Q = Qx  x = 3.5  48.7 = 170.4 m3/s

for which

 1=3
Q2
yc ¼
g B2
 1=3
170:42
¼ ¼ 3:59 m
9:8  8:02

A ¼ Byc ¼ 8:0  3:59 ¼ 28:7 m2


P ¼ B þ 2yc ¼ 8:0 þ 2  3:59 ¼ 15:2 m
A 28:7
R¼ ¼ ¼ 1:89 m
P 15:2

Therefore

8  3:52
x¼  3 ¼ 48:7 m
9:8  8:02 0:15  9:815:20:013
2
1:891=3 8:0
4.2 Spillway Structures 125

x < L = 75 m, so a critical flow section occurs within the side-channel length. Profile
computation must therefore begin at this section and work upstream for the subcritical
portion and downstream for the supercritical portion. The profiles are computed by solving
Eq. (4.15) in finite difference form. Because dy/dx cannot be calculated at the critical
section, the first step is done by iteration of Dy, using variables averaged over the first
length increment and subsequent calculations are done directly using variable values at the
nominated sections.
Calculations for the subcritical and supercritical profiles are shown below. In the tables

x distance along the channel,


Dx distance between nominated sections,
Dy change in flow depth between sections, iterated for first step, then = Dy/Dx  Dx,
y flow depth, = y at previous section +Dy,
A flow area, = B  y,
Q discharge at section, = 3.5  x,
B channel width,
Fr2 Froude number2, = Q2B/gA3,
P wetted perimeter, = B + 2  y,
R hydraulic radius, = A/P,
V velocity, = Q/A,
Sf energy gradient, = V2n2/R4/3,
Dy/Dx water surface gradient, by Eq. (4.15).

Calculations for the subcritical profile, upstream from critical section:


x Δx Δy (m) y (m) A (m2) Q B Fr2 P R (m) V (m/s) Sf Δy/ Δy (m)
(m) (m) (m3/s) (m) (m) Δx
48.7 3.591 28.72 170.45 8.0 1.000 15.18 1.89 5.93 0.002543
−3.7 −0.120 28.24 163.98 8.0 0.973 1.88 5.80 0.002462 −0.120
45 3.471 27.76 157.50 8.0 0.945 14.94 1.86 5.67 0.00238 0.033
40 −5.0 −0.167 3.304 26.43 140.00 8.0 0.866 14.61 1.81 5.30 0.002151 0.036
35 −5.0 −0.180 3.124 24.99 122.50 8.0 0.784 14.25 1.75 4.90 0.00192 0.038
30 −5.0 −0.188 2.936 23.49 105.00 8.0 0.694 13.87 1.69 4.47 0.001674 0.041
25 −5.0 −0.204 2.732 21.85 87.50 8.0 0.598 13.46 1.62 4.00 0.00142 0.044
20 −5.0 −0.222 2.510 20.08 70.00 8.0 0.494 13.02 1.54 3.49 0.001153 0.049
15 −5.0 −0.246 2.263 18.11 52.50 8.0 0.379 12.53 1.45 2.90 0.000869 0.056
10 −5.0 −0.281 1.983 15.86 35.00 8.0 0.250 11.97 1.33 2.21 0.000565 0.067
5 −5.0 −0.335 1.648 13.19 17.50 8.0 0.109 11.30 1.17 1.33 0.000242 0.087
0 −5.0 −0.437 1.211
126 4 Spillways

Calculations for the subcritical profile, downstream from the critical section:
x Δx Δy y A (m2) Q B (m) Fr2 P (m) R (m) V Sf Δy/ Δy
(m) (m) (m) (m) (m3/s) (m/s) Δx
48.7 3.591 28.72 170.39 8.0 0.999 15.18 1.89 5.93 0.002541
1.3 0.040 28.88 172.70 8.0 1.009 1.90 5.98 0.002571 0.040
50 3.630 29.04 175.00 8.0 1.020 15.26 1.90 6.03 0.002602 0.033
55 5.0 0.166 3.797 30.37 192.50 8.0 1.079 15.59 1.95 6.34 0.002791 0.022
60 5.0 0.108 3.904 31.23 210.00 8.0 1.180 15.81 1.98 6.72 0.003082 0.037
65 5.0 0.186 4.090 32.72 227.50 8.0 1.205 16.18 2.02 6.95 0.003196 0.024
70 5.0 0.118 4.208 33.66 245.00 8.0 1.283 16.42 2.05 7.28 0.003436 0.027
75 5.0 0.136 4.344

Repeating the calculations using iteration at each step resulted in a difference of 0.12 m
at x = 0 and no difference at x = 75 m for the nominated x increments.
The calculated water surface profile is shown below:

14
12 Channel bed
Elevation (m)

10 Water surface
8
6
4
2
critical section
0
0 20 40 60 80
Distance (m)

4.2.4 The Side Weir

A side weir is a weir along one side of a channel with its crest parallel to the main
flow direction in the channel. Side weirs are used to divert flow from the channel
into a lateral supply canal or wasteway. The discharge in the main channel will
decrease along the length of the side weir by the amount diverted. The design
requirements include the crest level (to determine the discharge at which diversion
begins), the crest length (to determine the magnitude of discharge diverted) and the
water surface profile along the length of the side weir (to determine the channel side
wall heights). Determining the diverted discharge and the side wall heights both
require computation of the water surface profile over the crest length.
The flow over the length of the crest is gradually varied, but cannot be described
by the normal gradually varied flow equation [Eq. (1.29)] because this implicitly
4.2 Spillway Structures 127

assumes a constant discharge along the profile. The momentum approach followed
for gradually varied flow with lateral inflow, as used for the side-channel spillway
[Eq. (4.15)], is also not applicable because it is based on the assumption that the
incoming discharge does not change the momentum of the flow in the side channel
because its initial direction is normal to the side-channel flow. In the lateral outflow
case, the diverted flow does carry momentum with it, so the longitudinal change in
momentum cannot be assumed to result only from the forces applied to the flow and
the varying Q. In the lateral inflow case, an energy analysis was not possible
because of the considerable and unquantifiable energy loss associated with the
turbulent mixing of the incoming flow. For the lateral outflow case, however, there
is no energy loss associated with the diverted flow in Bernoulli terms, because this
represents the energy per unit weight of water and is unaffected by changing
discharge.
The gradually varied flow with lateral outflow can therefore be described using
an energy analysis similar to that for normal gradually varied flow, but also
accounting for the variation of discharge with distance. As for the constant dis-
charge case, the change in specific energy along the channel results from the change
in bed elevation and the loss through friction,

dE
¼ So  Sf ð1:55Þ
dx

in which E is specific energy, x is the longitudinal direction, So is the bed slope and
Sf is the energy gradient. Therefore
 
d Q2
yþ ¼ So  Sf
dx 2gA2

in which y is the flow depth and A is the cross-sectional flow area. Because both
Q and A vary with x,

dy Q dQ Q2 dA dy
þ 2  ¼ So  Sf
dx gA dx gA3 dy dx

and because dA/dy = B, the surface width, and Q2B/gA3 = Fr2,

dy So  Sf  gA2 dx
Q dQ
¼ ð4:17Þ
dx 1  Fr2

which is similar to the equation for lateral inflow, except for the coefficient of the
third term in the numerator.
According to Eq. (4.17), the water surface will rise in the flow direction for
subcritical flow and drop in the flow direction for supercritical flow. It is also
possible to have a change from supercritical to subcritical flow through a hydraulic
jump within the weir crest length.
128 4 Spillways

Equation (4.17) can be solved numerically to define the water surface profile.
The solution requires evaluation of the term dQ/dx at each computational step. This
can be evaluated from the standard weir equation

2 pffiffiffiffiffi
q ¼ Cd 2gh3=2 ð7:2Þ
3

in which q is the discharge per unit crest length (=dQ/dx), Cd is an empirical


discharge coefficient and h is the height of the energy line above the crest. The
discharge coefficient is different from that for a normal weir because of the more
complicated geometry and flow pattern. The value has been found to depend on the
angle of the flow over the crest and the Froude number of the flow in the main
channel. Various formulations have been derived from laboratory measurements,
Borghei et al. (1999), for example, used laboratory data for Fr < 0.80 to propose
for sharp-crested side weirs with subcritical flow (using h = y – w)

w L
Cd ¼ 0:7  0:48Fr  0:3 þ 0:06 ð4:18Þ
y B

in which w is the crest height, y is the flow depth, L is the crest length and B is the
surface width in the main channel. Both Fr and y vary along the profile, and so Cd
needs to be calculated at each computational step.
Example 4.2
A canal with a 3.5 m wide rectangular cross section, a slope of 0.00080 and a Manning
n of 0.013 incorporates a 12.0 m long sharp-crested side weir with its crest 1.00 m
above the bed. Compute the water surface profile along the length of the weir if the
discharge in the channel downstream 10.0 m3/s.
Solution
The flow depth at the downstream end of the weir is the uniform flow depth in the canal.
Calculate this from the Manning and continuity equations,

A
Q ¼ R2=3 So1=2
n
A ¼ Wyo
n ¼ 0:013 for concrete
A Wyo
R¼ ¼
P W þ 2yo
So ¼ 0:00080

Therefore
 2=3
Wyo Wyo
Q¼ S1=2
n W þ 2yo o
4.2 Spillway Structures 129

i.e.
 2=3
3:5yo 3:5yo
10 ¼ 0:000801=2
0:013 3:5 þ 2yo

from which

yo ¼ 1:51 m

Compute the profile from this depth in an upstream direction over the length of the side
weir using Eq. (4.17) in finite difference form. At each step, the lateral outflow (=dQ/dx)
is calculated using Eq. (7.2) with Cd given by Eq. (4.18). Calculations are shown below.
In the table

x distance along the channel,


y flow depth, = y at previous section + Dy/Dx(x – x at previous section),
A flow area, = W  y,
Q discharge at section, = Q at previous section + dQ/dx  Dx,
V velocity = Q/A,
R hydraulic radius = A/(W + 2y),
Sf energy gradient = V2n2/R4/3,
Fr2 Froude number = Q2B/gA3,
Cd discharge coefficient, Eq. (4.18)
dQ/dx weir discharge per unit length, Eq. (7.2),
Dy/Dx water surface gradient, by Eq. (4.15),
z elevation of bed relative to level at downstream end of weir,
h water surface elevation = z + y,
Crest elevation of weir crest.

x y (m) A Q V R (m) Sf Fr2 Cd dQ/dx Δy/Δx z (m) h (m) Crest


(m) (m2) (m3/s) (m/s) (m)
0 1.510 5.29 10.00 1.89 0.81 0.0008 0.242 0.471 −0.507 0.0244 0 1.510 1
−1 1.486 5.20 10.51 2.02 0.80 0.000923 0.280 0.450 −0.450 0.0246 0.0008 1.487 1.0008
−2 1.461 5.11 10.96 2.14 0.80 0.001051 0.320 0.429 −0.397 0.0246 0.0016 1.463 1.0016
−3 1.437 5.03 11.35 2.26 0.79 0.001181 0.362 0.408 −0.348 0.0244 0.0024 1.439 1.0024
−4 1.412 4.94 11.70 2.37 0.78 0.001315 0.405 0.388 −0.303 0.0240 0.0032 1.416 1.0032
−5 1.388 4.86 12.00 2.47 0.77 0.001451 0.448 0.368 −0.263 0.0236 0.004 1.392 1.004
−6 1.365 4.78 12.27 2.57 0.77 0.001588 0.493 0.349 −0.227 0.0230 0.0048 1.370 1.0048
−7 1.342 4.70 12.49 2.66 0.76 0.001726 0.538 0.330 −0.195 0.0224 0.0056 1.347 1.0056
−8 1.319 4.62 12.69 2.75 0.75 0.001865 0.584 0.312 −0.166 0.0217 0.0064 1.326 1.0064
−9 1.298 4.54 12.86 2.83 0.75 0.002003 0.630 0.294 −0.141 0.0209 0.0072 1.305 1.0072
−10 1.277 4.47 13.00 2.91 0.74 0.002142 0.676 0.276 −0.119 0.0202 0.008 1.285 1.008
−11 1.257 4.40 13.11 2.98 0.73 0.00228 0.722 0.259 −0.099 0.0194 0.0088 1.266 1.0088
−12 1.237 4.33 13.21 3.05 0.72 0.002417 0.768 0.243 −0.083 0.0187 0.0096 1.247 1.0096

The calculated water surface profile is shown below:


130 4 Spillways

1.6
1.4
Elevation (m) 1.2
1.0
0.8 Water surface
0.6
Channel bed
0.4
Weir crest
0.2
0.0
-12 -10 -8 -6 -4 -2 0
Distance (m)

4.2.5 Shaft (Morning Glory) Spillways

In a shaft spillway, water flows over a circular control crest (Fig. 4.19), through a
transition, into a vertical shaft which changes to a horizontal conduit leading to an
exit downstream of the dam. A typical arrangement is shown in Fig. 4.20.
Shaft spillways are frequently used where there is insufficient space for a con-
ventional spillway and the flood discharges are not extreme. They are particularly
common for embankment dams with small catchment areas. One advantage of this
type is that the horizontal conduit may be used for river diversion during
construction.

Fig. 4.19 A shaft spillway crest


4.2 Spillway Structures 131

Fig. 4.20 Typical shaft


spillway arrangement

The control relationship for a shaft spillway is complex (as shown in Fig. 4.21d)
because the position of control can change as the water level changes.
A different control relationship is associated with each possible position of
control. In accordance with control principles, the actual discharge at any reservoir
water level is the lowest predicted value for any possible control and the resulting
composite control relationship is indicated by the solid curve in Fig. 4.21d. At low
heads, the crest is unsubmerged, as shown in Fig. 4.21a. Control is then at the crest
and the discharge is related to head by the usual weir equation, i.e. Q / H 3/2,
represented by curve ab in Fig. 4.21d. At higher heads, the throat, or transition
between the crest profile and the conduit, may control. In this case, the crest is
submerged and the structure flows full up to the end of the throat section, as shown
in Fig. 4.21b. The discharge is now described by an orifice-type equation, i.e. Q /
H1/2
a , represented by curve cd in Fig. 4.21d. At still higher heads, the entire structure
may flow full (Fig. 4.21c) and the discharge may be controlled by the resistance of
the conduit or the level of the tailwater, i.e. Q / H1/2 t , represented by curve ef in
Fig. 4.21d. If the control is by the conduit resistance, the water level at the exit will
most likely be at the top of the conduit, although flow may go through the critical
condition at this point if the critical flow depth is less than the height of the conduit;
if the tailwater controls, then the water level will be above the top of the conduit, as
determined by conditions further downstream. For conduit control, a certain
increase in discharge will cause a much greater increase in head than for crest
control, and shaft spillways are generally designed to operate under crest control at
the design flood discharge. The relative positions of the individual curves in
Fig. 4.21d, and hence the form of the composite curve, are determined by the
control feature characteristics such as the diameters of the crest, shaft and horizontal
conduit. The effects of these design parameters will be apparent in the following
discussion of the individual control relationships.

Crest Discharge

The crest, as for the straight overfall spillway, is best designed to the shape of the
underside of the nappe of the flow over a corresponding sharp-crested weir.
Because of the convergence of the flow, this shape is slightly different from that for
a straight crest. The United States Bureau of Reclamation (1973) presents tables of
coordinates of the underside nappe profiles for different ratios of total crest height to
radius, based on experimental results.
132 4 Spillways

Ha

(a) Crest control (b) Throat control

Ht

(c) Conduit or tailwater control

conduit / tailwater
control
Q α Ht 1/2 d
Reservoir water level

throat control
Q α Ha 1/2
b
crest control
Q α H 3/2

c e
a

Discharge

(d) Rating curve

Fig. 4.21 Rating curve for a shaft spillway

For low heads, the discharge will be determined purely by the crest character-
istics. The vertical transition beyond the crest will flow partly full, with the flow
following the sides of the shaft. As the head increases, the annular nappe will
become thicker and will eventually converge into a solid, vertical jet. When this
happens the boil of water above the crotch where the nappe converges will begin to
4.2 Spillway Structures 133

affect the flow over the weir. The discharge coefficient will therefore be different
from that for a straight spillway. The United States Bureau of Reclamation (1973)
assumes that the orifice flow which occurs once the nappe has converged and the
crest is submerged can still be described by a weir formula, provided the discharge
coefficient is specified appropriately. In terms of the head measured from the apex
of the crest and the length defined by its outer periphery (i.e. the position of the
equivalent sharp-crested weir), the discharge relationship is given as

Q ¼ Co ð2pRs ÞHo3=2 ð4:19Þ

in which Rs is the radius of the outer periphery of the crest and the subscripts o indicate
the values at the condition for which the crest profile is designed. Because of the
effects of submergence and back pressure in converging flows, Co depends on both Ho
and Rs. For the design head with the corresponding crest profile, the United States
Bureau of Reclamation (1973) provides a graphical relationship between Co and Ho/
Rs determined by model testing. Free flow occurs for values of Ho/Rs up to about 0.45,
and the value of Co is similar to that for a conventional overflow weir in this range.
Partial submergence occurs as Ho/Rs increases above 0.45, becoming complete at Ho/
Rs about 1.0, and Co reduces very considerably. The United States Bureau of
Reclamation (1973) also presents a diagram for adjusting the discharge coefficient for
heads different from that for which the crest profile was designed.
Many shaft spillway designs incorporate radial guide vanes on the crest to inhibit
vortex action, as recommended by the United States Bureau of Reclamation (1973),
for example. Novak et al. (2001), however, maintain that vortex formation is
actually advantageous because the spiral flow induced along the shaft walls sub-
stantially reduces vibration and pressure fluctuations associated with a free-falling
jet at low (and hence frequent) discharges. They recommend installing guide vanes
below the crest to induce vortex formation.

Throat Discharge

The shape of the throat or transition between the overflow crest and the outflow
conduit should be designed to follow the shape of a vertical jet falling under
gravity. The jet will have an area A = pR2 and a velocity V = (2 gh)1/2, where R is
the jet radius and h is the height to the water surface in the reservoir. By setting
Q = AV and expressing this in terms of R, an equation for the shape of the throat
can be obtained, i.e.
sffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Q
R¼ pffiffiffiffiffiffiffiffi ð4:20Þ
p 2gh

The shape defined by Eq. (4.20) will give the minimum size to accommodate the
flow without restriction and without developing pressures on the side of the shaft.
134 4 Spillways

The actual profile will be wider than this for at least some distance to accommodate
the crest profile, and flow will be under pressure in this region. Free flow will occur
if the radius of the jet, as defined by Eq. (4.20), becomes less than the radius of the
shaft, and the location where this happens is the point of control for the throat. By
substituting the corresponding value for h, Eq. (4.20) can be used as the discharge
equation under conditions of throat control.

Conduit Discharge

If the entire structure is flowing full, the position of control will be at the outlet of
the conduit. The rating curve for this condition can be derived simply by applying
energy conservation between the crest and the outlet and accounting for friction
losses using a resistance equation, such as that of Manning or Darcy–Weisbach.
Entrance losses at the inlet, bend losses and expansion losses will usually be
negligible in comparison with the friction losses; if they are not considered to be
negligible, they can be accounted for using the approaches and results described for
siphon spillways and culverts. Conduit control is actually an undesirable flow
condition and should be avoided because the transition to full flow may be
accompanied by intermittent aeration and sealing, causing erratic discharge,
vibrations and surges. To avoid this, the United States Bureau of Reclamation
(1973) recommends designing the conduit so that it will not flow full beyond the
throat section. To allow for air bulking and surging, it is recommended that the
conduit should not flow more than 75% full (in area) at the downstream end at the
maximum discharge.

4.2.6 Siphon Spillways

A siphon spillway is a closed conduit which rises above the water level in the
reservoir and discharges at a lower level downstream (Figs. 4.22 and 4.23). This
enables a greater head to be applied to the flow than if a free surface existed, and the
discharge for a given upstream water level is consequently much greater. Two
control mechanisms apply over different ranges of water level, and the rating curve
is therefore a combination of the relationships describing these.
The operation of a siphon spillway can be examined by considering a gradually
rising water level. Discharge will begin when the water level rises above the crest
level and will be related to the water level by the usual free flow relationship, i.e.
Q / H 3/2. As the water level rises, the velocity will increase and some air will be
entrained by the flow and removed from the cavity formed by the conduit or barrel
(Fig. 4.24). Provided air cannot enter the barrel, all the air will ultimately be
removed, and the siphon will be primed and running full. Once this has happened,
the discharge will be related to the total head across the structure by Q / h1/2.
4.2 Spillway Structures 135

Fig. 4.22 A three-barrel siphon spillway

ro

rc h

Energy Line

Fig. 4.23 A simple siphon spillway arrangement

Fig. 4.24 Initiation of priming in a model siphon spillway


136 4 Spillways

fully primed
Q α h 1/2
Reservoir water level

air-partialized flow

crest control
Q α H 3/2

Discharge

Fig. 4.25 Typical siphon spillway rating curve

Priming usually occurs when the water level is about a third of the throat height and
causes a sudden and considerable increase in discharge, as shown in Fig. 4.25.
Although superficially similar, the composite rating curve for a siphon differs
from that for a shaft spillway because the crest control relationship does not persist
to its intersection with the fully primed conduit relationship; flow switches suddenly
between the two types as priming and de-priming occur. Priming is accompanied by
air entrainment and the flow is a mixture of air and water for a range of water levels.
Some siphons are designed to operate under this air-partialized condition to prevent
hunting and give a more gradually varying rating curve. In these designs, the inlet
lip is carefully located and shaped to admit the right amount of air.
For siphons operating under fully primed conditions, full conduit flow will
persist until the water level drops sufficiently for air to enter the barrel once more.
Because the lip of the inlet is usually located below the crest level to prevent air
entry by gulping or vortices, an air vent is installed at about the crest level to control
priming and de-priming, as shown in Fig. 4.26. If the outlet is not submerged, air is
prevented from entering the barrel from the downstream side by installing a
deflector in the barrel or giving it an adverse slope, as shown in Fig. 4.26, to create
a water curtain.
Siphon spillways are frequently used where it is required to pass a high dis-
charge within a small range of water levels, or where there is a severe limitation on
space. A siphon can be formed by installing a hood over an overflow or a shaft type
spillway crest, and the outlet can be set to the elevation required to provide the
appropriate head, irrespective of the height of the crest. A siphon can respond
rapidly and operate at full capacity if the upstream water level rises suddenly. This
is useful for forebays of turbine plants or water treatment works to discharge
incoming water to waste in the event of sudden closure. The abrupt priming
characteristic can be a disadvantage, however, if sudden spates downstream are
4.2 Spillway Structures 137

water
curtain
minimum minimum
level level

vent vent

water
curtain

(a) Deflector type (b) S-barrel type

Fig. 4.26 Siphon sealing and air venting arrangements

undesirable, but this can be remedied by using multiple siphon barrels set to prime
at different levels. As indicated by the control relationship, once a siphon is fully
primed, a significant increase in water level accompanies any appreciable increase
in discharge, effectively limiting the capacity. Although compact, siphons are dif-
ficult and expensive to construct and need to be structurally stronger than con-
ventional spillways to withstand vibration. A further disadvantage is their inability
to pass debris and their tendency to clog (also with ice in cold climates).
The hydraulic design problems for siphon spillways are to define the geometry,
establish the rating curve (particularly for the fully primed condition), ensure that
priming takes place at the required water level and ensure that throat pressures are
high enough to avoid cavitation.
Siphon spillways are not as common as the other types, and there appear to be no
general guidelines for geometric design. Because the design discharge will occur in
the primed condition, there is no need to shape the crest profile after a free jet,

he
hb1 + hf
hf
h
hb2 + hf
h o + hf
hexp
Energy Line

Fig. 4.27 Energy loss through a fully primed siphon spillway


138 4 Spillways

which allows greater freedom in the design. Each design must be carried out for its
particular conditions, and physical models are often used for testing and refining
initial designs.
For unprimed flow, the control relationship is similar to that for a conventional
overflow spillway. The discharge coefficient would not be the same as for a con-
ventional spillway unless the standard crest profile was used, and would probably
also be slightly lower because of energy loss in the approach to the crest. However,
the design discharge is not intended to occur in the unprimed condition and great
accuracy is generally not required for the lower range of flows.
For primed flow, the spillway can be analysed as a closed conduit controlled at
the outlet, and a discharge equation is derived from energy considerations. This
requires accounting for the friction and other energy losses between the upstream
and downstream flows (Fig. 4.27). Conceptually this is easy, as the entrance, exit
and bend losses can be expressed as appropriate coefficients multiplied by the throat
velocity head, and friction loss can be evaluated using conventional resistance
equations. In practice, the calculation is not easy because the loss coefficients
depend on particular geometric characteristics and must be determined empirically,
and the usual friction equations are not reliable because the barrel is usually too
short for flow to be fully developed.
If the coefficients and friction factor are known, the energy loss can be estimated
by

h ¼ he þ hb1 þ hb2 þ ho þ hexp þ hf ð4:21Þ

in which h is the total head difference across the structure, he is the entrance and
converging transition loss, hb1 is the upper bend loss, hb2 is the lower bend loss, ho
is the outlet transition loss, hexp is the exit loss and hf is the friction loss.
The United States Bureau of Reclamation (1973) gives the following estimates
for these loss components for typical rectangular siphon structures, which can be
used for preliminary design purposes.

V2
he ¼ 0:2 ð4:22Þ
2g

V2
hb1 ¼ hb2 ¼ 0:42 ð4:23Þ
2g

for bends with a centre-line radius equal to 2.5 times the barrel height.
 
V 2 Vo2
ho ¼ 0:2  ð4:24Þ
2g 2g

for diverging outlets and


4.2 Spillway Structures 139

 
V2 V2
ho ¼ 0:1 o  ð4:25Þ
2g 2g

for converging outlets.


 2
A V2
hexp ¼ ð4:26Þ
Ao 2g
(which is actually just the velocity head at the outlet, V2o/2g).

V2
hf ¼ 0:25 ð4:27Þ
2g

In these expressions, V is the barrel velocity, Vo is the outlet velocity, A is the


barrel area and Ao is the outlet area. These relationships give indications of absolute
and relative magnitudes, and should be used with caution and judgement. The two
bend losses are unlikely to be equal, for example, as the upper bend is usually longer
and tighter than the lower one. Assuming the exit loss to be the exit velocity head
ignores the downstream flow conditions, and could be more rigorously accounted for
in the way described for culverts in Chap. 5, i.e. by assuming h to represent the
difference between upstream energy and the hydraulic grade line (water surface) at
the outlet. The water level at the outlet can be determined by applying momentum
conservation between this section and the tailwater as controlled from downstream.
The friction loss, hf, could also be better estimated for particular structures using a
resistance equation such as Manning or Darcy–Weisbach, although the boundary
layer will not be fully developed for some distance downstream from the inlet (the
equations relating boundary layer thickness to distance for a spillway (Sect. 4.3.2)
would give indications of the length of barrel required for the boundary layer from
the top and bottom surfaces to reach the centre line).
With all the loss components expressed in terms of the barrel velocity, Eq. (4.21)
provides a relationship between V (and hence Q = VA) and h, the head difference
across the structure. If the upstream and downstream flow velocities are small, h can
be assumed to be the difference between the corresponding water levels.
Tadayon and Ramamurthy (2013) used computational fluid dynamics modelling
to simulate the flow through a fully primed siphon spillway and used the results to
characterize the discharge capacity in terms of a discharge coefficient (Cd) in the
equation
pffiffiffiffiffiffiffiffi
q ¼ Cd d 2gh ð4:28Þ

in which q is the unit width discharge and d is the siphon barrel depth. They found
good agreement between their predictions and Cd values measured in two physical
models with different geometries. They make the point that numerical modelling is
less expensive and time-consuming than physical modelling.
140 4 Spillways

It is not possible to predict analytically or computationally the water levels at


which priming and de-priming will occur, and hence the ranges of levels for which
the unprimed and primed control relationships apply. Final designs of siphon
spillways usually require physical model testing to establish the priming conditions
and to refine the control relationships. Model tests are necessary for determining the
rating curves for air-regulated (whitewater) siphons.
Siphon spillways are vulnerable to cavitation, especially in the region of the
internal crest where the pressure is lowered considerably by the flow curvature. It is
widely accepted that the local absolute pressure head should not be less than 3.0 m,
i.e. about 6.71 m below atmospheric at sea level. The variation of atmospheric
pressure with altitude must be accounted for when specifying the allowable pres-
sure head, pc /c above sea level. The United States Bureau of Reclamation (1973)
specifies allowable subatmospheric pressure heads for conduits flowing full, which
can be represented by
 
pc
 ¼ 6:71  0:0010z ð4:29Þ
c allowable

where z is the elevation (in m) above sea level.


The pressure in the throat, pc, can be estimated using energy considerations.
Taking the datum at the crest level, the total energy at the crest, Hc, is

pc v2
Hc ¼ þ c
c 2g

in which vc is the local velocity. The pressure head at the crest is therefore

pc v2
¼ Hc  c ð4:30Þ
c 2g

Under the assumption of irrotational flow, the total energy over the crest section
is constant and the Bernoulli equation can be applied across the section. Hc can
therefore be found by subtracting the intervening losses from the upstream energy
level, i.e.

Hc ¼ H  losses ð4:31Þ

in which H is the height of the upstream energy line above the crest level and the
losses would include the entrance loss, some bend loss and some friction loss.
The velocity at the crest, vc, can be determined from ideal flow theory. The
region of flow beyond the developing boundary layer in the upper portion of the
siphon is irrotational and the flow through the upper bend is approximately the
same as for an irrotational vortex, for which the local velocity (v) is inversely
proportional to the radius of curvature (r), and so
4.2 Spillway Structures 141

vr ¼ vc rc ð4:32Þ

in which v is the local velocity at radius r and rc is the radius of the crest segment.
(The velocity is thus highest, and the pressure consequently lowest, at the crest.)
The velocity at the crest, vc, can be related to the discharge, Q, through the siphon
by determining the discharge as the integral of the local velocity across the section
and describing the local velocity by Eq. (4.32), i.e.
Z ro
Q¼ vbdr
rc
Z ro
rc
¼ vc bdr
rc r
ro
¼ vc rc b ln
rc

in which ro is the outer radius of the barrel and b is the width of the barrel. For a
given discharge, the velocity at the crest is therefore given by

Q
vc ¼ ð4:33Þ
rc b ln rroc

Substituting Hc from Eq. (4.31) and vc from Eq. (4.33) into Eq. (4.30) enables
the local pressure head at the crest to be determined, which can then be compared
with the allowable value according to Eq. (4.29).
Because of the assumptions required in analysing siphon and shaft spillways, the
results should be regarded as preliminary. Physical model studies are often used to
confirm and/or refine the analysis and design details.
Example 4.3
An S-shaped siphon spillway has a rectangular barrel which is 16.0 m long, 2.00 m wide
and 0.80 m deep, with a diverging outlet reaching a depth of 1.20 m at the exit. The
centre-line radius of the upper and lower bends is 2.0 m. The concrete inner surface has
a Manning n of 0.012. The spillway serves a reservoir at an altitude of 550 masl.

a. Using the United States Bureau of Reclamation (1973) typical energy loss estimates,
determine the upstream water level above the inner crest when the discharge is 15.2 m3/s
and the tailwater level is 8.0 m below the crest. Assume fully primed conditions, the
outlet to be submerged and the velocity heads upstream and downstream to be negligible.
b. Assess the likelihood of cavitation in the structure at these conditions.

Solution

a. The height of the water level above the crest, Hu, is given by

Hu ¼ h  8:0

where h is the total energy difference across the structure which is dissipated by friction,
entrance, outlet, exit and bend losses (as shown in Fig. 4.27), i.e.
142 4 Spillways

h ¼ he þ hb1 þ hb2 þ ho þ hexp þ hf

where

V2
he ¼ 0:20
2g
V2
hb1 ¼ hb2 ¼ 0:42
2g
 2 
V Vo2
ho ¼ 0:20 
2g 2g

with

A D
Vo ¼ V ¼V
Ao Do
0:80
¼V ¼ 0:67V
1:20
so

V2   V2
ho ¼ 0:2 1  0:672 ¼ 0:11
2g 2g
 2 2  2 2
A V D V
hexp ¼ ¼
Ao 2g Do 2 g
 2 2
0:8 V V2
¼ ¼ 0:44
1:2 2 g 2g

V2
hf ¼ 0:25
2g

(Note that calculating hf by the Manning equation would give a slightly smaller loss.)
and

Q 15:2
V¼ ¼ ¼ 9:50 m/s
A 2:0  0:80
Therefore
9:502
h ¼ ð0:20 þ 0:42 þ 0:42 þ 0:11 þ 0:44 þ 0:25Þ ¼ 8:47 m
2  9:8
and
Hu ¼ 8:47  8:00 ¼ 0:47 m

b. The likelihood of cavitation can be assessed by comparing the pressure head at the crest
with the permissible value.
The pressure head at the crest is given by

pc v2
¼ Hc  c
c 2g
4.2 Spillway Structures 143

in which Hc and v-c are the energy and local velocity at the crest. Hc can be estimated by
subtracting losses from the upstream energy up to the crest. It is assumed that the losses
comprise the entrance loss, half of the first bend loss and friction over the length of half
the first bend (a quarter circle). The friction loss this calculated is an overestimate, and
therefore conservative, because the boundary layer will not be fully established over this
short distance. Then

1 Lbend
Hc ¼ Hu  he  hb1  Sf
2 2
where Lbend is the length of the first bend and Sf is the friction gradient. Using the USBR
loss coefficients and Manning’s equation for Sf,

V2 1 V 2 2pr V 2 n2
Hc ¼ Hu  0:20  0:42 
2g 2 2g 4 R4=3
2
9:50 1 9:502 2p2:0 9:502 0:0122
¼ 0:47  0:20  0:42   4=3
2  9:8 2 2  9:8 4 2:00:80
2ð2:0 þ 0:8Þ

¼ 0:47  0:92  0:97  0:22


¼ 1:64 m

The velocity at the crest is given by Eq. (4.33), i.e.

Q
vc ¼
rc b ln rroc
D
rc ¼ rcentre line 
2
0:80
¼ 2:00  ¼ 1:60 m
2
D
ro ¼ rcentre line þ
2
0:80
¼ 2:00 þ ¼ 2:40 m
2

Therefore
15:2
vc ¼ ¼ 11:7 m/s
1:60  2:0  ln 2:40
1:60

and

pc 11:72
¼ 1:64  ¼ 8:62 m
c 2  9:8

The permissible pressure head for no cavitation is given by Eq. (4.29), i.e.
 
pc
 ¼ 6:71  0:0010 z
c allowable

where z is the altitude.


144 4 Spillways

So
 
pc
¼ 6:71 þ 0:0010  550 ¼ 6:16 m
c allowable

The actual pressure head at the crest is lower than the permissible value, so cavitation is
very likely for this design and appropriate modifications to the geometry should be made.

4.2.7 Chutes

A chute is a steep channel which is often used in a spillway configuration to convey


the discharge from a control section to the river downstream (Fig. 4.28). Chute
spillways are the most common type for embankment dams because of their sim-
plicity, adaptability and overall economy. They may be used in conjunction with
straight (Fig. 4.29) or side-channel (Fig. 4.16) spillways or even through a col
separate from the embankment.
This arrangement significantly reduces the size of concrete structure required and
avoids disruption of the embankment construction and the problems of sealing
between the fill and the concrete. Often the spillway structure is cut into an abut-
ment or col and the excavation material used for embankment fill. The chute itself is
light and can be constructed on a wide range of foundation materials.
Flow in a chute is invariably supercritical and fast. The major hydraulic prob-
lems experienced in designing chutes are interference waves, translatory waves,

Fig. 4.28 A chute leading from a side-channel spillway


4.2 Spillway Structures 145

Fig. 4.29 Alternative


overflow spillways for an
embankment dam

(a) Looking upstream

(b) Plan

cavitation and aeration. Some of these problems are also experienced in other
structures and spillway components, and this discussion is not restricted in appli-
cation to chutes only.
Interference Waves
Interference waves are shock waves set up by any disruption of supercritical flow.
The patterns and depths of standing waves and some control measures have been
discussed in the chapter on transitions. Application to chutes would be similar.
Translatory Waves
Supercritical flow with very high velocity or on very steep slopes will become
unstable and a series of roll waves will form. These present the problems that higher
freeboards than otherwise are necessary, and that they can cause unsteady flow
impulses in terminal structures.
Roll waves can occur in channels with a wide range of slopes
(0.02 < S0 < 0.35). They can be prevented by artificially roughening the chute
surface, but this may promote cavitation. They can also be prevented by designing
the geometry so that the ratio of flow depth to wetted perimeter is large enough to
inhibit their formation. Novak et al. (2001) recommend ensuring a ratio of at least
0.1 for the maximum discharge, and accepting the presence of roll waves at lower
discharges when extra freeboard is not required and the effect on the terminal
structure is less crucial.
More rigorous criteria for the onset of unstable flow and methods for predicting
some characteristics of roll waves have been developed. Some of these are pre-
sented by French (1985).
146 4 Spillways

4.2.8 Stepped Chutes and Spillways

Flow at the downstream end of a spillway or chute is fast and high in kinetic
energy, and its release into downstream channels can cause considerable damage.
The excess energy is usually dissipated by deflecting the flow jet away from the
structure or in a stilling basin at the end of the slope (see Chap. 6). An alternative is
to provide a series of steps along the spillway face or chute, dissipating much of the
energy before the end of the slope and hence requiring a smaller terminal energy
dissipator (Fig. 4.30). This design is ideally suited to roller-compacted concrete
dam spillways, where steps are created in the construction process.
The characteristics of stepped chutes and spillways have been comprehensively
presented by Chanson (2002) and research into different aspects continues.
Flow down a series of steps can occur as either ‘nappe’ or ‘skimming’ flow
(Fig. 4.31), depending on the discharge and the step geometry. Nappe flow occurs
as a succession of freely falling nappes, with a fully or partially developed hydraulic
jump or decelerating supercritical flow on each step, and voids in the step spaces
(Fig. 4.31a–c). Skimming flow occurs as a uniform stream over the step edges, with
recirculating vortices within the step spaces (Fig. 4.31d). A gradual but rather
chaotic transition occurs between the two main regimes. Nappe flow is usually
associated with relatively low discharges or relatively large steps on relatively mild
slopes, and skimming flow with higher discharges, smaller steps and steep slopes.
Both types can occur in the same structure, but Chanson (2002) recommends that
the transition condition be avoided if possible because of the load fluctuations on
the structure associated with oscillations between the flow types. It is important to
be able to predict the flow type for design purposes because the associated energy
dissipation and flow depths are different.

Fig. 4.30 A laboratory model of a stepped chute


4.2 Spillway Structures 147

h yc
(a) Nappe flow, fully developed
hydraulic jump
l

(b) Nappe flow, partially


developed hydraulic jump

(c) Nappe flow, no hydraulic


jump

(d) Skimming flow

Fig. 4.31 Flow types in a stepped channel

Many empirical formulations have been presented for distinguishing between the
flow types, the differences between them arising from the different data used for
their determination as well as different interpretations or definitions of the types
(Fig. 4.32). Chanson (2002) defines the upper limit of nappe flow as

yc h
¼ 0:89  0:4 ð4:34Þ
h l

where yc is the critical flow depth related to the unit width discharge by q = (gy3c )0.5,
h is the step height and l is the step length, and the lower limit of skimming flow is
given by
148 4 Spillways

3.5

3.0 Eqn (4.38)

2.5 Eqn (4.37)

2.0 skimming flow


yc /h
1.5
Eqn (4.36)
1.0 Eqn (4.35)

Eqn (4.34)
0.5
nappe flow
0
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.7
h/l

Fig. 4.32 Criteria for distinguishing flow types in a stepped channel

yc h
¼ 1:2  0:325 ð4:35Þ
h l

Conditions between these limits correspond to the transition zone. Boes and
Hager (2003) define the onset of skimming flow, with no reference to a transition,
which is given by

yc h
¼ 0:91  0:14 ð4:36Þ
h l

James et al. (1999) and Chamani and Rajaratnam (1999) found skimming flow to
be more and more difficult to obtain with decreasing chute slope, leading to non-
linear criteria. James et al. (1999) proposed Eq. (4.37) for the upper limit of nappe
flow, which agrees well with Chanson’s criterion [Eq. (4.34)] for steep slopes.
 1:07
yc h
¼ 0:541 ð4:37Þ
h l

Chamani and Rajaratnam (1999) proposed Eq. (4.38) for the onset of skimming
flow.
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 

h yc 1 yc  0:34
¼ 0:89  þ 1:5  1 ð4:38Þ
l h h
4.2 Spillway Structures 149

Equations (4.37) and (4.38) are virtually coincidental, suggesting negligible


transition, although James et al. observed a small transition zone for low channel
slopes and a large transition for steep slopes (>*h/l = 0.7). Chanson (1996) pre-
sented an analysis to predict the onset of skimming flow, assuming the transition to
occur when the air cavity beneath the falling nappe disappears. The result predicts a
trend similar to Eqs. (4.37) and (4.38), although less steep at low values of h/l.
Energy dissipation mechanisms are different for the two types of flow. In nappe
flow, energy dissipation occurs through jet impact on the step surface and through
turbulence within the hydraulic jump, while in skimming flow it occurs through
mixing between the primary flow and the recirculating water in the step spaces.
Various formulations for energy loss have been presented, some of which are
presented below.
For nappe flow, Chanson (1993, 1994) used a combination of theoretical and
empirical analysis to quantify the energy loss along the chute as
yc 0:275 yc 0:55 !
DH 0:54 þ 1:72
¼1 h h
ð4:39Þ
Hmax 3
2 þ Hydam
c

in which DH is the total head loss along the chute, Hmax is the energy head upstream
of the structure and Hdam is the height of the dam structure. James et al. (1999)
found Eq. (4.39) to overestimate the energy loss for high values of yc/h (>*1.3). It
is usually useful for designing an energy-dissipating terminal structure to calculate
the residual energy, as Hres = Hmax − DH.
For skimming flow, Chanson (1993, 1994) proposed that the energy dissipation
is given by
 1=3  fe 2=3 !
DH
fe
cos a þ 12 8 sin
¼1 8 sin a a
ð4:40Þ
Hmax 3
2 þ Hdam
yc

in which a is the channel slope and fe is the effective friction factor. The effective
friction factor is influenced significantly by aeration and is difficult to predict.
Chanson (2002) suggests fe = 0.20 as an average, first approximation, value for
stepped chutes.
Equations (4.39) and (4.40) apply for long chutes where (quasi-) uniform flow
conditions are attained. Slightly modified forms of these equations were also pre-
sented for chutes controlled by underflow crest gates. Energy loss on low chutes
may be considerably less than predicted by these equations. As a general rule, the
rate of energy dissipation increases with dam height and decreases with increasing
discharge, so the steps become less effective during flood flows.
Various modifications to step geometry have been proposed in attempts to increase
energy dissipation. Chinnarasri and Wongwises (2004, 2006) compared the perfor-
mance of upwardly sloping steps and end sills at the ends of the steps, and found
modest improvement. James et al. (2001) tested a wedge-notch step geometry with the
150 4 Spillways

step edge deflecting flow towards the centre of the chute and towards the sides on
alternate steps. This geometry had little effect on energy dissipation in the nappe flow
regime, but increased it considerably in the transitional and skimming regimes.
Most results for stepped chutes and spillways are empirical. The limited pro-
totype data available makes for considerable uncertainty in design, suggesting that
physical modelling would be advisable. The high degree of aeration, especially with
skimming flow, limits the reliability of physical modelling, however. Chanson
(2002) suggests that to ensure proper scaling of air entrainment and flow resistance
with Froude similitude, the scale should not be less than 1:10 to 1:12.

4.3 Cavitation and Aeration on Spillways and Chutes

4.3.1 Cavitation

Cavitation occurs when cavities or ‘bubbles’ of water vapour and other liberated
gases form in flowing water. These may be carried by the flow to regions of higher
pressure where they implode. Severe damage can result if implosion of cavities is
repeated in quick succession against a concrete or steel surface where pressures as
high as 1500 MPa have been recorded (Lesleighter 1988). Dong et al. (2010) cite
some impressive reports of cavitation damage. At Hoover Dam, cavitation was
triggered by a bulge in the concrete tunnel lining and led to erosion through the
lining and 13.7 m into the underlying rock over a length of 35 m (Warnock 1947).
Similar damage occurred at the Liujiaxia Hydropower Station in China, where three
8 mm steel bars protruding through the concrete surface precipitated cavitation and
led to erosion of a hole which is 100 m long and 4.8 m deep (Liu 1983). Cassidy
and Elder (1984) report other cases: On a chute in the Reza Shah Kebir Project in
Iran, cavitation caused a hole which eroded through the 1.5 m concrete floor of the
chute and a further 1 m into the underlying rock. On the overflow spillway of the
Bratsk Dam in Russia, damage included a hole of 7.5 m by 10.5 m in plan and
1.2 m deep. Kells and Smith (1991) give details of further cases.
The consequences of cavitation are therefore not trivial, and the conditions under
which it occurs are not extraordinary. Assessing its likelihood is essential in the
design of any hydraulic structure conveying water at high velocities, and this
requires understanding its origins, recognizing the structural features associated
with it and being able to prevent or control it. Falvey (1990) and Kells and Smith
(1991) have provided comprehensive and detailed treatment of the phenomenon
and recommended design methods for its prevention.
Cavities form in water if the local pressure is the same as the water vapour pressure,
which depends on temperature. Cavities can therefore be induced by increasing the
temperature of the water to where the vapour pressure is equal to the ambient pressure;
this is ‘boiling’, which therefore takes place at different temperatures at different
altitudes. More relevant here is cavitation induced by the reduction of local pressures
4.3 Cavitation and Aeration on Spillways and Chutes 151

to the vapour pressure corresponding to the ambient temperature. The reduction in


pressure necessary to initiate cavitation can result from local flow velocities being
sufficiently increased by small changes in boundary geometry. Cavitation problems
can be anticipated where velocities exceed about 20 m/s and irregularities exist on the
flow surface, such as associated with joints or poor concrete finishes. Convex cur-
vatures, such as on spillway crests or in chute transitions, also contribute to creating
conditions favourable for cavitation by reducing pressure.
Predicting the conditions at which cavitation is likely to occur is not practically
possible through theoretical analysis, but many empirical relationships have been
developed from experiments with different boundary geometries by (for example)
Ball (1963, 1975, 1976) and Johnson (1963) and presented by Cassidy and Elder
(1984). These authors present graphs defining the critical velocity at which incipient
cavitation can be expected for different geometries of discrete roughness features
such as positive and negative vertical offsets. A commonly used indicator of cav-
itation potential is the ‘cavitation index’, r, which accounts for both flow velocity
and pressure. This index is derived by equating the energy at two points in the flow
in terms of the Bernoulli equation, i.e.

Vo2 po V2 p
þ þ zo ¼ þ þz ð4:41Þ
2g c 2g c

in which V is velocity, p is pressure, g is gravitational acceleration, c is the specific


weight of water and z is elevation. The subscript o represents a reference location
and the unsubscripted terms apply at a position on the boundary.
Assuming the elevation terms to be relatively insignificant, Eq. (4.41) can be
rearranged to
 2
V p  po
Cp ¼ 1  ¼ qVo2
ð4:42Þ
Vo
2

where q is the fluid density and Cp is the pressure coefficient or Euler number.
For cavitation conditions, the boundary pressure p becomes the vapour pressure
pv and the pressure coefficient takes its minimum value. The resulting parameter is
the cavitation index expressed as
po  pv
r¼ qVo2
ð4:43Þ
2

which can also be expressed as

Ho  Hv
r¼ Vo2
ð4:44Þ
2g

in which H is the pressure head (p/c).


152 4 Spillways

Different reference locations have been used by different researchers, such as the
undisturbed approach condition, the top of the turbulent boundary layer, the top of
roughness elements or protrusions potentially causing cavitation and the level of an
offset potentially causing cavitation. Using the undisturbed approach conditions
will usually be conservative.
It is important to note that both the reference and vapour pressures are absolute
values, i.e. they should include atmospheric pressure and the gauge pressure.
(Atmospheric pressure at sea level is 101.3 kPa (or 10.3 m of water), decreasing to
about 81.0 kPa at 2000 m above sea level and about 61.5 kPa at 4000 m above sea
level. The vapour pressure of water is 1.23 kPa at 10 °C, 2.33 kPa at 20 °C and
4.23 kPa at 30 °C.) The pressure should also include any centrifugal components
present and the slope term for steep chutes, i.e.
 
y V2
p ¼ pa þ c y cos h  ð4:45Þ
g r

in which pa is the atmospheric pressure, y is the flow depth, h is the bed slope, r is
the radius of curvature of the bed and the + and − signs correspond to concave and
convex surfaces, respectively.
The U S Army Corps of Engineers (1995) has consolidated much of the
experimental work and presented curves defining incipient cavitation for different
situations; those for vertical offsets into and away from the flow are shown in
Fig. 4.33. Other curves presented by the U S Army Corps of Engineers indicate

2.5

D
2.0
Cavitation Index

1.5

D
1.0

0.5

0
0.1 1 10 100
D (mm)

Fig. 4.33 Cavitation characteristics for abrupt offsets into and away from the flow (adapted from
U S Army Corps of Engineers 1995)
4.3 Cavitation and Aeration on Spillways and Chutes 153

Table 4.1 Criteria for prevention of cavitation damage (adapted from Falvey 1983)
Cavitation Design measures
index
>1.80 No cavitation expected
0.25 − 1.80 Protection by surface treatment, including grinding of roughness elements to
specified chamfers
0.17 − 0.25 Protection by modifying chute profile (e.g. reducing curvature) or including
aeration devices
0.12 − 0.17 Protection by including aeration devices
<0.12 Protection probably not possible

significant reduction in cavitation potential with rounded or chamfered offsets into


the flow, although Dong et al. (2010) found a chamfer to have little effect.
It is not only discrete roughness forms, such as construction joints, that can
initiate cavitation, but so can a rough surface, such as poorly finished or deteriorated
concrete. Arndt and Ippen (1967) studied flow over triangular grooves and sug-
gested that cavitation could occur if the cavitation index exceeds about four times
the Darcy–Weisbach friction factor.
In practical designs, it is difficult to predict local pressures accurately. Surface
irregularities are also sometimes unavoidable. It is therefore necessary to take
measures to prevent cavitation if there is any danger of it occurring. Falvey (1983)
(as presented by Kells and Smith 1991) proposed design measures for preventing
cavitation, depending on anticipated values of the cavitation index (Table 4.1).
Aeration provides effective prevention of cavitation by increasing the com-
pressibility of the water near the solid surface. Volkart and Rutschmann (1984)
report that a concentration of air bubbles of only 0.1% by volume increases the
mean compressibility by a factor of about 10. The compressibility introduced by air
absorbs the impact of the collapse of cavities and effectively eliminates pitting.
Aeration occurs naturally on spillway and chute surfaces, but can also be induced or
enhanced by the installation of aerating devices if this is insufficient to prevent
cavitation. The use of aerators is discussed in the next section, after an explanation
of self-aeration.

4.3.2 Aeration

Supercritical flow on a very steep slope will become aerated even without the
installation of air entrainment devices. This is obviously beneficial for cavitation
protection and also contributes to energy dissipation. However, aeration also causes
bulking and hence an increase in flow depth, requiring higher side walls than for
unaerated flow. When designing a chute it is necessary to predict the aerated flow
depth for side wall sizing, and also the air concentration to check if this is sufficient
154 4 Spillways

Fig. 4.34 An aerated chute leading to a stilling basin

to prevent cavitation. Figure 4.34 shows a chute with self-aeration assisted by


aeration notches.
The location where self-aeration begins is quite obvious on spillway faces with
the onset of ‘white water’ flow, and is generally assumed to occur where the
turbulent boundary layer has developed sufficiently to penetrate the full depth of
flow. Aeration then begins at the water surface and extends through the flow until it
is fully aerated some distance further downstream. The development of flow from a
control section along a steep slope is illustrated in Fig. 4.35.
Locating the inception of aeration is an important first step in describing the
development of aeration along the spillway or chute surface for the purposes of
providing cavitation prevention devices and estimating the bulked flow depth.
Falvey (1980) has described some early approaches used for determining the

Fig. 4.35 Development of non-aerated flow


flow and self-aeration on a
chute or spillway face H boundary layer

inception point
x δ
limit of air penetration
L y
partially aerated flow
x* fully aerated flow

θ
4.3 Cavitation and Aeration on Spillways and Chutes 155

distance of the inception point from the spillway or chute crest (L). The assumption
that self-aeration begins where the turbulent boundary layer reaches the water
surfaces suggests that this distance can be determined from descriptions of the
growth of the boundary layer. Probably the earliest application of this notion was by
Hickox (1939) who proposed

L ¼ 14:7q0:53 ð4:46Þ

in which q is the unit width discharge.


Various relationships between boundary layer thickness (d) and distance (x) have
been proposed, including the following:
Annemuller (1958):

d
¼ 0:010 ð4:47Þ
x

Beta et al. (1963):

d
¼ 0:010 to 0:016 ð4:48Þ
x

Schlichting (1968):
 0:2
d Vx
¼ 0:37 ð4:49Þ
x m

in which V is the free stream velocity (calculated at the water surface assuming no
energy loss from the head on the crest) and m is the kinematic viscosity of the water.
Following from Bauer (1954), various researchers have accounted empirically
for the effect of surface roughness on the development of the boundary layer using
the relationship
 b
d x
¼a ð4:50Þ
x ks

in which a and b are constants and ks is the (Nikuradse) equivalent roughness of the
surface. Bauer (1954) proposed a = 0.024 and b = 0.135. Castro-Orgaz (2009) has
provided expressions for a and b in terms of the exponent of the power law for the
vertical velocity profile in a turbulent boundary layer. Using previously obtained
results, Wood et al. (1983) quantified a and b to obtain
 0:11  0:10
d x x
¼ 0:0212 ð4:51Þ
x Hs ks
156 4 Spillways

Fig. 4.36 Energy diagram


for developing flow on a H ≈ 3/2 yc ~yc
chute or spillway
Hs = H + L sin θ – y cos θ
= V 2/2g
L sin θ δ
L
y

in which Hs is the potential flow velocity head at x, i.e. the distance from the
upstream total energy level and the local water surface (Fig. 4.36).
Application of Eqs. (4.47) to (4.51) to determine L requires estimation of the
flow depth (y) at L to define the boundary layer thickness at the onset of aeration.
Novak et al. (2001) suggest using standard nonuniform flow calculations in the
non-aerated zone, although the accuracy of this is questionable because this theory
assumes fully developed turbulent flow; the ideal flow part of the flow is unaffected
by the boundary roughness and the usual resistance coefficients do not apply. An
alternative is to assume purely ideal flow with no developing boundary layer. Under
this assumption, there is no energy loss up to the inception point (Fig. 4.26) and
both y and L will be slightly underestimated. The flow depth at any location within
the assumed ideal flow (and particularly at the inception point) is related to the
discharge through continuity by
q
y¼ ð4:52Þ
V

where V is the velocity at that location, calculated as (see Fig. 4.36)

V ¼ ð2gHs Þ1=2 ¼ ð2gðH þ L sin h  y cos hÞÞ1=2 ð4:53Þ

For a spillway, the value of H can be calculated from the spillway equation. For
a chute, it may be assumed that the depth is critical (yc) at the beginning of the chute
and then H = 3/2 yc.
The distance to the inception point (L) can thus be found by a double iteration on
y and L using Eqs. (4.52) and (4.53) for y and one of Eqs. (4.47)–(4.51) for L. An
appropriate procedure would be: assume initial values for L and y from which V can
be calculated using Eq. (4.53); revise the estimate of y until the value given by
Eq. (4.52) matches the assumed value; calculate L using the equation selected from
(4.47) to (4.51) and compare this with the value assumed; repeat the procedure until
the L values converge.
Equations (4.50) and (4.51) also require estimates of ks. An appropriate value
can be obtained from tables for different surfaces in standard textbooks (e.g.
4.3 Cavitation and Aeration on Spillways and Chutes 157

Henderson 1966); ks for very smooth concrete is about 0.15 mm. A value could
also be estimated from a known resistance coefficient, such as Manning’s n, by
expressing it as the friction factor (f) and then using this to solve for ks in Eq. (1.37)
or (1.38).
Castro-Orgaz (2009) accounted for both the irrotational and boundary layer
regions of flow. He provided analytical solutions of the energy and momentum
equations applicable to the irrotational flow and boundary layer, respectively, to
describe the development of the water surface profile and the boundary layer. These
can be applied directly to determine the location where the boundary layer reaches
the surface. Castro-Orgaz also used his equations to formulate an equation to give
the inception length by iteration.
Wilhelms and Gulliver (2005b) suggest that the depth at the inception point can
be assumed to be the unaerated uniform flow depth determined by the surface slope
and roughness, and can be calculated by a standard resistance equation, such as
Manning’s or Chézy’s. This simplifies the calculation of L considerably, as no
iteration for y is required, although iteration for L is necessary in Eqs. (4.49)–(4.51).
The notion that aeration commences where the boundary layer reaches the water
surface may lead to overestimation of the distance to inception because intermittent
turbulence occurs up to 1.2d above the bed (Falvey 1980), so aeration may actually
begin before the boundary layer reaches the water surface. Falvey (1980) presents a
different approach to finding the inception point, attributed to Bormann (1968). This
involves the application of two equations for the energy loss coefficient Cf, one
relating it to the boundary layer thickness and the other to the distance Reynolds
number. This approach also requires iteration, the correct value of L being that
which gives the same value of Cf by the two different equations.
The distance from the inception point to where the flow can be considered to be
fully aerated appears not to have been reliably established, but Falvey (1980) infers
from the experimental results of Bormann (1968) that it is of the same order of
magnitude as the distance to the inception point. More refined analysis of partially
and fully aerated nonuniform flow is discussed by Wood et al. (1983), Haindl
(1984) and Ackers and Priestley (1985).
Sizing the side walls of a chute requires estimation of the flow depth, which is
greater than for unaerated flow because of the bulking effect of the air. It is quite
difficult to define the surface of the flow because the mixture of air and water
becomes predominantly air gradually rather than abruptly. For analysis purposes, it
is often defined as the level at which the air concentration integrated to that level is
a specified value, such as 98%, 95% or 90%. Most researchers have regarded the
flow as a well-mixed continuum of air and water, but Wilhelms and Gulliver
(2005a) have suggested that two forms of conveyed air exist: ‘entrained air’ which
has been drawn into the flow and is transported as bubbles, and ‘entrapped air’
which is transported with the flow in the roughness or waves of the very irregular
water surface. Various ways have been proposed to calculate the effective bulked
flow depth, yb, to which a small freeboard is usually added for designing side walls.
158 4 Spillways

Wood (1984) proposed two simple relationships for estimating the bulked flow
depth (assuming this to be the depth to where the integrated air concentration is
90%), i.e.
 2 1=3
q
yb ¼ cyc ¼ c ð4:54Þ
g

in which yc is the critical flow depth and c is an empirical coefficient in the range
0.32 < c < 0.37, and
yb  yo pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
¼ 0:1 0:2Fr 2  1 ð4:55Þ
yo

in which yo is the unaerated uniform flow depth and Fr is the Froude number of this
flow.
Most other relationships for predicting yb recognize its dependence on the
concentration of air in the flow, C = Qa/(Qa + Q), where Qa is the discharge of air
and Q is the discharge of water. The ratio of water discharge to total discharge is
defined by q1 = Q/(Qa + Q) which for a rectangular channel is equal to the ratio
yo/yb. It follows that
yo
yb ¼ ð4:56Þ
1C

Calculating yb through Eq. (4.56) requires estimation of C and yo. Various


methods have been proposed for estimating C.
Novak et al. (2001) present relationships for C (in SI units) developed from
Anderson’s data (ASCE Task Committee 1961; Anderson 1965). For rough spill-
ways (ks = 1.2 mm in Anderson’s experiments),
 
S
C ¼ 0:7226 þ 0:743 log ð4:57Þ
q0:20

for the range of conditions 0.16 < S/q0.2 < 1.4. For smooth spillways,
 0:385
S
C ¼ 0:5027 ð4:58Þ
q2=3

for the range 0.23 < S/q2/3 < 2.3. In Eqs. (4.57) and (4.58), S is the channel slope.
The applicability of these equations can be checked by applying the usual criterion
for hydraulically rough and smooth flow in terms of Re*. According to Novak et al.
(2001), Eqs. (4.57) and (4.58) tend to underestimate C at the lower ends of the
defined ranges and as for all empirically determined relationships should be used
with caution beyond the ranges specified.
4.3 Cavitation and Aeration on Spillways and Chutes 159

Novak et al. (2001) also present equations proposed by Wood (1991) for the
average concentration (in %) as functions of the chute slope (in degrees), i.e.

3
C¼ h for 0 \h\ 40 ð4:59Þ
2

and

C ¼ 45 þ 0:36 h for 40 \h\70 ð4:60Þ

ICOLD (1992) provides a similar relationship, i.e.

C ¼ 0:75ðsin hÞ0:75 ð4:61Þ

Wilhelms and Gulliver (2005b) propose predicting the bulked flow depth with
an equation that is equivalent to Eq. (4.56) and with the concentration specified as
the total conveyed air (the sum of entrained and entrapped air) to the depth of 98%
air. They have determined that the entrapped air concentration can be assumed to be
0.23 at this level (and 0.14 and 0.073 for depths of 95% and 90% concentration,
respectively). The development of entrained air concentration downstream from the
inception point is described by
 

Ce ¼ Ce1 1  eax =yi ð4:62Þ

in which x* is the distance along the surface measured from the inception point, yi is
the depth at the inception point, a is the dimensionless spatial rate at which the
entrained air content approaches the equilibrium value Ce∞ and has a value of
0.010 ± 0.003. For 15° < h < 75°, Ce∞ can be determined as
 
Ce1 ¼ 0:656 1  e0:0356ðh10:9Þ ð4:63Þ

with h in degrees.
Equations (4.57) to (4.63) can be used to calculate the bulking effect through
Eq. (4.56). The unaerated flow depth, yo, must first be determined, however. This is
generally regarded as the uniform flow depth, although Falvey (2007) suggests that
the correction should be applied at points along a computed gradually varied flow
profile. Although Wilhelms and Gulliver (2005b) define yo in their expression of
Eq. (4.56) as the depth at the inception point, they still calculate it as the uniform
flow depth. The issue is further complicated by the fact that aeration reduces the
resistance to flow, and therefore the resistance coefficient used in conventional
resistance equations needs to be modified for aerated flow, and the yo used in
Eq. (4.56) reduced accordingly. Ackers and Priestley (1985) proposed adjusting the
Darcy–Weisbach friction factor according to
160 4 Spillways

fa
¼ 1  1:9C2 for C\0:65 ð4:64Þ
f

or

fa
¼ 0:2 for C [ 0:65 ð4:65Þ
f

in which f is the standard value and fa is the value to be used for aerated flow.
Example 4.4
A long rectangular chute leading from a critical control has a width of 5.0 m, a slope of
0.60 and a ks value of 0.30 mm. If the discharge is 12 m3/s, determine

a. the distance from the chute entry to where self-aeration of the flow would begin, using the
formulation of Wood et al. (1983),
b. the fully developed air concentration using the results of Wilhelms and Gulliver (2005b)
and
c. the bulked flow depth once the air concentration is fully developed.

Solution

a. The distance to the inception point, L is found by trial using Eq. (4.51) with x = L and
d = y, the flow depth at the inception point, i.e.

y
L¼  0:11  0:10 ðAÞ
L L
0:0212 Hs ks

Assuming ideal flow up to the inception point, the depth y is found by trial from
Eq. (4.53) while satisfying Eq. (4.52).
Calculations with the correct trial values are shown below:
Try L = 21.2 m.
With L = 21.2 m, try y = 0.156 m
Then
Hs ¼ H þ L sin h  y cos h

with
h ¼ arctan 0:60 ¼ 31:0

and
sffiffiffiffiffi!
3 3 3 q
2
H ¼ yc ¼
2 2 g
0sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1
3 @ 3 ð12:0=5:0Þ2 A
¼ ¼ 1:26 m
2 9:8
4.3 Cavitation and Aeration on Spillways and Chutes 161

So
Hs ¼ 1:26 þ 21:2 sin 31:0  0:156 cos 31:0
¼ 12:05 m

and
pffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
V¼ 2 g Hs ¼ 2  9:8  12:05 ¼ 15:4 m/s

Check:

q 12:0=5:0
y¼ ¼ ¼ 0:156 m
V 15:4

which is equal to the trial value. Therefore y = 0.156 m.


Now calculate RHS of (A) with the trial values:

0:156
RHS ¼  21:2 0:11   ¼ 21:1 m
21:2 0:10
0:0212 12:05 0:0003

which is approximately equal to the trial value of L. Therefore


L = 21 m.
b. The total air concentration is the sum of entrapped air and entrained air, i.e.

C ¼ Centrapped þ Ce1

where Centrapped = 0.23 to the depth of 98% air and Ce∞ is the equilibrium entrained air
concentration, given by Eq. (4.63), i.e.
 
Ce 1 ¼ 0:656 1  e0:0356ðh10:9Þ
 
¼ 0:656 1  e0:0356ð31:010:9Þ ¼ 0:32

This equation is valid for 15° < h < 75°; here, h = 31° falls well within this range.
Therefore

C ¼ 0:23 þ 0:32 ¼ 0:55

c. The bulked flow depth is given by Eq. (4.56), i.e.

yo
yb ¼
1C
in which C is the fully developed total air concentration and yo is the unaerated flow
depth, i.e. not bulked by entrained air, but with reduced friction associated with the air
concentration. This is calculated using the aerated flow friction factor which is related to
the unaerated value by Eq. (4.64) for C = 0.55 < 0.65, i.e.
162 4 Spillways

fa
¼ 1  1:9C 2
f

From Eq. (1.38),


1 12 R
pffiffiffi ¼ 2 log
f ks

with (using y = 0.156 m)


Wy

W þ 2y
5:0  0:156
¼ ¼ 0:147 m
5:0 þ 2  0:156

Therefore
1 12  0:147
pffiffiffi ¼ 2 log ¼ 7:54
f 0:00030

So
f ¼ 0:018

And C ¼ 0.55 (above)


Therefore
 
fa ¼ 0:018 1  1:9  0:552 ¼ 0:0077

Then yo is found from the Darcy–Weisbach equation together with continuity, i.e.
sffiffiffiffiffi
8gpffiffiffiffiffiffiffiffi
Q ¼ AV ¼ Wyo RSo
fa

Therefore
rffiffiffiffiffiffiffiffiffiffiffiffiffiffisffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
8  9:8 5:0yo
12:0 ¼ 5:0yo 0:60
0:0077 5:0 þ 2yo

from which, by trial, yo = 0.159 m


Therefore

0:159
yb ¼ ¼ 0:35 m
1  0:55

As mentioned before, aeration can significantly reduce the likelihood of cavitation,


but a minimum concentration of air is necessary for this to be effective.
Equations (4.57)–(4.63) enable estimates to be made of the average concentration
over the mixture depth, but the distribution is not uniform and it is the concentration
at the boundary that can be related to cavitation. Novak et al. (2001) suggest that
this should be at least 7% to be effective, and that the average concentration should
probably exceed 30% to ensure this. Wilhelms and Gulliver (2005b) related average
4.3 Cavitation and Aeration on Spillways and Chutes 163

and surface air concentrations as measured by several investigators and concluded


that an average concentration of 20% would be sufficient to ensure greater than 8%
at the surface. ICOLD (1992), together with Eq. (4.57) for the average concen-
tration, gives the concentration at the chute surface as

Co ¼ 1:25h3 for 0\h\40 ð4:66Þ

with h in radians, or

Co ¼ 0:65 sin h for h [ 40 ð4:67Þ

The description of the development of aeration presented above suggests that the
crest region and upper face of a spillway are vulnerable to cavitation, as the
inception point of aeration is some distance downstream of the crest and full
aeration even further along the spillway face. Even once full aeration has been
attained, Eqs. (4.62) and (4.66) suggest that the necessary average air concentration
might not be obtained by self-aeration for relatively mild slopes. Where
self-aeration is insufficient, it can be enhanced by the installation of devices for
introducing air to the underside of the flow. Various aerator configurations have
been suggested, some of which are illustrated in Fig. 4.37.

(a) Offset (b) Ramp

(c) Offset with air gallery (d) Ramp with air gallery

(f) Ramp and offset (f) Ramp and offset with air gallery

Fig. 4.37 Types of aerator (adapted from Cassidy and Elder 1984)
164 4 Spillways

The basic design of an aerator is an offset of the surface away from the flow,
sometimes including a ramp upwards to enhance the deflection of the flow from the
surface, and often with an enlarged area for air access. The general idea is to reduce the
pressure locally in order to entrain air. Obviously, without access to air, the reduction
in local pressure would actually increase the likelihood of cavitation. As the bubbles
of air introduced by an aerator will rise through the flow, the concentration of air will
decrease with distance downstream and a series of aerators will be required along the
length of the spillway or chute.
The design of aerator devices requires determination of
– whether or not aeration is required,
– the type and geometry of aerator to be used,
– the location of the most upstream device,
– the spacing between slots to maintain aeration as air bubbles rise to the water
surface and
– the size of the air supply duct.
Further details of aeration and cavitation and its prevention can be obtained from
the work of Volkart and Rutschmann (1984), Rutschmann and Volkart (1988), May
(1987) and Falvey (1980, 1990).

Problems
4:1 A side-channel spillway is to be designed for a small dam. The spillway crest
is straight, 25 m long and has a conventional ogee form above the side
channel. The side channel has a rectangular cross section with a width of
5.0 m, a longitudinal slope of 0.080 and Manning’s n = 0.020. The channel
extends along the full length of the spillway, with a 3.20 m high broad-crested
weir at the end, leading to a steep chute. Define the water surface profile along
the side channel when the water level in the reservoir is 0.80 m above the
spillway crest.
4:2 A lateral spillway channel conveys the water from a spillway crest that is
120 m long and discharges at 3.5 m3/s/m. The channel has a trapezoidal cross
section with a base width of 3.0 m, side slopes of 0.5H:1V and a longitudinal
slope of 0.150. Manning’s n is 0.0150. The channel begins at the upstream end
of the spillway and leads to a chute through a gradual transition downstream of
the end of the spillway. Determine the location of the critical section within the
side channel.
4:3 A side channel conveys water from a 25 m long ogee spillway crest to a chute.
The channel has a rectangular cross section with a width of 5.0 m, a longi-
tudinal slope of 0.080 and Manning’s n of 0.014. The channel begins at the
upstream end of the spillway and extends for 10.0 m beyond the downstream
end, where it terminates in a 3.0 m high broad-crested weir leading to the
chute. When the water level in the reservoir is 1.0 m above the crest, deter-
mine the flow depths in the channel
4.3 Cavitation and Aeration on Spillways and Chutes 165

a. immediately upstream of the broad-crested weir,


b. at the downstream end of the spillway crest and
c. midway along the spillway crest.
4:4 A canal with a 3.0 m wide rectangular cross section, a slope of 0.00050 and a
Manning’s n value of 0.013 incorporates an 8.0 m long sharp-crested side weir
with its crest 0.85 m above the bed.
a. If the discharge downstream of the weir is 4.5 m3/s, determine the flow
depths at the upstream and downstream ends of the weir and the discharge
diverted by the weir.
b. Sketch and classify the water surface profile upstream of the weir and deter-
mine the distance to where the flow depth is 0.90 times the uniform depth.
4:5 Produce a rating curve for the shaft spillway shown below. The discharge
coefficient for the crest is 2.07 in terms of the outer periphery length and the head
above the crest apex, and may be assumed constant over the range of conditions
for which discharge is crest controlled. The diameter of the horizontal conduit
and the vertical shaft is 2.0 m, and this connects to the crest profile 3.0 m below
the crest apex. Manning’s n for the shaft and conduit is 0.018.

4.30 m
H
3.0 m

15 m

S = 0.0040 2.0 m

125 m

4:6 An s-shaped siphon spillway has a rectangular, prismatic barrel which is


16.0 m long, 2.0 m wide and 1.3 m deep. The inner radius of curvature of the
bends is 1.5 m.
a. Determine the discharge under primed conditions when the reservoir level
upstream is 0.30 m above the crest and the tailwater level is 5.0 m below
it. The outlet is water sealed and you may neglect velocity heads upstream
and downstream.
b. Estimate the pressure head at the crest and comment on the possibility of
cavitation.
c. Estimate the length of overfall spillway that would be required to pass the
same discharge at the same upstream water level.
166 4 Spillways

4:7 A long rectangular chute forms part of the spillway structure of a dam at an
altitude of 1800 m. The chute is 2.0 m wide and leads from the reservoir
through a critical flow control. It has a slope of 0.25 and a Manning’s n of
0.013 and is designed for a discharge of 7.5 m3/s.
a. Using various methods, calculate the distance from the crest to where
self-aeration of the flow would be expected to begin. (Also confirm the
units to be used in Hickox’s (1939) equation.)
b. If a construction joint is located in the vicinity of the beginning of
self-aeration, what maximum vertical offset into and away from the surface
could be allowed without raising concerns about cavitation?
c. Using various methods, estimate the fully developed average air concen-
tration and the depth of the fully aerated uniform flow.

References

Ackers, P., & Priestley, S. J. (1985). Self-aerated flow down a chute spillway. In Proceedings of
the 2nd International Conference on the Hydraulics of Floods and Flood Control, Cambridge,
Paper A1, Cranfield: British Hydromechanics Research Association.
Anderson, A. G. (1965). Influence of channel roughness on the aeration of high-velocity, open
channel flow. In Proceedings, 11th International Association of Hydraulics Research
Congress, Leningrad, Vol. 1, Paper 1.37.
Anderson, R. M., & Tullis, B. P. (2012). Comparison of piano key and rectangular labyrinth weir
hydraulics. Journal of Hydraulic Engineering, 138(4), 358–361.
Annemuller, H. (1958). Luftaufnahme Durch Fliessended Wasser (Air Entrainment in Flowing
Water) (p. 146). Heft: Theodor-Rehvock Flussbaulaboratorium Universitat Fridericiana
Karlsruhe.
Arndt, R. E. A., & Ippen, A. T. (1967). Cavitation near surfaces of distributed roughness,
Hydrodynamics Report No. 104, Massachusetts Institute of Technology, June 1967.
ASCE Task Committee. (1961). Aerated flow in open channels. Transactions ASCE, 25, 456.
Ball, J. W. (1963). Construction finishes and high velocity flow. Journal of the Construction
Division, ASCE, 89(CO2), 91–110.
Ball, J. W. (1975). Cavitation damage caused by surface irregularities subjected to high velocities.
In ASCE Hydraulics Division Specialty Conference, Seattle, August 1975.
Ball, J. W. (1976). Cavitation from surface irregularities in high velocity flow. Journal of the
Hydraulics Division, ASCE, 112(HY9), 1283–1297.
Bauer, W. J. (1954). Turbulent boundary layer on steep slopes. Transactions ASCE, 119, 2719.
Beta, G., Jovanovic, S., & Bukmirovic, V. (1963). Nomographs for Hydraulic Calculations, Part 1,
Jaroslav Cerni Institute for Development of Water Resources, Belgrade, Yugoslavia, Vol. X,
No. 28, OTS 63-11451/ 3(in Serbo-Croat).
Boes, R., & Hager, W. H. (2003). Hydraulic design of stepped spillways. Journal of Hydraulic
Engineering, 129(9), 671–679.
Borghei, S. M., Jalili, M. R., & Ghodsian, M. (1999). Discharge coefficient for sharp-crested side
weir in subcritical flow. Journal of Hydraulic Engineering, 125(10), 1051–1056.
Bormann, K. (1968). Der Abfluss in Schussrinnen Unter Berucksichtigung der Luftaufnahme
(Discharge in chutes considering air entrainment), Versuchsantalt fur Wasserbau der
Technischen Hochschule Munchen, Bericht Nr 13.
References 167

Bradley, J. N., & Peterka, A. J. (1957). The hydraulic design of stilling basins. Proceedings of
American Society of Civil Engineers, 83(HY5), 1401–1406.
Cassidy, J. J. (1970). Designing spillway crests for high-head operation. Journal of the Hydraulics
Division, ASCE, 96(HY3), 745–753.
Cassidy, J. J., & Elder, R. A. (1984). Spillways of high dams, Chap 4. In P. Novak (Ed.),
Developments in Hydraulic Engineering, Vol. 2. Elsevier Applied Science Publishers.
Castro-Orgaz, O. (2009). Hydraulics of developing chute flow. Journal of Hydraulic Research, 47
(2), 185–194.
Chadwick, A & Morfett, J (1986). Hydraulics in Civil and Environmental Engineering, E & F N
Spon.
Chamani, M. R., & Rajaratnam, N. (1999). Onset of skimming flow on stepped spillways. Journal
of Hydraulic Engineering, 125(9), 969–971.
Chanson, H. (1993). Stepped spillway flows and air entrainment. Canadian Journal of Civil
Engineering, 20(3), 422–435.
Chanson, H. (1994). Comparison of energy dissipation between nappe and skimming flow regimes
on stepped chutes. Journal of Hydraulic Research, 32(2), 213–218.
Chanson, H. (1996). Prediction of the transition nappe/skimming flow on a stepped channel.
Journal of Hydraulic Research, 34(3), 421–429.
Chanson, H. (2002). The Hydraulics of Stepped Chutes and Spillways, A A Balkema.
Chinnarasri, C., & Wongwises, S. (2004). Flow regimes and energy loss on chutes with upward
inclined steps. Canadian Journal of Civil Engineering, 31(5), 870–879.
Chinnarasri, C., & Wongwises, S. (2006). Flow patterns and energy dissipation over various
stepped chutes. Journal of Irrigation and Drainage Engineering, 132(1), 70–76.
Crookston, B. M., & Tullis, B. P. (2013a) Hydraulic design and analysis of labyrinth weirs. I:
Discharge relationships, Journal of Irrigation and Drainage Engineering, 139(5), 363–370.
Crookston, B. M., & Tullis, B. P. (2013b) Hydraulic design and analysis of labyrinth weirs. II:
Nappe aeration, instability and vibration, Journal of Irrigation and Drainage Engineering, 139
(5), 371–377.
Dong, Z., Wu, Y., & Zhang, D. (2010). Cavitation characteristics of offset-into-flow and effect of
aeration. Journal of Hydraulic Research, 48(1), 74–80.
Falvey, H. T. (1980). Air-Water Flow in Hydraulic Structures, Engineering Monograph No. 41,
United States Department of the Interior, Bureau of Reclamation, Denver.
Falvey, H. T. (1983). Prevention of cavitation on chutes and spillways. In Proceedings of the
Conference on Frontiers in Hydraulic Engineering, American Society of Civil Engineers,
Cambridge, MA, August, pp. 432–437.
Falvey, H. T. (1990). Cavitation in Chute and Spillways, Engineering Monograph No. 42, United
States Department of the Interior, Bureau of Reclamation, Denver.
Falvey, H. T. (2007). Discussion of “Gas transfer, cavitation and bulking in self-aerated spillway
flow”. Journal of Hydraulic Research, 45(6), 859.
French, R. H. (1985). Open-Channel Hydraulics. McGraw-Hill.
Hager, W. H. (1988). Discharge characteristics of gated standard spillways. Water Power and Dam
Construction, 40(1), 15–26.
Haindl, K. (1984). Aeration at hydraulic structures, Chap. 3. In Developments in Hydraulic
Engineering, Vol. 2. Elsevier Applied Science Publishers.
Henderson, F. M. (1966). Open Channel Flow. Macmillan.
Hickox, G. H. (1939). Air entrainment on spillway faces. Civil Engineering, 9, 89–96.
ICOLD (1992). Spillways, Shock Waves and Air Entrainment. Bulletin 81, International
Commission on Large Dams, Paris.
James, C. S., Comninos, M., & Palmer, M. W. (1999). Effects of slope and step size on the
hydraulics of stepped chutes. Journal of the South African Institution of Civil Engineering,
41(2), 1–6.
James, C. S., Main, A. G., & Moon, J. (2001). Enhanced energy dissipation in stepped
chutes. Proceedings of the Institution of Civil Engineers, Water & Maritime
Engineering, 148(4), 277–280.
168 4 Spillways

Johnson, V. E. (1963). Mechanics of cavitation. Journal of the Hydraulics Division, ASCE, 89


(HY3), 251–275.
Kells, J. A., & Smith, C. D. (1991). Reduction of cavitation on spillways by induced air
entrainment. Canadian Journal of Civil Engineering, 18, 358–377.
Khode, B. V., Tembhurkar, A. R., Porey, P. D., & Ingle, R. N. (2012). Experimental studies on
flow over labyrinth weir. Journal of Irrigation and Drainage Engineering, 138(6), 548–552.
Knight, A. C. E. (1989). Design of efficient side-channel spillway. Journal of Hydraulic
Engineering, 115(9), 1275–1289.
Leite Ribeiro, M., Bieri, M., Boillat, J.-L., Schleiss, A. J., Singhal, G., & Sharma, N. (2012).
Discharge capacity of piano key weirs. Journal of Hydraulic Engineering, 138(2), 199–203.
Lesleighter, E. J. (1988). Cavitation in hydraulic structures. In Proceedings of the International
Symposium on Model-Prototype Correlation of Hydraulic Structures, ASCE/IAHR, Colorado
Springs, USA, August, pp. 74–94.
Liu, C. (1983). A study on cavitation inception of isolated surface irregularities. Collected
Research Papers, Water Conservancy and Hydroelectric Power Research Institute, Hydropower
Press, Beijing, Vol. 13, pp. 36–56 (in Chinese).
Machiels, O., Pirotton, M., Pierre, A., Dewals, B., & Erpicum, S. (2014). Experimental parametric
study and design of piano key weirs. Journal of Hydraulic Research, 52(3), 326–335.
May, R. (1987). Cavitation in hydraulic structures, Report SR79, HR Wallingford, Wallingford,
United Kingdom.
Murphy, T. E. (1973). Spillway crest design, Report MP H-73-5, U S Army Engineer Waterways
Experiment Station, Vicksburg, Mississippi, USA.
Novak, P., Moffat, A. I. B., Nalluri, C., & Narayanan, R. (2001). Hydraulic Structures (3rd ed.).
London: Unwin Hyman.
Reese, J. R., & Maynord, T. M. (1987). Design of spillway crests. Journal of Hydraulic
Engineering, 113(4), 476–490.
Rutschmann, P., & Volkart, P. (1988). Spillway chute aeration. International Water Power & Dam
Construction, January.
Schlichting, H. (1968). Boundary Layer Theory. McGraw-Hill.
Tadayon, R., & Ramamurthy (2013). Discharge coefficient for siphon spillways. Journal of
Irrigation & Drainage Engineering, 139(3), 267–270.
U S Army Corps of Engineers. (1995). Hydraulic design of spillways, Technical Engineering and
Design Guides as adapted from the U S Army Corps of Engineers, No. 12, ASCE Press,
American Society of Civil Engineers, New York.
United States Bureau of Reclamation. (1973). Design of Small Dams (2nd ed.). Washington, DC:
United States Department of the Interior.
Volkart, P., & Rutschmann, P. (1984). Air entrainment devices (air slots), Nr 72, Mitteilungen der
Versuchsanstalt für Wasserbau, Hydrologie und Glaziologie, Eidgenössischen Technischen
Hochschule, Zürich.
Warnock, J. E. (1947). Cavitation in hydraulic structures: experiences of the Bureau of
Reclamation. Transactions of the American Society of Civil Engineers, 112, 43–58.
Wilhelms, S. C., & Gulliver, J. S. (2005a). Bubbles and waves description of self-aerated spillway
flow. Journal of Hydraulic Research, 43(5), 522–531.
Wilhelms, S. C., & Gulliver, J. S. (2005b). Gas transfer, cavitation, and bulking in self-aerated
spillway flow. Journal of Hydraulic Research, 43(5), 532–539.
Wood, I. R. (1984). Air entrained flow in hydraulic structures. In H. Kobus (Ed.) Proceedings of
the International Association for Hydraulics Research, Symposium on Scale Effects in
Modelling Hydraulic Structures. Esslingen, Akademie Verlag.
Wood, I. R. (1991). Free surface air entrainment on spillways. In I. R. Wood (Ed.) Air Entrainment
in Free Surface Flows: IAHR Hydraulic Structures Design Manual, Vol. 4, Balkema,
Rotterdam.
Wood, I. R., Ackers, P., & Loveless, J. (1983). General method for critical point on spillways.
Journal of Hydraulic Engineering, 109(2), 308–312.
Chapter 5
Culverts

5.1 Introduction

A culvert is a closed conduit, usually relatively short, which is used where the
course of a natural or artificial channel is obstructed by, for example, a road or rail
embankment (Fig. 5.1). Culvert-type structures are also frequently used as outlet
works or spillways for small dams, particularly those constructed to provide
detention storage for urban storm water. The most significant characteristic of
culverts from a hydraulic point of view is that they may flow either full or partly
full, and they are sometimes classified as ‘hydraulically long’ or ‘hydraulically
short’ depending on this condition. Although probably the most common of all
hydraulic structures, their hydraulics is also amongst the most complex for a
number of reasons. First, there are three possible positions of control, and all may
govern at different discharges. Second, minor energy losses are significant in
relation to friction and must be accounted for. Third, different conditions of sub-
mergence may occur at the inlet, requiring different control relationships at the same
location for different discharges.
Culvert design is done by trial analysis, i.e. a size and geometry is selected and
this is analysed and modified until a satisfactory structure is found. The usual
situation is that a discharge is specified, which must be passed without exceeding a
specified limit on the upstream water level or headwater (HW). The purpose of the
analysis is therefore to establish the relationship between discharge and headwater,
Q(HW).
The effective control on discharge through a culvert is generally exercised either
by the inlet, the barrel or the tailwater conditions. It is common to refer to the last
two collectively as outlet control (e.g. U S Bureau of Public Roads 1965) because
the end of the barrel coincides with the outlet and controlling tailwater, and the
analysis is similar in principle for both. The hydraulic characteristics of inlet- and
outlet-controlled culverts are treated separately in the following discussion. It is
generally not possible to establish whether a culvert is inlet or outlet controlled

© Springer Nature Switzerland AG 2020 169


C S James, Hydraulic Structures,
https://doi.org/10.1007/978-3-030-34086-5_5
170 5 Culverts

Fig. 5.1 A single rectangular box (a) and multiple circular (b) culverts

beforehand; analyses must be done making both assumptions and the correct
condition is identified from the results.

5.2 Inlet Control

If a culvert is inlet controlled, the discharge is determined solely by the headwater


depth (representing the available energy) and the geometry of the inlet section
(specifically the cross-sectional shape, area and edge geometry). The length, size
and roughness of the barrel and the tailwater level do not affect the capacity.
Because the capacity is restricted at the inlet, the barrel is actually larger than it need
be, and it will flow only partly full.
There are several possible water surface profiles for flow under inlet control, and
the required relationship between Q and HW depends on which actually occurs. If
both the inlet and the outlet of the culvert are unsubmerged, the profile will be as
shown in Fig. 5.2a. Because there is a continuous free surface, the flow must pass
through critical at the inlet control section and flow will be supercritical through at
least part of the barrel. A hydraulic jump to subcritical flow may occur within the
barrel (broken line) or in the channel downstream of the outlet. For a rectangular
box culvert, this condition can hold until the water level upstream is 1.5 times the
barrel height (D) above the bottom of the control section (assuming the approach
velocity is negligible and the energy level is at the water surface). For design
purposes, the maximum height is usually assumed to be 1.2 times the barrel height,
to allow for some velocity head, entrance loss and shape effects.
If the inlet is submerged and the outlet is unsubmerged, the profile will be as
shown in Fig. 5.2b. The flow is the same as for an orifice or a sluice gate, and flow
will again be supercritical for at least some length of the barrel. As for the first case,
a hydraulic jump to subcritical flow may occur within the barrel (broken line) or
downstream of the outlet.
If the outlet is submerged a hydraulic jump will occur in the barrel, which will
flow full downstream from that position. The inlet may be submerged (Fig. 5.2c) or
5.2 Inlet Control 171

HW yc

(a) Inlet and outlet unsubmerged

HW

(b) Inlet submerged, outlet unsubmerged

air vent

HW

(c) Inlet and outlet submerged

HW yc

(d) Inlet unsubmerged, outlet submerged

Fig. 5.2 Flow profiles for inlet-controlled culverts

unsubmerged (Fig. 5.2d). If it is submerged an air vent is required to maintain the


inlet-controlled condition. If venting is not provided the air could be evacuated,
resulting in surging and erratic discharge. Complete evacuation of the air would be
analogous to priming of a siphon, and the point of control would change to the
tailwater.
For establishing a discharge relationship, two basic types of inlet-controlled flow
must be recognized. For low headwaters, the inlet will be unsubmerged and the
relationship will have the same form as the weir equation, i.e. Q / HW3/2, where
HW is the headwater depth (approximating the energy level if the approach velocity
is negligibly small). For higher headwaters, when the inlet is submerged, the flow
172 5 Culverts

Fig. 5.3 Rating curve for an


inlet-controlled box culvert

Q α HW 1/2

Headwater, HW
submerged inlet

Q α HW 3/2

HW ≈ 1.2D

unsubmerged inlet

Discharge, Q

will be orifice type and the relationship will have the form Q / HW1/2. The rating
curve will therefore be complex, as shown in Fig. 5.3.
Unlike for shaft and siphon spillways, the compound nature of the rating curve
does not result from a change in the position of control, but rather from a change in
the nature of the control at the inlet section.
For unsubmerged flow, the discharge is controlled by the critical flow section
and is given by

Q ¼ Ac Vc

in which Ac is the flow area and Vc is the velocity, both at the critical section. For a
rectangular box culvert Ac = Byc, where B is the width and yc is the critical flow
depth and Vc = (gyc)1/2. Therefore
pffiffiffiffiffiffiffiffi
Q ¼ B yc g yc

and, again for a rectangular section, yc = 2/3HW (assuming no energy losses before
the control section and negligible approach velocity), so
 3=2
pffiffiffi 2
Q¼B g HW 3=2
3

To account for approach velocity and energy loss before the control section, an
empirical discharge coefficient is introduced. The discharge equation is then
 3=2
pffiffiffi 2
Q ¼ CD B g HW 3=2 ð5:1Þ
3

For typical box culvert conditions, Stephenson (1981) gives CD = 0.92.


5.2 Inlet Control 173

For a circular culvert, the critical flow depth cannot be so easily defined in terms
of the geometry and energy. It must be determined from the general definition of the
Froude number, i.e.

Q2 B
Fr2 ¼ 1 ¼ ð5:2Þ
g A3

in which B is the surface width. Stephenson (1981) gives the following discharge
equation for Ec/D < 0.8, where Ec is the critical specific energy which for negligible
approach velocity can be represented by HW and D is the pipe diameter.
 
pffiffiffi 5=2 HW 1:9
Q ¼ 0:48 CB gD ð5:3Þ
D

with CB = 0.92, as for box culverts.


For the submerged condition (Fig. 5.2b), the discharge will be related to head-
water depth by an orifice-type equation. This can be derived by equating specific
energy at sections immediately upstream and downstream of the inlet. If the
approach velocity is negligible, the upstream specific energy can be represented by
HW, and

V2
HW ¼ y þ
2g

in which y and V are the flow depth and velocity at the vena contracta. Then
pffiffiffiffiffiffipffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
V¼ 2 g HW  y

The discharge is given by the product of this velocity and the cross-sectional
flow area. The area can be expressed as CcA, where Cc is a contraction coefficient
and A is the cross-sectional area of the inlet. The flow depth at the vena contracta
can be expressed as ChD, where Ch is a contraction coefficient and D is the height of
the inlet. The discharge is therefore given by
pffiffiffiffiffiffipffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Q ¼ Cc A 2 g HW  Ch D ð5:4Þ

The coefficients Cc and Ch depend on the inlet geometry. Stephenson (1981)


gives the values listed in Table 5.1. The geometry of the side-tapered box culverts
referred to in this table is shown in Fig. 5.4. Note that no barrel characteristics or
tailwater conditions appear in the inlet control equations or in the table of
coefficients.
Similar equations and loss coefficients have been published by the U S Bureau of
Public Roads (1965 and others), Bodhaine (1976) (also presented by French 1985)
and Ramsbottom et al. (1997).
174 5 Culverts

Table 5.1 Submerged inlet Geometry Cc Ch


control coefficients (adapted
from Stephenson 1981) Side-tapered box culverts
15°–26° wingwalls and top bevel 0.59 0.84
26°–90° wingwalls, no bevel 0.59 0.84
26°–45° wingwalls and top bevel 0.64 0.86
45°–90° wingwalls and side bevel 0.64 0.86
Circular culverts
Square edge 0.57 0.79
Bevel edge 0.65 0.83

bevel 4
1
HW symmetrical
D flare angles
15o – 90o

(a) Elevation (b) Plan

Fig. 5.4 Side-tapered inlet for box culverts (adapted from Stephenson 1981)

5.3 Outlet Control

Flow through a culvert is outlet controlled if the relationship between discharge and
headwater is determined by the barrel characteristics and, sometimes, the tailwater
level. In this case, the discharge is limited by the barrel and the inlet is larger than
necessary. The culvert will therefore flow either full or subcritical. Again, various
water surface profiles are possible and these are shown in Fig. 5.5.
If the inlet and outlet are both unsubmerged, as in Fig. 5.5a, flow will be free
surface and subcritical throughout the culvert. If the flow is actually controlled by
the barrel, rather than the tailwater, then critical flow will occur at the outlet section
(broken line).
If the inlet is submerged and the outlet is unsubmerged, as in Fig. 5.5b, then the
barrel will flow full over at least part of its length. If the tailwater level controls and
is below the top of the outlet, the water surface will separate from the soffit at some
point, but flow will remain subcritical. If control is by the barrel and the critical flow
depth is greater than the height of the barrel, it will flow full throughout its length
and emerge as a jet (upper broken line). If the critical depth is less than the height of
the culvert, then the flow will separate from the soffit at some point and flow
through critical at the outlet (lower broken line).
If both the inlet and outlet are submerged, as shown in Fig. 5.5c, then the culvert
will flow full and the tailwater level will control the discharge.
5.3 Outlet Control 175

HW
yc

(a) Inlet and outlet unsubmerged

HW

yc

(b) Inlet submerged, outlet unsubmerged

HW

(c) Inlet and outlet submerged

Fig. 5.5 Flow profiles for outlet-controlled culverts

he
hf H

HW hexit
hv

D
d
LSo So

Fig. 5.6 Energy losses through an outlet-controlled culvert

An equation for the discharge can be developed by accounting for all the energy
losses up to the control point. The losses through a culvert with inlet and outlet
submerged are shown in Fig. 5.6.
A common approach for describing the control relationship for barrel and tail-
water control is to relate the energy upstream to the hydraulic grade line (i.e. the
176 5 Culverts

water surface) at the outlet section (e.g. U S Bureau of Public Roads 1965). This
obviates the need to estimate exit losses and the downstream velocity head if
control is by the tailwater, but requires determination of the outlet flow depth (d).
The difference between the upstream energy level and the hydraulic grade line at
the outlet, H, has three components, viz.

H ¼ he þ hf þ hv ð5:5Þ

in which he is the entrance loss, hf is the friction loss and hv is the velocity head in
the barrel (V2/2g).
The entrance loss can be evaluated in terms of the velocity head in the culvert,
i.e.

V2
he ¼ k e ð5:6Þ
2g

in which V is the velocity in the culvert and ke is an empirical coefficient that


depends on the inlet geometry. Values of ke for different inlet geometries for pipe
and box culverts have been presented by Stephenson (1981) and are reproduced in
Table 5.2.
Table 5.2 Entrance loss coefficients for outlet-controlled culverts (adapted from Stephenson
1981)
Type of structure and entrance design ke
Pipe
Projecting from fill, socket end 0.2
Projecting from fill, square cut end 0.5
Headwall or headwall and wingwalls:
Socket end of pipe or rounded 0.2
Square edge 0.5
Mitered to conform to fill slope 0.7
End-section conforming to fill slope 0.5
Bevelled edges, 33.7° or 45° bevels 0.2
Side- or slope-tapered inlet 0.2
Metal pipe projecting from fill, no headwall 0.9
Reinforced concrete box sections
Headwall parallel to embankment, no wingwalls
Square on three edges 0.5
Round three edges to radius 1/12 barrel or bevelled three sides 0.2
Wingwalls 30°–75° to barrel
Square edge at crown 0.4
Crown edge rounded to radius 1/12 barrel or bevelled top 0.2
Wingwall 10°–25° to barrel, square edged at crown 0.5
Wingwalls parallel (extension of sides), square at crown 0.7
Side- or slope-tapered inlet (Fig. 5.3) 0.2
5.3 Outlet Control 177

The wide range of values of ke shows that the efficiency of a culvert can be
increased substantially by careful design of the inlet geometry, even if flow is outlet
controlled.
The friction loss, hf, can be estimated by a resistance equation. According to the
Darcy–Weisbach and Manning equations

f L V 2 V 2 n2 L
hf ¼ ¼ 4=3
8gR R

in which f is the friction factor, L is the length of the barrel, R is the hydraulic radius
of the barrel and n is Manning’s resistance coefficient. The friction loss can be
expressed in terms of the velocity head as
   
f L V2 19:6 n2 L V 2
hf ¼ ¼ ð5:7Þ
4R 2g R4=3 2g

Equation (5.5) can then be written as


  2
fL V
H¼ ke þ þ1 ð5:8Þ
4R 2g

using the Darcy–Weisbach friction factor, or


  2
19:6 n2 L V
H¼ ke þ 4=3
þ 1 ð5:9Þ
R 2g

The relationship between Q and HW is sought, and so H must be related to HW.


From the geometry of Fig. 5.6,

H ¼ HW þ L So  d ð5:10Þ

in which So is the slope of the barrel and d is the height of the hydraulic grade line
above the bottom of the barrel, i.e. the water depth at the outlet.
For a specified discharge (Q), the headwater (HW) required can be calculated by
inserting V = Q/A into either Eq. (5.8) or Eq. (5.9) to obtain H, and then using
Eq. (5.10) for HW. If the discharge corresponding to a given value of HW is
required, Eqs. (5.8) and (5.10) can be combined to give
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 gðHW þ L So  d Þ
V¼ ð5:11Þ
ke þ 4f LR þ 1

using the Darcy–Weisbach f, or


178 5 Culverts

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 g ðHW þ L So  d Þ
V¼ ð5:12Þ
ke þ 19:6 R4=3
n2 L
þ1

using Manning’s n.
Multiplying V by the cross-sectional area of the barrel gives the discharge in
terms of the culvert characteristics and the headwater level.
Calculation of either Q or HW requires estimation of the level of the hydraulic
grade line at the exit, d. If the barrel is the control and the critical flow depth, yc, is
greater than the barrel height, then the water will issue from the end of the barrel as
a jet and d = D. If yc is less than the barrel height, then flow will pass through
critical at the outlet. Because of the strong vertical curvature of the flow near the
critical condition, the pressure distribution will not be hydrostatic and the hydraulic
grade line will not be at the water surface, but some distance above it. The actual
position is not easy to determine and the U S Bureau of Public Roads (1965)
recommends that d should be taken as midway between yc and the barrel height in
this case.
If the outlet is submerged, the culvert is controlled by the tailwater level and d is
the water level at the outlet. The tailwater level must first be determined from the
characteristics of the flow in the channel downstream, which may require gradually
varied flow computations. The value of d can then be found by applying the
momentum conservation principle (either through the force–momentum flux
equation or using the momentum function) between the outlet section and a section
where the tailwater is defined, similarly to the analyses of an abrupt expansion-type
transition and a submerged sluice gate. If the discharge is known and the upstream
water level (HW) is required, the analysis is straight forward and explicit: the
downstream subcritical flow is determined from the channel conditions through a
resistance equation if flow is uniform, or from another downstream control if not;
the depth d is then calculated from the momentum analysis; HW is then calculated
using Eq. (5.11) or Eq. (5.12) with V = Q/A. If HW is specified and the corre-
sponding discharge is required, then three equations (downstream control,
momentum conservation and either Eq. (5.11) or Eq. (5.12) must be solved
simultaneously. (If a number of discharges for a range of HW values are required, it
may be easier to generate a rating curve for a range of discharges and use this to
read off values.)
For cases of outlet control where free flow occurs through the barrel, the analysis
will require gradually varied flow computations from the downstream control depth
to the headwater.
As with all conveyance structures with multiple possible control features, it may
not be initially obvious whether a culvert will operate under inlet control or outlet
control. The design will usually proceed by specifying a trial geometry and ana-
lysing this under both control assumptions. The analysis that predicts the higher
value of HW or the lower value of Q will indicate the governing control. This can
then be compared with the design requirement, and the geometry adjusted until the
design is satisfactory.
5.3 Outlet Control 179

Example 5.1
A 2.4 m square concrete (n = 0.013) box culvert has a side-tapered inlet with 30° flared
wingwalls and a top bevel. The barrel is 30 m long with a slope of 0.0050 and opens
abruptly into a long, 4.8 m wide rectangular channel.
Determine the headwater depth for a discharge of 30 m3/s. At this discharge, the uniform
flow depth in the downstream channel is 3.82 m.
Solution
Conditions of both inlet and outlet control must be analysed, with the greater headwater
depth indicating the actual condition.
Assuming unsubmerged inlet control, the discharge and headwater depth are related by
Eq. (5.1), i.e.
 3=2
pffiffiffi 2
Q ¼ CD B g HW 3=2
3

Therefore
!2=3
Q
HW ¼ pffiffiffi 3=2
CD B g 23

with
CD ¼ 0:92

Therefore
!2=3
30
HW ¼ pffiffiffiffiffiffiffi 3=2 ¼ 3:99 m
0:92  2:4  9:8 23

This HW value is more than 1.5 times the culvert height, so the inlet must be submerged.
For submerged inlet control, the discharge and headwater depth are related by Eq. (5.4),
i.e.
pffiffiffiffiffiffipffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Q ¼ Cc A 2 g HW  Ch D

Therefore
 2
Q
HW ¼ pffiffiffiffiffiffi þ Ch D
Cc A 2 g

with
Cc ¼ 0:64 ðTable5:1Þ

Ch ¼ 0:86 ðTable 5:1Þ

A ¼ D2 ¼ 2:42 ¼ 5:76 m2
180 5 Culverts

Therefore
 2
30
HW ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffi þ 0:86  2:4 ¼ 5:44 m
0:64  5:76  2  9:8

Assuming outlet control:

HW ¼ H þ d  L So ðrefer to Fig: 5:6Þ

V2
H ¼ he þ hf þ
2g
  2
2 g n2 L V
¼ ke þ 4=3 þ 1
R 2g

with ke ¼ 0:20 ðTable 5:2Þ

Q 30
V¼ ¼ ¼ 5:21 m/s
A 5:76
A 5:76
R¼ ¼ ¼ 0:60 m
P 4  2:4
Therefore
 
2  9:8  0:0132  30 5:212
H¼ 0:20 þ þ 1 ¼ 1:93 m
0:604=3 2  9:8

d is calculated from the force–momentum flux equation


between the outlet section and uniform flow downstream, i.e.

yo
d F2
F1

1 2
F1  F2 ¼ q QðV2  V1 Þ

1
F1 ¼ c d 2 B
2
1
¼ 9:8  103 d 2 4:80 ¼ 23:5 d 2  103 N
2
1
F2 ¼ c y2o W
2
1
¼ 9:8  103  3:822  4:80 ¼ 343  103 N
2
Q 30
V2 ¼ ¼ ¼ 1:64 m/s
A2 4:8  3:82
Q 30
V1 ¼ ¼ ¼ 5:21 m/s
A1 2:42
5.3 Outlet Control 181

Therefore

23:5 d 2  103  343  103 ¼ 103  30  ð1:64  5:21Þ

and so
 
30ð1:64  5:21Þ þ 343 0:5
d¼ ¼ 3:17 m
23:5

Therefore

HW ¼ 1:93 þ 3:17  30  0:0050 ¼ 4:95 m

HW for inlet control (5.44 m) is greater than for outlet control (4.95 m). Therefore, the
culvert is inlet controlled and HW = 5.44 m. This will be the water level upstream if
ponding occurs and the approach velocity is negligible.

Problems
5:1 A 2.4 m square concrete box culvert with a side-tapered inlet is 45 m long and
is set on a slope of 1:200. The entrance is unbevelled and has wingwalls with
flare angles of 40°. The culvert opens abruptly into a rectangular channel with
a width of 4.5 m, a slope of 1:1000 and a Manning’s n of 0.025.
a. Determine the headwater depth at the design discharge of 26 m3/s.
b. Determine the discharge when the headwater is 4.11 m.

References

Bodhaine, G. L. (1976). Measurement of peak discharge at culverts by indirect methods, in


Techniques of water resources investigations of the United States geological survey, Book 3,
Chapter A3. Washington: U S Geological Survey.
French, R. H. (1985). Open-channel hydraulics. McGraw-Hill Book Company.
Ramsbottom, D., Day, R., & Rickard, C. (1997). Culvert design manual. CIRIA Report 168,
London.
Stephenson, D. (1981). Stormwater hydrology and drainage, Elsevier Scientific Publishing
Company.
U S Bureau of Public Roads. (1965). Hydraulic charts for the selection of highway culverts.
Hydraulic Engineering Circular No 5.

Further Reading

Balkham, M., Fosbeary, C., Kitchen, A., & Rickard, C. (2010). Culvert design and operation
guide. CIRIA.
Chapter 6
Energy Dissipation Structures

6.1 Introduction

It is often necessary to provide for the dissipation of excess kinetic energy in free
surface flow situations. The most common instance is where water is released from
a reservoir, where the storage has caused a local increase in potential energy which
is converted to kinetic energy as the water is released, either over a spillway or
through outlet works. The high kinetic energy must be dissipated before the flow
reaches the natural river, or severe erosion could result.
Energy dissipation is also often required in artificial canal systems at the bottom
of drop structures.
There are two fundamental approaches to energy dissipation:
1. Inducement of turbulence. This can be affected by imposing sudden changes in
direction, such as the impact at the base of a free overfall or on baffles, or
through sudden expansions, particularly the hydraulic jump.
2. Discharge as a free jet. Flow may be directed as a free jet into a plunge pool or
on to a concrete apron, thereby causing an impact and change in direction, as
described above. The structure exit can be designed to promote atomization of
the flow and the water drops are then retarded by air resistance before impacting
on any surface.
There are many ways in which these basic approaches can be used in practice,
and various structures have been developed by intuition and model studies. In most
cases, theoretical analysis is not possible and design is carried out by using standard
geometries based on empirical results, or by using case-specific model studies. The
hydraulic jump is a very important energy dissipation feature and it can be analysed
to some extent.

© Springer Nature Switzerland AG 2020 183


C S James, Hydraulic Structures,
https://doi.org/10.1007/978-3-030-34086-5_6
184 6 Energy Dissipation Structures

6.2 The Hydraulic Jump

A hydraulic jump is the sudden change from (upstream-controlled) supercritical


flow to (downstream-controlled) subcritical flow with a free surface. The rapid
expansion associated with this change causes a significant energy loss and so it is
often induced where energy dissipation is required. The design problem is to ensure
that the jump occurs where it is required, to ensure its satisfactory performance, and
to size the bounding channel geometry to contain it effectively. In terms of per-
formance, the jump must be stable and dissipate energy effectively. Designing the
local channel geometry requires knowledge of the change in flow depth across the
jump, the location and distance over which this will occur and, if possible, the shape
of the intervening water surface profile. These characteristics depend on the entry
and tailwater flow conditions and any design features used to control the jump.

6.2.1 Hydraulic Jump Characteristics

Even for the simplest situation of an unaided hydraulic jump on a plane, horizontal
surface, its form, behaviour and effectiveness for energy dissipation, all vary with
the approaching flow condition and distinctly different types are recognized. These
can be classified according to the value of the Froude number on the upstream side,
Fr1 (Chow 1959).
The limiting case is obviously for Fr1 = 1.0, for which there can be no jump.
For 1.0 < Fr1 < 1.7, an undular jump will form. This has no distinct broken
wavefront, and flow changes from supercritical to subcritical through a series of
unbroken standing waves. The energy difference between the upstream and
downstream flows is small and is not all ‘lost’; some are propagated downstream in
the wave train. A broken wavefront will form if Fr1 exceeds about 1.25, but the
actual value depends on the roughness of the surface, being lower for a smooth
surface than for a rough one.
For 1.7 < Fr1 < 2.5, a weak jump forms. The downstream water surface is
fairly smooth, the velocity distribution is quite uniform across the depth and the
energy loss is less than about 20% of the upstream value.
For 2.5 < Fr1 < 4.5, an oscillating jump forms. The upstream flow enters the
jump as a jet which oscillates between the bottom and the water surface with no
regular period. The resulting pulsating type of flow produces irregular surface
waves which can travel long distances downstream.
For 4.5 < Fr1 < 9.0, the jump is steady. It is stable, free from surface waves and
energy dissipation is effective, in the range of 45–70% of the incoming value.
For Fr1 > 9.0, a strong jump forms. Energy dissipation is effective, up to 85%,
but the behaviour is unsatisfactory. The incoming velocity is very high and the
difference between upstream and downstream water depths is large. The fast jet
intermittently entrains slugs of water at the face of the jump, causing waves and a
6.2 The Hydraulic Jump 185

rough water surface downstream. The water surface becomes increasingly rough as
Fr1 increases and is usually unacceptable for values greater than about 13.
All of these jump types are encountered in energy dissipation design, and some
require the use of structural features to ensure satisfactory performance. The weak
jump requires no such features and the design must just ensure an appropriate
downstream water level and provide for the length of the jump, which is relatively
short.
The oscillating jump presents wave problems. Baffle blocks and other commonly
used appurtenances have little effect on waves. The problem is generally not serious
for wide flows, such as on aprons below dams, but can be serious in canals, where
the width is small and waves can travel long distances, causing damage to unlined
banks and riprap and surging at measuring devices. Variations in channel geometry
and alignment cause reflections which can dampen or intensify waves. Special
types of stilling basin (such as the United States Bureau of Reclamation (USBR)
Basin IV described in Sect. 6.3) are required for this condition or wave suppressors
can be used. Peterka (1978) recommends two wave suppressor designs.
The raft-type wave suppressor, shown in Fig. 6.1, is the most effective type
when additional submergence is not possible. The rafts are porous panels set at
fixed heights and spacing. Surges over the rafts are dissipated by the flow through
the holes. In experiments, the wave height was found to be reduced by about 50%
by each raft. The rafts are set with their top surfaces at the mean water surface and
are made deep enough so that the wave troughs do not separate from the underside.
The ratio of open area to total area of raft is in the range of 1:6–1:8. The spacing
between rafts must be at least three times the raft width.
The underpass-type wave suppressor, shown in Fig. 6.2, is the most effective
type, but is more expensive than the raft type. It consists of a horizontal roof with a
vertical headwall high enough to force all the flow to pass beneath the roof. The
height of the roof may be set to accommodate a wide range of water levels and

w 3w min

Fig. 6.1 Raft-type wave suppressor (adapted from Peterka 1978)


186 6 Energy Dissipation Structures

Fig. 6.2 Underpass-type wave suppressor

wave heights. The degree of wave suppression for any particular roof height is
determined by the length of the roof.
Peterka (1978) provides some design recommendations for the underpass-type
wave suppressor, but these are based on limited experimental results. Practical
design of either the raft or underpass structures usually requires specific model
studies.
The steady jump is the most desirable and wherever possible the incoming flow
should be induced within the appropriate range of Froude number. Baffles and sills
are sometimes incorporated in stilling basins for this condition to shorten the length
required and to stabilize the jump position. This is particularly important where the
discharge is not precisely known, or is expected to vary. The effects of these
features are discussed below and their incorporation in standard designs is pre-
sented in Sect. 6.3.
For Fr1 greater than about 10, the jump becomes very sensitive to tailwater depth
and a tailwater greater than the conjugate of the inflowing depth is advisable to
ensure that the jump is kept within the stilling basin. For such high Froude numbers,
the difference between the conjugate depths is large and high training walls are
required. For high spillways, stilling basins are therefore not usually economically
competitive with other types of energy dissipators, such as buckets or splitters.
Analysis of the hydraulic jump for design purposes is necessary to enable pre-
diction of the sequent depth ratio, the energy loss, the location and length of the
jump, and the water surface profile through the jump. These issues have been the
subject of many experimental investigations over many years. Comprehensive
reviews of progress have been provided by Carollo et al. (2007, 2009), Hager et al.
(1990) and Schulz et al. (2015).
The sequent depth ratio can be predicted by applying conservation of momentum
and continuity. The upstream (1) and downstream (2) flow conditions are related by

P
M1  M2 ¼ ð6:1Þ
c

in which M is the momentum function, P is the force applied to the flow between
the sections in the upstream direction and c is the unit weight of water.
For a simple hydraulic jump (with no features to apply a force to the flow) on a
horizontal bed, and assuming the effect of boundary shear through the jump to be
6.2 The Hydraulic Jump 187

Fig. 6.3 Simple hydraulic Lj


jump on horizontal bed
Lr

y y2

y1 x

1 2

negligible, Eq. (6.1) can be developed to obtain a relationship between the


upstream and downstream (sequent) flow depths (see Fig. 6.3).
In this case

M1 ¼ M2

which, for rectangular sections, can be written as

q2 y2 q2 y2
þ 1¼ þ 2
gy1 2 gy2 2

By substituting q = V1 y1, this can be developed to give


qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
y2 1
¼ 1 þ 8Fr1  1 2
ð6:2Þ
y1 2

and by substituting q = V2 y2,


qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
y1 1
¼ 1 þ 8Fr2  1 2
ð6:3Þ
y2 2

Hager and Bremen (1989) found that the sequent depth ratio is also influenced
by the inflow Reynolds number (Re1 = V1y1/m, where V1 is the inflow velocity and
m is the kinematic viscosity). They proposed correction to Eq. (6.3) in terms of Fr1,
Re1 and y1/b where b is the channel width, but the error in ignoring this effect is
unlikely to exceed about 3%.
The neglect of boundary shear in the development of Eqs. (6.2) and (6.3) results
in exaggerated estimates of y2/y1 (e.g. Hughes and Flack 1984; Corollo et al. 2007,
2009; Schulz et al. 2015). The ratio y2/y1 reduces consistently with increasing bed
roughness, the effect becoming more pronounced with increasing Fr1. Various
formulations have been proposed for prediction, all based on laboratory
188 6 Energy Dissipation Structures

information. For example, Corollo et al. (2007) proposed that Eq. (6.2) be adjusted
according to
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
y2 1
¼ 1 þ 8ð1  bÞFr21  1 ð6:4Þ
y1 2

with
 0:75 !
2 ks
b ¼ arctan 0:8 ð6:5Þ
p y1

in which ks is d50 size of the bed roughness particles. Equation 6.5 was formulated
and tested with data for ks up to 32 mm, ks/y1 in the range 0 to 2.025 and Fr1 up to
about 10.
Once the flow conditions on either side of the jump are known, the energy loss
can easily be found as the difference between the specific energy values. By solving
the energy and momentum relationships simultaneously, the following general
expression for energy loss (DE) can be derived:

ð y2  y1 Þ 3
DE ¼ ð6:6Þ
4y1 y2

Peterka (1978) presents experimental results which agree very closely with the
predictions of Eq. (6.6). The energy loss varies from a negligible amount for Fr1
less than about 2 up to around 85% for Fr1 = 20.
The length of a hydraulic jump is an important design parameter because it
determines the size of stilling basin required or, in the absence of a basin, the
location beyond which bed protection is not required. The length cannot be pre-
dicted theoretically, and many experimental studies have been carried out to for-
mulate empirical prediction relationships (e.g. Peterka 1978; Mehrotra 1976; Hager
et al. 1990; Carollo et al. 2007, 2012; Schulz et al. 2015). The length of a hydraulic
jump (Lj) is difficult to measure because the end is not distinct and specifying its
location involves some degree of subjectivity; the end of the jump may be assumed
to be where the flow depth reaches its maximum or where the flow depth becomes
approximately uniform. The length of the roller (Lr) may be more objectively
defined and measured, but the difference is not great for practical purposes. The
length of the jump, as well as the sequent depth, is reduced by bed roughness
(Hughes and Flack 1984; Carollo et al. 2007).
In most prediction formulations, the length is normalized in terms of the
upstream flow depth and expressed as a function of the upstream Froude number
(Fr1) or the ratio of sequent depths (y2/y1). Peterka (1978) used a large data set to
produce a graphical relationship for Lj/y1(Fr1) for a wide range of Fr1 (Fig. 6.4); this
can be approximated by
6.2 The Hydraulic Jump 189

180
160 Peterka (1978)
140
Equation (6.8)
120
Lr /y1 or Lj /y1

100 Equation (6.9)


80
60 Equation (6.10)

40
Equations (6.11) &
20 (6.13)
0 Equation (6.12)
0 5 10 15 20
Fr1

Fig. 6.4 Comparison of predictions of jump or roller length on a smooth bed

Lj
¼ 0:13Fr21 þ 11Fr1  12 ð6:7Þ
y1

Hager et al. (1990) proposed different relationships for Lr/y1 for different ranges
of Fr1. For 2.5 < Fr1 < 8

Lr
¼ 8ðFr1  1:5Þ ð6:8Þ
y1

For Fr1 > 8, they found the jump length to depend also on the aspect ratio,
y1/b. For y1/b < 0.10
 
Lr Fr1
¼ 12 þ 160 tanh ð6:9Þ
y1 20

and for y1/b > 0.10


 
Lr Fr1
¼ 12 þ 100 tanh ð6:10Þ
y1 12:5

Using their own data together with those of Hughes and Flack (1984) and Hager
et al. (1990), Carollo et al. (2007) proposed three alternative formulations, i.e.
 
Lr y2
¼ 4:616 1 ð6:11Þ
y1 y1

Lr 2:244
¼  1:272 ð6:12Þ
y1 y1
y2
190 6 Energy Dissipation Structures

both of which require prior adjustment for bed roughness if required and explicitly
accounting for bed roughness,
  
Lr ks
¼ 6:525 exp 0:60 ðFr1  1Þ ð6:13Þ
y1 y1

Equations (6.8) to (6.13) for smooth beds (ks = 0) are plotted together with
Peterka’s (1978) curve in Fig. 6.4. Peterka’s curve indicates longer jumps than the
others, mainly because of his different definition of length (Hager and Li (1992)
have also suggested that Lj/Lr is approximately equal to 4/3). The Hager et al. and
Carollo et al. (6.11) and (6.12) relationships give similar predictions for Fr1 up to
about 16, with Carollo et al. (6.12) departing from these for Fr1 greater than about
10. The difference between Eqs. (6.9) and (6.10) of Hager et al. shows the effect of
channel width to be considerable and increasing with Fr1 above 8.
Relationships for Lj/y2 = f(Fr1) have also been proposed (e.g. Peterka 1978;
Mehrotra 1976; Hughes and Flack 1984). These show Lj/y2 to be approximately
constant for Fr1 in the desirable range of 4.5–9. The data of Peterka (1978) (see also
Fig. 6.14) show Lj/y2  6.1 in this range, while the relationships of Hager et al.
(1990) (Eqs. (6.8) and (6.9)) and Carollo et al. (2007) (Eqs. (6.11) and (6.13))
suggest Lr/y2  4.1.
Knowledge of the water surface profile through the jump is also important in
stilling basin design because it will determine the height of the retaining walls.
Valiani (1997) produced an analytical solution describing the surface profile, i.e.

 3 ! !1=3
y y2 x
¼ 1þ 1 ð6:14Þ
y1 y1 Lj

in which y is the water depth within the roller, x is the distance from the entrance to
the roller and Lj is the length of the jump (the distance between the sequent depths),
assumed to be the same as Lr (Fig. 6.1). This relationship was shown by
Castro-Orgaz and Hager (2009) to be supported by data for Fr1 in the range of about
2.5 to about 9, and considered by them to give reasonable results for practical
purposes. Chow (1959) presented a dimensionless graph for describing the profile
based on the experimental results obtained by Bakhmeteff and Matzke (1936). This
shows a similar trend with Fr1 as Eq. (6.14), but a more gradual initial rise of the
surface profile for all values of Fr1.
The jump occurs at the location where the sequent depth relationship is satisfied
(Eqs. (6.2) or (6.3)). However, the length of the jump between the sequent depths
may need to be accounted for if accurate location is required. The flow depth on at
least one side of a hydraulic jump will be nonuniform, and its profile must be
computed to enable the jump to be located. If the downstream (subcritical) flow is
uniform, the jump will begin at the location on the upstream (supercritical) profile
where the sequent depth of the uniform flow occurs. The jump length can then be
calculated from this depth according to one of the above relationships to define the
6.2 The Hydraulic Jump 191

M3 sequent

A Lj
M2

y2

M3 B y1

Fig. 6.5 Location of a hydraulic jump between nonuniform profiles

location of the end of the jump. If the upstream (supercritical) flow is uniform, then
the jump will end at the location on the downstream (subcritical) profile where the
sequent depth of the uniform flow occurs. The jump length is then calculated using
the upstream, uniform, flow depth to locate the beginning of the jump. If the flow is
nonuniform both upstream and downstream, then both nonuniform profiles must be
computed, as well as a profile of the sequent depths of one of them. An initial
estimate of the jump location is then given by the intersection of one of the water
surface profiles with the sequent profile of the other. The upstream flow depth at
this location enables a jump length to be calculated, which is then fitted between the
water surface profile and the sequent profile. This will define a location different
from the initial estimate, so the jump length calculation may need to be refined.
Figure 6.5 shows an example for a jump between an M3 profile and an M2 profile
(as discussed by Chow 1959). The intersection of the M2 profile with the sequent
M3 profile (point A) defines a supercritical flow depth on the M3 profile (point B)
for calculating the jump length (Lj). The jump length is then fitted between the M3
sequent curve and the M2 profile to locate the sequent depth pair at the beginning
and end of the jump.
Example 6.1
Water is released from a storage facility into a long, 4.5 m wide, rectangular,
concrete-lined channel on a slope of 0.00040. The design discharge is 15.0 m3/s,
resulting in a velocity at the entrance to the channel of 12.0 m/s.
Determine

a. the flow depths immediately upstream and downstream of the hydraulic jump in the
channel,
b. the energy dissipated in the jump,
c. the location of the jump in the channel,
d. the length of the jump and roller and
e. the water surface profile through the jump.
192 6 Energy Dissipation Structures

Solution

Q = 15 m3/s W = 4.5 m
V = 12 m/s

So = 0.00040
1

Determine the flow depth at section 1 from the continuity equation,

Q ¼ AV
Q ¼ 15 m3 =s
A ¼ Wy1 ¼ 4:5y1
V ¼ 12 m/s

Therefore 15 ¼ 4.5y1  12
so
15
y1 ¼ ¼ 0:278 m
4:5  12
Determine the uniform flow depth in the channel from the Manning and continuity
equations,

A 2=3 1=2
Q¼ R So
n
A ¼ Wyo
n ¼ 0:013 for concrete
A Wyo
R¼ ¼
P W þ 2yo
So ¼ 0:00040

Therefore
 2=3
Wyo Wyo
Q¼ 1=2
3So
n W þ 2yo

i.e.
 2=
4:5 yo 4:5 yo 3 1
15 ¼ 0:00040 =2
0:013 4:5 þ 2 yo

from which
yo ¼ 2:06 m

Determine the sequent depth of y1 from Eq. (6.2),


6.2 The Hydraulic Jump 193

qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
y2 1
¼ 1 þ 8Fr21  1
y1 2
V 12
Fr1 ¼ pffiffiffiffiffiffiffi ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ 7:27
gy1 9:8  0:278

Therefore

1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2 
y2 ¼ 0:278 1 þ 8  7:27  1
2
¼ 2:72 m

a. y*2 > yo, so the hydraulic jump is some distance along the channel, and at the jump, the
subcritical flow will be yo and the supercritical flow will be its sequent depth, given by
Eq. (6.3), i.e.
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
yo sequent 1
¼ 1 þ 8Fr2o  1
yo 2
Vo 15=ð4:5  2:06Þ
Fro ¼ pffiffiffiffiffiffiffi ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ 0:360
gyo 9:8  2:06

Therefore

1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2ffi 
yo sequent ¼ 2:06 1 þ 8  0:360  1
2
¼ 0:440 m

b. The energy dissipated by the jump can be calculated as the difference between the specific
energy (E) values before and after it, or from Eq. (6.6) which was derived from this, i.e.
 3
yo  yo sequent
DE ¼
4  yo sequent  yo
ð2:060  0:440Þ3
¼
4  0:440  2:06
¼ 1:173 m

c. The jump will be located where the sequent of the uniform flow depth occurs on the M3
profile from the entrance to the channel,

M3 yo = 2.060 m
yo sequent
y1 = 0.278 m
= 0.440 m
x

The distance x can be found using the Direct Step Method, i.e. by solving
194 6 Energy Dissipation Structures

DE
Dx ¼
So  Sf

sequentially for increments of y between y1 and yo sequent. Sf is the average energy


gradient for the increment, as calculated for each location from the Manning equa-
tion, i.e.

V 2 n2
Sf ¼
4
R =3
Using 10 flow depth increments, x is found to be 54 m. (Similar to Example 1.8.)

d. The length of the jump or roller can be related to the supercritical


Froude number by Eqs. (6.7)–(6.12).

Vyo sequent
Fryo sequent ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

g yo sequent
Q
Vyo sequent ¼
W yo sequent
15
¼
4:5  0:440
¼ 7:58 m/s

Therefore

7:58
Fryo sequent ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
9:8  0:440
¼ 3:65

The hydraulic jump length, by Eq. (6.7), is therefore

Lj
¼ 0:13 Fr2yo sequent þ 11 Fryo sequent  12
yo sequent

Therefore
 
Lj ¼ 0:440 0:13  3:652 þ 11  3:65  12
¼ 11:6 m

Figure 6.4 indicates that for Fr1 = 3.65, all the equations for the roller length ((6.8)
to (6.12)) give similar results. Using Eq. (6.8), for example,

Lr  
¼ 8 Fryo sequent  1:5
yo sequent

Therefore
6.2 The Hydraulic Jump 195

Lr ¼ 0440  8ð3:65  1:5Þ


¼ 7:6 m

e. The water surface profile through the jump is given by Eq. (6.14), i.e.

 3 ! !1=
y yo x 3
¼ 1þ 1
yo sequent yo sequent Lj

Using Lj = Lr = 7.6 m as calculated above, the water surface profile is as shown


below:

2.5

2.0
Flow Depth (m)

1.5

1.0

0.5

0.0
-2 0 2 4 6 8 10
Distance (m)

6.2.2 Controlled Hydraulic Jumps

A hydraulic jump occurs if the upstream supercritical flow and downstream sub-
critical flow satisfy Eq. (6.1). The upstream conditions are controlled by the dis-
charging structure and are usually not able to be manipulated for the design.
Ensuring a jump therefore requires that the subcritical flow conditions, controlled
from further downstream, are appropriate. For the simple hydraulic jump in a
horizontal channel, the required downstream depth is given by Eq. (6.2), and is
denoted y*2 hereafter. If the existing downstream conditions do not meet this
requirement, then additional measures must be designed to ensure the jump takes
place where required. The different possible conditions, using an ogee spillway
structure, for example, are illustrated in Fig. 6.6.
The ideal situation is where the tailwater (yt) is exactly equal to y*2, as shown in
Fig. 6.6a. The momentum equation is satisfied and the jump forms immediately and
unaided. This condition is unlikely if water is being released into a natural channel,
196 6 Energy Dissipation Structures

yt = y2*
y1

(a) yt = y2*

y2* yt yt

(b) yt < y2*

y2* yt yt
yt

(c) yt > y2*

Fig. 6.6 Effect of tailwater on hydraulic jump formation

but could possibly be contrived for a fixed discharge in an artificial canal by


designing the geometry appropriately. In practice, there may be inaccuracy in some
design assumptions and some feature (such as a row of baffle blocks) is usually
installed to control the jump position and ensure that it does not move downstream.
If the tailwater is less than y*2, an unaided jump would occur some distance
downstream, as shown in Fig. 6.6b. The supercritical flow would assume an M3
profile, in which the flow depth increases downstream until it reaches the sequent
depth of yt, where the jump occurs. The tailwater level can be increased using a
weir, sill or an upward step created by excavating a basin just downstream of the
structure. As indicated by Eq. (6.1), any feature that imposes an upwardly directed
force against the flow reduces the downstream momentum function value, and
hence flow depth required for a jump to occur.
If the tailwater is greater than y*2, the hydraulic jump is forced upstream and may
be submerged, as shown in Fig. 6.6c. Although this is a safe condition because the
position is easily fixed, the energy dissipation is less effective than for an unsub-
merged jump. An unsubmerged jump can be ensured by raising the level of the
apron, thereby creating a downward step to the tailwater. Another design option for
the case where the tailwater is greater than the required conjugate depth is to install
a sloping entrance to the stilling basin or apron. The jump will then advance up the
slope to the position where momentum balance is achieved.
6.2 The Hydraulic Jump 197

In practice, the relationship between the sequent depth required for jump for-
mation and the tailwater depth determined by downstream conditions may vary with
discharge. It would be highly unlikely for yt to be equal to y*2 over the whole range of
expected discharges. It would be quite common, however, for yt to be either greater
or less than y*2 over the entire range. In these cases, the designs discussed above
would be satisfactory for a wide range of discharges, although ideal only for the
particular value used for the design. It is possible, however, for the relationship
between yt and y*2 to reverse within the expected range of discharges. If y*2 > yt at low
discharges and y*2 < yt at high discharges, a combination of the designs described
above would be effective, i.e. the provision of a sill or upward step to raise the
tailwater at low discharges and a sloping apron to allow the advance of an unsub-
merged jump at high discharges. If y*2 < yt at low discharges and y*2 > yt at high
discharges, it is not possible to design a satisfactory configuration for all discharges.
The most practical approach would be to provide an upward step to raise the tail-
water at high discharges and accept a submerged jump at low discharges.
Commonly used features used to reduce the tailwater depth requirement are sills,
baffle blocks, abrupt upward steps and sloping approach surfaces. These features
add stability to the jump and also increase the energy dissipation. Jumps associated
with such features are classified according to their locations. In particular, Type A
jumps are wholly located upstream or downstream of the controlling feature and
Type B jumps extend across the feature. Analyses for these features are presented in
the following sections.
Sloping Surfaces
Many designs for energy dissipators induce hydraulic jumps on sloping surfaces, such
as at the ends of spillways or drop structures or sloping entrances to stilling basins.
A sloping entrance or apron assists in stabilizing the jump location and ensuring its
occurrence over a range of discharges. The design again requires knowledge of the
sequent depth ratio, the length of the jump and its location. The momentum analysis
for sequent depths is complicated considerably if the jump occurs in a channel with a
slope steep enough that gravitational forces need to be considered. Evaluation of the
forces acting on the free body is then difficult and empirical input is essential. The
weight component of the free body is difficult to estimate because the length and shape
of the jump are not well defined. A further complication is that significant aeration
occurs, changing the specific weight of the water, and hence affecting estimation of
both the free body weight component and the pressure forces.
A number of distinct situations are recognized for hydraulic jumps associated
with sloping channel beds (e.g. Chow 1959; Henderson 1966; French 1985), as
shown in Fig. 6.7. These situations are characterized by the position of the jump,
defined by its length, Lr, the distance of its beginning from the change in slope,
l and the relationship between the sequent depths, y1 and y2.
In the illustrated cases, yt is the tailwater depth, and y1 and y2 are the upstream
and downstream sequent depths. The sequent depth for y1 on a horizontal bed (i.e.
as given by Eq. (6.2)) is denoted as y*2. If the jump begins at or after the end of the
198 6 Energy Dissipation Structures

Lr

y2 = y2* Type A
y1
θ = yt

Lr
Lr
l y1
y1 y2 = yt Type E
Type B y2
θ
θ

Lr

Lr
y1 Type C
y2 = yt
y2
θ Type F

Lr y1

y1 y2 yt Type D
θ

Fig. 6.7 Types of hydraulic jumps in sloping channels (adapted from French 1985)

slope then y2 = y*2 = yt and the jump will be Type A. In this case, the usual analysis
and results apply. If yt is greater than that needed for a jump on a horizontal bed, the
jump will advance onto the slope and the jump will be Type B, C or D. Here the
slope introduces a weight component which enhances the upstream value of M, and
also provides a force to affect the change in flow direction. If yt is insufficient to
force the entire jump onto the slope, the jump is Type B. If yt is just great enough
for the end of the jump to coincide with the end of the slope, then the jump is
Type C. For greater values of yt, the jump will occur completely on the slope as
Type D. Ohtsu and Yasuda (1991) found that for these types if yt/y1 > 3.0 and
h > about 20° a surface roller does not form and the upstream supercritical flow
plunges directly into the subcritical flow. Type E jumps occur on long slopes with
no break to horizontal. Type F jumps are unusual, but may occur below some drop
structure designs.
The analysis for Type A jumps is straight forward, and the results already
presented for horizontal channels apply. No analytical solution is yet available for
jumps of Types B, C and D but useful empirically based relationships have been
developed.
Kindsvater (1944) developed the following equation for the sequent depth ratio
for jumps completely on the slope, i.e. Types C and D.
6.2 The Hydraulic Jump 199

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi !
y2 1 cos 3h
¼ 1 þ 8Fr1 2
1 ð6:15Þ
y1 2 cos h 1  2N tan h

in which N is an empirical factor related to the jump length and h is the channel
slope.
Equation (6.15) can be written as
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
y2 1
=
¼ 1 þ 8G21  1 ð6:16Þ
y1 2

in which y1′= y1/cos h, and with

G21 ¼ C21 Fr21 ð6:17Þ

and

cos3 h
C21 ¼ ð6:18Þ
1  2N tan h

N has been found to be primarily a function of h and Peterka (1978) presented a


graphical relationship between N and tanh for 5° < h < 16°, which can be repre-
sented by

N ¼ 0:938 lnðtan hÞ þ 0:258 ð6:19Þ

Rajaratnam (1967) related C1 directly to h as

C1 ¼ 100:027h ð6:20Þ

with h in degrees.
For the same situation, Ohtsu and Yasuda (1991) proposed
y2  
¼ 0:077 h1:27 þ 1:41 ðFr1  1Þ þ 1 ð6:21Þ
y1

which agrees well with Eqs. (6.16)–(6.20) over practically realistic ranges of Fr1
and h.
Various empirical formulations have been presented for the sequent depth ratio
for Type B jumps and have been reviewed and tested by Carollo et al. (2011).
Solution of all the equations requires knowledge of the location of the beginning of
the jump, in terms of its distance from the junction between the sloping and hor-
izontal beds (l) or the height of this point above the horizontal bed level. For most
design applications, this calculation is not necessary, however. The sequent depths
are determined by the upstream and downstream controls, and the position of the
200 6 Energy Dissipation Structures

jump adjusts to these. If the tailwater level necessary to hold a jump against a
specified inflow depth is required, a value can be found by trial that ensures a
suitable jump type and location. The jump location is defined by its length (Lj) and
the distance of its beginning from the slope junction (l).
Ohtsu and Yasuda (1991) proposed equations for the length of the jump for
4 < Fr1  14 and for two ranges of slope. For h < 19°

Lj
¼ 5:75 tan h þ 5:70 ð6:22Þ
y2

and for 19° < h < 60°


 
Lj yt
¼ 4:6   1 þ 5:7 ð6:23Þ
y2 y2

Equation (6.22) is in accordance with the graphical relationship of Peterka


(1978) over the specified range of Fr1, but Peterka’s graph indicates considerable
dependence on Fr1 beyond this range. He presented no data for slopes greater than
17°. Ohtsu and Yasuda (1991) showed Eq. (6.23) to represent roller length (Lr) and
jump length (Lj) data equally well.
Ohtsu and Yasuda (1991) proposed Eq. (6.24) for the distance l, applicable for
the full range of h up to 60°, 6  Fr1  14 and 1.1  yt/y*2  3.0.
! 0:75
l 2:3 yt
¼  0:8 1 ð6:24Þ
y2 ðtan hÞ 0:73 y2

This agrees well with the graphical relationship of Peterka (1978).


In these analyses, it is usually assumed that y1 will not vary along the slope over
the distance within which the hydraulic jump may be expected. This is a reasonable
assumption for relatively high head structures.
A suggested application of the relationships for describing the characteristics of
a hydraulic jump associated with a transition from a sloping to a horizontal channel
proceeds as follows:
– Calculate y1, yt and y*2. y1 is the flow depth upstream of the jump, and is
determined by the upstream control feature. yt is the tailwater depth, determined
from the downstream control; if this is a required quantity for the design, a trial
value is assumed and adjusted until a value resulting in a satisfactory condition
is produced. y*2 is the sequent depth of y1 as would occur on a horizontal surface
(i.e. for a Type A jump), given by Eq. (6.2).
– Compare yt with y*2. If yt < y*2, then a Type A jump will occur at some location
further downstream. If yt > y*2, then the jump will occur at least partially on the
slope.
– If yt > 3y*2, then there will be no distinct roller and the upstream supercritical
flow will plunge as a jet directly into the tailwater.
6.2 The Hydraulic Jump 201

– If yt < 3y*2, then an unsubmerged jump will occur. Calculate the length of the
jump (Lj) from Eq. (6.22) if h < 19° or Eq. (6.23) if 19° < h < 60°. Calculate
the location of the beginning of the jump (l) from Eq. (6.24).
– Compare Lj with l. If Lj > l, then the jump is Type B, and is now fully described.
If Lj = l, then the jump is Type C and is also now fully described. If Lj < l, then
the jump is Type D.
– For jump Type D, if h > 19°, then there will be no distinct roller and the
upstream supercritical flow will plunge as a jet directly into the tailwater. If
h < 19°, then the sequent depth (y2) is calculated from Eq. (6.21) or from
Eqs. (6.16) to (6.19). The end of the jump will be located where y2 occurs on the
subcritical S1 profile.
The energy loss is best determined by calculating the energy values upstream
and downstream of the jump relative to the same datum and taking the difference.
For Type A jumps, Eq. (6.6) may be used.
Example 6.2
The channel in Example 6.1 is modified to have an initial steep slope leading to a
horizontal reach. A control structure further downstream induces a flow depth of 4.30 m
in the vicinity of the slope transition.
Locate the hydraulic jump in the channel and determine the associated energy loss

a. if the initial slope is 0.25 (14°) and


b. if the initial slope is 0.10 (5.7°).

Solution

a. Slope = 0.25
Assume the jump begins close enough to the exit that the supercritical flow at the toe of
the jump is approximately the same as at the exit. The analysis could be refined if
necessary by computing the gradually varied profile before the jump.
Determine if the jump occurs on the horizontal (Type A) or sloping (Type B or C or D)
surface:
yt = 4.30 m (given),
y*2 = 2.72 m (from Example 6.1),

yt > y*2, so the hydraulic jump occurs on the slope.


Determine if the jump is submerged or unsubmerged:
yt < 3y*2 (4.30 < 3  2 .72 = 8.16 m), so an unsubmerged jump will occur on the
slope.
Calculate the length of the jump:
The slope of 0.25 (14°) is less than 19° so the jump length is given by Eq. (6.22),
i.e.

Lj
¼ 5:75 tan h þ 5:70
y2
202 6 Energy Dissipation Structures

So Lj ¼ 2:72ð5:75 tan 14 þ 5:70Þ


¼ 19:4 m

Calculate the location of the beginning of the jump, from Eq. (6.24), i.e.
! 0:75
l 2:3 yt
¼  0:8 1
y2 ðtan hÞ 0:73 y2

So
! 0:75
2:3 4:30
l ¼ 2:72  0:8 1
ð0:25Þ0:73 2:72
¼ 10:0 m

Lj > l, so the jump is Type B and the water surface profile is

19.4 m

10.0 m
0.28 m
4.30 m

14o

1 2

The energy loss can be determined by relating the total energy values at sections 1 and
2, with the datum selected at the horizontal bed level,

Loss ¼ H1  H2
   
V2 V2
¼ y1 cos h þ 1 þ l tan h  y2 þ 2
2g 2g
0  1
  15=ð4:5  4:30Þ2
12:02
¼ 0:28  cos 14 þ
o
þ 10:0 tan 14  @4:30 þ
o A
2  9:8 2  9:8

¼ 10:11  4:33
¼ 5:78 m

b. Slope = 0.10
As for the slope of 0.25, an unsubmerged jump will occur on the slope.
Calculate the length of the jump:
The slope of 0.10 (5.7°) is less than 19° so the jump length is given by Eq. (6.22), i.e.
6.2 The Hydraulic Jump 203

Lj
¼ 5:75 tan h þ 5:70
y2

So
Lj ¼ 2:72ð5:75 tan 5:7 þ 5:70Þ
¼ 17:1 m

Calculate the location of the beginning of the jump, from Eq. (6.24), i.e.
! 0:75
l 2:3 yt
¼  0:8 1
y2 ðtan hÞ0:73 y2

So
! 0:75
2:3 4:30
l ¼ 2:72 0:73
 0:8 1
ð0:10Þ 2:72
¼ 20:9m

Lj < l, so the jump is Type D. The sequent depth ratio is most easily obtained from
Eq. (6.21), i.e.

y2  
¼ 0:077h1:27 þ 1:41 ðFr1  1Þ þ 1
y1

So
 
y2 ¼ 0:278 0:0775:71:27 þ 1:41 ð7:27  1Þ þ 1
¼ 3:96 m

Alternatively, using Eq. (6.16),


qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
y2 1
=
¼ 1 þ 8G21  1
y1 2

with

= y1 0:278
y1 ¼ ¼ ¼ 0:279
cos h cos 5:7
with G21 given by Eqs. (6.17) and (6.20), i.e.

G21 ¼ C21 Fr21


C1 ¼ 100:027h ¼ 100:0275:7 ¼ 1:43

So
G21 ¼ 1:432 7:272 ¼ 108

Therefore
204 6 Energy Dissipation Structures

 
1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
y2 ¼ 0:279 1 þ 8  108  1
2
¼ 3:96 m

Alternatively, using Eq. (6.16) with G21 given by Eqs. (6.17), (6.18) and (6.19), i.e.

G21 ¼ C21 Fr21


cos3 h
C21 ¼
1  2N tan h

N ¼ 0:938 lnðtan hÞ þ 0:258


¼ 0:938 lnðtan 5:7 Þ þ 0:258
¼ 2:42

So
cos3 5:7
C21 ¼
1  2  2:42 tan 5:7
¼ 1:91

and

G21 ¼ 1:91  7:272 ¼ 101

Therefore
 
1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
y2 ¼ 0:279 1 þ 8  101  1
2
¼ 3:83 m

The three results for y2 are similar; accept y2 = 3.96 m.


The water surface profile is
20.9 m
17.1 m

3.96 m 4.30 m
0.28 m
5.7o

1 2

The energy loss can again be determined by relating the total energy values at Sects. 6.1
and 6.2, with the datum selected at the horizontal bed level,
6.2 The Hydraulic Jump 205

Loss ¼ H1  H2
   
V2 V2
¼ y1 cos h þ 1 þ l tan h  y2 þ 2
2g 2g
 2

12:0
¼ 0:278  cos 5:7 þ þ 20:9  tan 5:7
2  9:8
!
ð15=ð4:5  3:96ÞÞ2
 3:96 þ
2  9:8
¼ 9:71  4:00
¼ 5:71 m

Sills
A sill is a transverse, thin, vertical obstruction extending over the full flow width.
The effect of sills on hydraulic jump performance has been investigated by
(amongst others) Rajaratnam and Murahari (1971), Ranga Raju et al. (1980), Ohtsu
(1981), Ohtsu et al. (1996) and Hager and Li (1992).
Following from previous work, Hager and Li (1992) defined different types of
jumps associated with sills, depending on the height of the sill (s), its distance from
the toe of the jump (Ls) and the tailwater depth (y2) (Fig. 6.8). The Type A jump is
similar to the simple jump, with the sill located at the end of the surface roller. The
sill is relatively small and has little effect on the jump location or sequent depth
ratio. A Type B jump occurs with a slightly lower tailwater, allowing the jump to
move downstream and the deflection of the bottom current by the sill causing a
small standing wave just downstream of the sill. A further reduction in tailwater
depth allows a second surface roller to form. This is regarded as the limiting
condition for practical energy dissipation design, and is designated the Type Bm (B
minimum) jump. A Type C jump with a pulsating standing wave is also recognized,
but should be avoided in design. An extreme situation with a large disturbance
wave over the sill and flow returning to supercritical downstream can also occur.
Hager and Li (1992) (with some reliance on data provided by Bretz (1987))
proposed accounting for the effect of the sill on the sequent depth ratio by applying
an empirical correction to the ratio for the simple hydraulic jump (Eq. (6.2)), i.e.
y2
Y¼ ¼ Y   DYf  DYs ð6:25Þ
y1

in which Y* is the sequent depth ratio for a simple jump, as calculated by Eq. (6.2),
DYf is a correction to account for wall friction (if applicable) and DYs is a correction
to account for the effect of the sill. The correction DYs is given by

DYs ¼ 0:7 S0:7 þ 3Sð1  KÞ2 for K [ 0:5 ð6:26Þ

in which S = s/y1, K ¼ Ls =Lr . Lr is the roller length of a simple hydraulic jump
given by Eqs. (6.8) or (6.9) or (6.10). This equation shows that the sequent depth
206 6 Energy Dissipation Structures

y2 Type A
y1 s

Ls

y2
Type B
y1 s

Ls

y2 Type Bm
y1 s

Ls

Fig. 6.8 Hydraulic jumps controlled by sills

ratio decreases with increasing sill height and also depends on the distance between
the sill and the toe of the jump, the effect increasing with decreasing distance.
A typical design problem would be to determine the sill height and location to force
a jump to occur between known entry (y1) and tailwater (y2) depths. Because
Eq. (6.26) includes both the sill height and its location (Ls), one of these needs to be
specified for a design, and the other calculated from Eqs. (6.25) and (6.26) (DYf
would be negligible in most cases.) The sill height should be constrained by the
upper limit for appearance of a Type Bm jump, given by

1
Smax ¼   ð6:27Þ
9ð1  KÞ3

The design should be checked for the full range of expected discharges.
To maximize energy dissipation, the sill should be high and close to the toe of
the jump. Hager and Li (1992) suggest, however, that the potential for scour in the
tailwater region should also be considered. They recommend the Type A jump if
the tailwater channel is easily erodible and the Type Bm only if the tailwater channel
is highly resistant to scour.
6.2 The Hydraulic Jump 207

A bottom roller occurs downstream of a sill and Hager and Li (1992) suggest
including its length within the stilling basin. Using Bretz’s (1987) data for the
extent of the bottom roller, they proposed the length of the basin (LB) to be given by
 
LB 4 1=
¼ 1  0:6 S 3 ð1  K Þ ð6:28Þ
Lr 3

Example 6.3
For the situation described in Example 6.1, determine the height and location of a sill to
induce a hydraulic jump near the outlet.
Solution
A hydraulic jump near the outlet without a sill would require a tailwater depth to give a
sequent depth ratio according to Eq. (6.2), i.e.
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 
y2 1
Y ¼ ¼ 1 þ 8 Fr21  1
y1 2
1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2 
¼ 1 þ 8  7:27  1 ¼ 9:79
2
The actual tailwater (y2) is the uniform flow depth (=2.06 m, from Example 6.1), which
is less than y*2. The amount by which the sill must reduce the required tailwater, DYs
(=Dy/y1), is given by Eq. (6.25), i.e.

DYs ¼ Y   Y

with Y = y2/y1, and assuming the effect of friction within the jump to be negligible.
Therefore

2:06
DYs ¼ 9:79  ¼ 2:38
0:278
Equation (6.26) relates DYs to the step height and distance,

DYs ¼ 0:7 S0:7 þ 3 Sð1  KÞ2 for K [ 0:5

with

s

y1

and

Ls

LR

From Eq. (6.8)


208 6 Energy Dissipation Structures

LR ¼ y1 8ðFr1  1:5Þ


¼ 0:278  8ð7:27  1:5Þ ¼ 12:8 m

An appropriate design can be obtained by choosing a sill height (s), solving Eq. (6.26)
for K and checking the limit for K and the maximum sill height for the onset of a Bm
jump, according to Eq. (6.27),
1
Smax ¼
9 ð1  KÞ3

The smallest sill height that satisfies both limiting conditions is s = 0.78 m, for which

s 0:78
S¼ ¼ ¼ 2:81
y1 0:278

which gives, from Eq. (6.26),

Ls
K¼ ¼ 0:79
LR

So

Ls ¼ 0:79  12:8 ¼ 10:1 m

and
1
Smax ¼ ¼ 12
9ð1  0:79Þ3

which is greater than the 2.81 m selected.


The length of basin required is given by Eq. (6.28), i.e.
 
LB 4 1=
¼ 1  0:6 S 3 ð1  K Þ
LR 3
So
 
4 1
LB ¼ 12:28  1  0:6  2:81 =3 ð1  0:79Þ
3
¼ 12:3 m

The water surface through the jump is then

y2 = 2.06 m
s = 0.78 m
y1 = 0.278 m

Ls = 10.1 m

LB = 12.3 m
6.2 The Hydraulic Jump 209

Baffle Blocks
Ranga Raju et al. (1980) considered the effect of a row of baffle blocks in com-
bination with a sill. Their results for baffle blocks can be extracted to describe their
effect independently. From dimensional consideration, they showed that
 
FB Lb h
¼ f 1 ; ; Fr 1 ; g ð6:29Þ
F2 y1 y1

in which FB is the force on the blocks per unit width, F*2 is the hydrostatic force
associated with the sequent depth for a simple jump (i.e. 0.5 cy*2
2 ), Lb is the distance
between the toe of the jump and the front of the block, h is the block height and η is
the blockage ratio (=w/(w + z), where w is the block width and z is the spacing
between blocks). Equation (6.29) was expressed as
 
W1 W2 FB Lb
¼ f2 ð6:30Þ
F2 y1

with W1 accounting for h/y1 and W2 accounting for η.


The force on the blocks is related to the upstream and downstream conditions by
Eq. (6.1), which can be expressed as
 2   
FB y1 y1
¼ 1 þ 2Fr21  2Fr21 1 ð6:31Þ
F2 y2 y2

in which F2 is the hydrostatic force per unit width at the downstream section (i.e.
0.5 cy22).
Using data from Basco and Adams (1971) and Murahari (pers. comm. to Ranga
Raju), Ranga Raju et al. (1980) developed relationships for Eq. (6.30) and for W1
and W2. These were presented graphically, but can be approximated by the fol-
lowing equations:
  
W1 W2 FB Lb
¼ 0:56 exp 0:045 ð6:32Þ
F2 y1

for trapezoidal blocks (such as the standard shape in USBR stilling basins), and
  
W1 W2 F B Lb
¼ 0:52 exp 0:054 ð6:33Þ
F2 y1

for rectangular blocks (as used by Murahari), with


 0:7
h
W1 ¼ 3:08 ð6:34Þ
y1
210 6 Energy Dissipation Structures

and

W2 ¼ 0:33 g0:99 ð6:35Þ

For a design application, the location and geometry of baffles would be required
to hold a jump between the upstream flow depth (y1) and the tailwater depth (y2)
which is less than the sequent depth for a simple jump (y*2). The force on the baffles
can first be calculated from Eq. (6.31) and the simple hydraulic jump sequent depth
from Eq. (6.2). Then two of the three design parameters (h, η and Lb) must be
specified and the third determined through Eqs. (6.32) to (6.35); Lb should be well
within the undisturbed jump length. (Note that for standard stilling basins, USBR
uses η = 0.5, i.e. w = z.) Ranga Raju et al. (1980) observed that a second row of
baffle blocks had little effect on the drag force, and the relationship represented by
Eq. (6.32) applied equally well to this situation. An additional row of baffles may
assist to even out the velocity distribution within the jump and reduce downstream
scour, but does not further reduce the sequent depth ratio.
Example 6.4
For the situation described in Example 6.1, determine the dimensions and location of a
row of baffle blocks to induce a hydraulic jump near the outlet,
Solution
The force on the baffles per unit width (FB) can be calculated from the upstream and
tailwater conditions by Eq. (6.31). FB is then related to the block height (h), blockage
ratio (η) and distance between the toe of the jump and the baffles (Lb) by Eqs. (6.32),
(6.34) and (6.35). Two of these variables must be selected and the third calculated from
the equations.
Calculate FB by Eq. (6.31):
 2   
FB y1 y1
¼ 1 þ 2Fr21  2 Fr21 1
F2 y2 y2
F2 ¼ 0:5 c y22
¼ 0:5  9:8  103  2:062 ¼ 20:8  103 N/m width
Fr1 ¼ 7:27 from Example 6:1

Therefore
     !
0:278 2 2 0:278
FB ¼ 20:8  10 3
1 þ 2  7:27  2  7:27
2
1
2:06 2:06
¼ 14:2  103 N/m width

Lb and η will be chosen and h calculated.


Ranga Raju et al. (1980) recommend that Lb should be well within the undisturbed jump
length (L*r ); assume Lb/L*r = 0.30. (Note that h decreases with Lb, so a small value is
preferable.)
6.2 The Hydraulic Jump 211

For their stilling basins, USBR recommends η = 0.50; assume this value.
Calculate the left-hand side of Eq. (6.32) for USBR standard shape blocks,
  
w1 w2 FB Lb
¼ 0:56 exp 0:045
F2 y1
Lb
¼ 0:30 ðchosenÞ
LR
LR ¼ 12:8 ðfrom Example 6:3Þ

So

Lb ¼ 0:30  12:8 ¼ 3:84 m

and
  
w1 w2 FB 3:84
¼ 0:56 exp 0:045 ¼ 0:300
F2 0:278

Calculate W2 for η = 0.50 from Eq. (6.35), i.e.

w2 ¼ 0:33 g0:99
¼ 0:33  0:500:99 ¼ 0:655

and
F2 ¼ 0:5c y2
2
y2 ¼ 2:72 m ðfrom Example 6:1Þ

So

F2 ¼ 0:5  9:8  103  2:722 ¼ 36:3  103 N/m width

w1 w2 FB
Now calculate W1 from ¼ 0:300 as calculated above, i:e:
F2

0:300 F2 0:300  36:3  103


w1 ¼ ¼ ¼ 1:17
w2 FB 0:655  14:2  103

Calculate h from Eq. (6.34), i.e.


 0:7
h
w1 ¼ 3:08
y1

So

 1=   1
w1 0:7 1:17  =0:7
h¼ y1 ¼ 0:278 ¼ 1:11 m
3:08 3:08
212 6 Energy Dissipation Structures

For their stilling basin baffles, USBR recommends the spacing and width to be 0.75
times the baffle height; accept this recommendation, so width and spacing will be 0.75 
1.11 = 0.83 m.
The water surface through the jump is

y2 = 2.06 m
h = 1.11 m
y1 = 0.278 m

Lb = 3.8 m

Upward and Downward Steps


An upward (positive) step can be used to stabilize a hydraulic jump if the tailwater
depth is lower than that required for an unaided jump to form. In practice, this is
usually obtained by excavating the stilling basin to lower its bed relative to the fixed
level downstream.
Hager and Bretz (1986) classify jumps with positive steps as Type A if the step
coincides with the end of the roller, and Type B (with a lower tailwater) if the step
is within the length of the roller (Fig. 6.9).
The relationship between the sequent depths and the height of the step can be
described by Eq. (6.1) provided the force on the step can be quantified. Forster and
Skrinde (1950) tested Type A jumps and assumed a hydrostatic pressure distribu-
tion on the step, related to the flow depth immediately upstream of it. This
assumption led to overestimation of Y and they used their results to produce a
diagram for relating Y, Fr1 and S. Hager and Sinniger (1985) and Hager and Bretz
(1986) assumed that the water level before and after the step would be the same,
and the hydrostatic pressure force per unit width would then be equal to qg(y2 +
s/2)s. Equation (6.1) can then be written as
   2  
q2 y21 q y22 s
þ  þ ¼ y2 þ s ð6:36Þ
gy1 2 gy2 2 2

This was found to underestimate the force, which is not actually hydrostatic and
includes a dynamic contribution associated with the jet beneath the jump. Better
agreement with Forster and Skrinde’s (1950) and Hager and Bretz’s (1986) data
was found by assuming a uniform pressure distribution over the step height, cal-
culated from the total depth; the unit width force is then equal to qg(y2+s)s, and
Eq. (6.1) becomes
   
q2 y2 q2 y2
þ 1  þ 2 ¼ ðy2 þ sÞs ð6:37Þ
gy1 2 gy2 2
6.2 The Hydraulic Jump 213

Lr

y2 Type A

y1 s

Lr

Type B
y2

y1 s

Fig. 6.9 Hydraulic jumps with upward step

This assumption produces better agreement with Hager and Bretz’s (1986) data
for both Type A and Type B jumps, suggesting that a greater than hydrostatic force
on the step persists for jumps at least up to the Type A position. The force would
tend to hydrostatic for jumps located further upstream than Type A. Other more
complicated adjustments to the pressure distribution have been proposed by Fiuzat
(1986) and Karki and Mishra (1986) but provide inconsequential improvement.
Hager and Sinniger (1985) and Hager and Bretz (1986) expressed Eq. (6.37) as a
relationship between Fr1, s and the sequent depths as
 
Y ð Y þ SÞ 2 þ S2  1
Fr21 ¼ ð6:38Þ
2ðY  1ÞÞ

Application of Eq. (6.37) to Hager and Bretz’s (1986) data shows it to over-
estimate the required step height for given values of y1 and y2 by about 25%; its
application would therefore lead to a conservative design.
As for sills, the relationship expressed by Eqs. (6.37) and (6.38) would be
expected also to depend on the position of the jump, as the force on the step must
depart further from hydrostatic as the toe of the jump becomes closer to the step
(Fiuzat 1986). This effect is yet to be satisfactorily quantified, as the documented
experiments were performed with fixed toe locations. Hager and Bretz (1986)
positioned their Type B jumps with the step about halfway along the length of the
roller, which they showed to be about 4.25(y2 + s) for Type B jumps and about
4.75(y2 + s) for Type A jumps. Forster and Skrinde (1950) fixed the distance from
the toe to the step at 5(y2 + s) for all their experiments. As the results from these
different sources are quite consistent, the effect of the jump location on the sequent
depth ratio is probably not great. Knowing the location of the jump is important for
214 6 Energy Dissipation Structures

designing the basin length, however, and there is evidence (Fowler 2019) that the
distance of the toe of the jump from the step varies consistently with the step height.
It is important to note that the tailwater depth determined from downstream
conditions cannot always be assumed to occur immediately after the step, and the
possibility of sweep out by the incoming supercritical flow must be considered. If
the tailwater depth for a given step size or the step size for a given tailwater depth is
less than indicated by Eq. (6.37), the supercritical flow may sweep out past the step
and push the hydraulic jump a considerable distance downstream. To prevent
sweep-out, the subcritical depth just after the step must be greater than the sequent
depth of the sweeping out supercritical flow at that location. Hager and Sinniger
(1985) maintain that these conditions are similar for abrupt steps, so sweep-out is
unlikely. This is not so for smooth steps, however, and sweep-out is highly likely.
Moreover, with a smooth step, preventing sweep-out will induce a subcritical flow
depth before the step considerably greater than the sequent depth of the approaching
supercritical flow, pushing the hydraulic jump far upstream and most likely sub-
merging the upstream control. However, as the objective of a stilling basin is to
effect energy dissipation, a smooth step would be less appropriate anyway.
Example 6.5
For the situation described in Example 6.1, determine the height of an upward step to
induce a hydraulic jump near the outlet.
Solution
The step height can be calculated from the momentum equation with an assumption of
the pressure distribution on the step. Equation (6.37) applies for both Types A and B
jumps, i.e.

M1  M2 ¼ ðy2 þ sÞs
q2 y2
M1 ¼ þ 1
gy1 2
Q 15
q¼ ¼ ¼ 3:33 m3 =s/m
W 4:5
So

3:332 0:2782
M1 ¼ þ ¼ 4:11 m2
9:8  0:278 2
Similarly

3:332 2:062
M2 ¼ þ ¼ 2:67 m2
9:8  2:06 2
Therefore

4:11  2:67 ¼ 1:44 ¼ ðy2 þ sÞs

i.e.
6.2 The Hydraulic Jump 215

s2 þ y2 s  1:4 ¼ 0

Therefore
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ð2:06Þ ð2:06Þ2 4  1  ð1:44Þ

21
¼ 0:55 m OR  2:61 m

So

s ¼ 0:55 m

It is not possible to determine the position of the jump, so the basin length should
accommodate the length of at least a Type A roller. The minimum length is therefore

Lb ¼ Lr ¼ 4:75 ðy2 þ sÞ
¼ 4:75 ð2:06 þ 0:55Þ ¼ 12:4 m; say 15 m

A downward (negative) step can be used to stabilize the position of a hydraulic


jump if the tailwater depth is greater than required for an unaided jump to form.
This can be incorporated in a design by setting the apron or channel bed level at the
exit from a structure above the bed level of the downstream channel.
As for an upward step, the jump can occur at different locations relative to the
step, and different types are recognized (e.g. Hsu 1950; Rajaratnam and Ortiz 1977;
Hager and Bretz 1986) (Fig. 6.10). Similar to the situations with an upward step, a
jump is Type A if it occurs wholly upstream of the step and Type B if it extends
over the step. The transition from Type A to Type B as the tailwater level reduces
passes through an unstable sequence of standing wave and then upward and
downward oriented jets. Hager and Bretz (1986) define the minimum Type B jump
as the limiting type, where the jump is wholly downstream of the step with the
supercritical depth at the toe equal to that just before the step. For lower tailwater
depths, the jump would occur as a simple type further downstream. Design requires
analysis of Types A and minimum B as well as quantification of the standing wave
height to establish side wall heights.
For the Type A jump, the force on the step is equal to qg(y2 − s/2)s and Eq. (6.1)
is written as
   2  
q2 y21 q y22 s
þ  þ ¼  y2  s ð6:39Þ
gy1 2 gy2 2 2

or, in terms of Fr1,


 
Y ðY  SÞ2 1
Fr21 ¼ ð6:40Þ
2ð Y  1Þ
216 6 Energy Dissipation Structures

y2 Type A
y1
s

Transition
y2 profiles
y1
s

Type B
y1 y2
s

minimum
y1 Type B
y2 = y2*
s y1

Fig. 6.10 Hydraulic jumps with downward step

For the Type B jump, Hsu (1950) assumed the force on the step to be hydrostatic
as determined by the upstream flow depth, i.e. equal to qg(y1 + s/2)s, i.e.
    
q2 y2 q2 y2 s
þ 1  þ 2 ¼  y1 þ s ð6:41Þ
gy1 2 gy2 2 2

leading to
 
Y Y 2  ð 1 þ SÞ 2
Fr21 ¼ ð6:42Þ
2ð Y  1Þ

Hager (1985) found better agreement with data by assuming for Type B jumps
that the incoming flow is similar to a free jet. The pressure is then atmospheric at
the top of the step and hydrostatic over the step height; the force on the step is then
equal to qgs2/2 and Eq. (6.1) is written as
6.2 The Hydraulic Jump 217

   
q2 y2 q2 y2 s2
þ 1  þ 2 ¼ ð6:43Þ
gy1 2 gy2 2 2

or, in terms of Fr1,

Y ðY 2  S2  1Þ
Fr21 ¼ ð6:44Þ
2ð Y  1Þ

For the minimum B jump, the jet reaches the channel bed just after the step, with
the same flow depth as immediately upstream. The sequent depth ratio is therefore
as for a simple jump on a horizontal bed.
Hsu (1950) presented experimental data confirming the applicability of
Eqs. (6.39) and (6.41) for the sequent depth ratios for Type A and B jumps, and
particularly the conditions at the transition between them.
Hager and Bretz (1986) found the length of the roller to be approximately 3.5y2
for Type A jumps (decreasing slightly with Fr1) and about 4.25y2 for Type B jumps.
They accounted for variation with Fr1 in their response to the discussion by Ohtsu
and Yasuda (1987), but acknowledged the limited data available.
A downward step design should allow for the side walls to accommodate the
height of the standing wave in the transition between Type A and B jumps. Based
on data from Rajaratnam and Ortiz (1977) and their own experiments, Hager and
Bretz (1986) propose the maximum wave height to be given by

ðY þ SÞmax ¼ 2Fr1 ð6:45Þ

6.3 Standard Stilling Basins

The purpose of a stilling basin is to effect energy dissipation by means of a hy-


draulic jump within the basin, so that the high energy supercritical flow entering the
basin is converted to lower energy subcritical flow before reaching the downstream
channel. The aims of the design are to ensure formation of the jump, to keep it
stable in one position and to make it as short as possible.
Features commonly used to control hydraulic jumps were discussed in the
previous section, and these are usually used in combination for stilling basin design.
Just as the effects of the features acting individually were determined empirically,
so must their effects in combination. Original stilling basin design therefore
invariably requires physical modelling testing. Because many stilling basins need to
be designed for similar conditions, various general-purpose, standard designs have
been developed. In particular, the United States Bureau of Reclamation (USBR)
(Bradley and Peterka 1957; Peterka 1978) has proposed a number of designs for
specific applications and conditions. These are applicable for a wide range of
218 6 Energy Dissipation Structures

sloping floor
baffle piers
chute blocks
end sill
floor level

(a) Longitudinal section

chute baffle end sill


blocks piers

(b) Plan

Fig. 6.11 Features of USBR stilling basins

conditions, but care should be taken to comply with the specified restrictions;
otherwise, specific model studies should be carried out.
A typical USBR stilling basin will incorporate some or all of the features
illustrated in Fig. 6.11.
The sloping floor at the entrance assists in maintaining an unsubmerged jump
under a range of discharges, as described in the previous section. The floor level is
set to match the tailwater level to the downstream subcritical flow depth required for
jump formation. This should be checked for the full range of discharges expected,
with the tailwater level determined by the downstream channel characteristics.
The chute blocks are sometimes used to form a serrated entrance. The blocks
disperse the incoming flow by splitting it in jets and raising some of the flow off the
floor, as well as enhancing aeration. This has the effect of stabilizing the jump,
shortening its length and generally improving the performance.
The sill is placed at the end of the basin to ensure that a jump occurs and to
control its position over the whole operating range of discharges. It also protects the
channel bed from direct action of the current. Sometimes it is also used to set up a
reverse roller, which directs any mobilized bed material back towards the structure
to prevent undermining (Fig. 6.12a).

(a) (b)

scour hole

Fig. 6.12 Sill at end of USBR stilling basin


6.3 Standard Stilling Basins 219

For large basins with high incoming velocities, the sill is usually dentated, as
shown in Fig. 6.12b. This helps to diffuse the remaining part of the high-velocity jet
which may reach the end of the basin, resulting in a more uniform velocity profile,
less wave action and less likelihood of the intermittent entrainment and downstream
displacement of water masses in the jump.
Baffle piers are blocks placed on the basin floor. These assist in energy dissi-
pation by impact action. By exerting additional force on the control volume, they
reduce the downstream depth required for a jump to form, and therefore allow a
lower tailwater to be tolerated. Baffle piers are used most in small structures with
low velocities. They are unsuitable for high velocities because of the likelihood of
cavitation.
Three standard USBR stilling basin designs incorporating these features are
recommended by Bradley and Peterka (1957), Peterka (1978) and United States
Bureau of Reclamation (1987). They have been developed through analysis of
many successful designs as well as laboratory studies, as described in detail by
Peterka (1978).
The USBR Basin II (note that Basin I refers to a simple, unaided hydraulic jump)
is intended for use with high dam spillways and large canal structures, with high
incoming velocities and Froude numbers greater than 4.5. The recommended
geometry and dimensions, in terms of the inflow and sequent tailwater depths, are
shown in Fig. 6.13 and ensure safe, conservative design for spillways up to about

α
y1
y2*
0.02 y2*

h1 = y1 h2 = 0.2 y2*

LII
slope = 2:1
(a) Longitudinal section

0.5 y1
w1 = y1 s2 = 0.15 y2*
s1 = y1
w2 = 0.15 y2*

(b) Plan

Fig. 6.13 USBR Basin II geometry (adapted from Peterka 1978)


220 6 Energy Dissipation Structures

60 m high with discharges up to about 45 m3/s/m of basin width. This type does not
incorporate baffle piers because of the possibility of cavitation at high velocities.
The chute block spacing may be varied to avoid having fractional blocks, but the
width and spacing should be kept equal. The width and spacing of the sill dentates
indicated in Fig. 6.13 are recommended maximums and may be reduced in narrow
basins to increase their number; again the width and spacing should be kept equal.
The basin length, LII, is equal to the hydraulic jump length, which is related to the
entry Froude number, as shown in Fig. 6.14. This shows the significant reduction in
jump length affected by the stilling basin features. The basin floor level should be
set so that the tailwater is at least at the sequent depth of y1 for all discharges
anticipated. The incipient sweep-out level is about 5% less than this, so setting for
the sequent depth provides this margin of safety. The basin performance is not
affected appreciably by the slope of the entrance, but for slopes greater than 45o the
transition to the stilling basin floor should be curved (rather than sharp as shown in
Fig. 6.13a) with a radius of at least 4y1.
The water surface profile, which can also be considered as the pressure profile,
can be approximated by a straight line at angle a drawn from the tailwater level at
the end of the basin (Fig. 6.13). The value of a depends on Fr1 according to

a ¼ 5:06 ln Fr1  1:43 ð6:46Þ

The USBR Basin III incorporates baffle piers, resulting in a more compact and
therefore more economical design. Baffle piers introduce vulnerability to cavitation,
however, so the flow conditions are more restrictive than for Basin II. The design is
appropriate for entry Froude numbers greater than 4.5 but with an upper limit on
entry velocity of about 15–18 m/s and on discharge of about 18 m3/s/m of basin
width. The recommended geometry and dimensions are shown in Fig. 6.15.

4
L/y2*

3
Basin I, IV
2
Basin II
1 Basin III

0
2 4 6 8 10 12 14 16
Fr1

Fig. 6.14 USBR Basin lengths (adapted from Peterka 1978)


6.3 Standard Stilling Basins 221

y1
y2*
0.2 h3
0.5 y2 *
slope = 1:1
h1 = y1 h3 h4

0.8 y2*
slope = 2:1
LIII

(a) Longitudinal section

0.5 y1 0.375 h3
w1 = y1
s1 = y1 w3 = 0.75 h3
s3 = 0.75 h3

(b) Plan

Fig. 6.15 USBR Basin III geometry (adapted from Peterka 1978)

The height of the baffle piers is related to the entry Froude number according to

h3
¼ 0:59 þ 0:17 Fr1 ð6:47Þ
y1

The width and spacing may deviate from 0.75 times the height as recommended,
but the sum of the widths should remain equal to the sum of the spaces. The baffle
pier shape may be as shown, or they may be cubes, but the corners should not be
rounded as this would reduce the effectiveness of the flow separation producing
energy-dissipating eddies. The recommended distance of the baffle piers from the
downstream face of the chute blocks should be adhered to. The end sill should be
solid rather than dentated as for Basin II. Its height is not crucial, but the USBR
recommends a slight variation with Fr1 according to

h4
¼ 1:0 þ 0:056 Fr1 ð6:48Þ
y1
222 6 Energy Dissipation Structures

The basin length (LIII) is given by the lowest curve in Fig. 6.14, showing that the
appurtenances reduce the jump length to less than half of the length of a simple
jump.
As for Basin II, the tailwater level should correspond to the sequent of the inflow
depth, although the incipient sweep-out level is about 15–18% lower, giving a
greater margin of safety. The entrance transition should also be rounded for entry
slopes steeper than 45%.
The water surface profile (and pressure profile) shows a very steep rise from
about half the tailwater depth to the full depth in the vicinity of the baffle piers
(Fig. 6.15a).
The USBR Basin IV is designed for low-head dam and canal structures where
the entry Froude number is in the range 2.5–4.5, i.e. corresponding to the oscillating
jump where wave suppression is necessary. This basin is therefore an alternative to
the wave suppressors described in Sect. 6.2. The recommended geometry and
dimensions are shown in Fig. 6.16.
The design is intended to reduce appurtenances to a minimum. The usual chute
blocks are replaced by larger deflector blocks with sloping top surfaces. Their width
can be increased up to a maximum of y1, but the width to spacing ratio should be
kept at 1:2.5. The sill (which is optional) can be sized as for Basin III, i.e. with a
slope of 2:1 and a height according to Eq. (6.48). This geometry does not reduce
the hydraulic jump length, so the recommended basin length is as indicated by the
top curve in Fig. 6.13.
The floor level should be set for a tailwater depth equal to 1.1 times the sequent
depth of the incoming flow. Performance will be satisfactory for the full range of
discharges if this is designed for the maximum discharge.

2 y1 min
5o slope

2 y1 h4

LI
slope = 2:1
(a) Longitudinal section

s1 = 2.5 w1
w1 = 0.75 y1

(b) Plan

Fig. 6.16 USBR Basin IV geometry (adapted from Peterka 1978)


6.3 Standard Stilling Basins 223

0.2 y1
0.2 hs
y1 y1 hs = 0.2 y2 * slope = 2:1

L1
L = 3 y2*
(a) Longitudinal section

0.375 y1 min

0.7 y1
0.7 y1
0.15 y2*

(b) Plan

Fig. 6.17 Geometry of USBR Basin for low Fr1 (adapted from George 1978)

Another basin design developed by the Bureau of Reclamation (George, 1978)


for low Froude numbers (2.5 to about 6.0) includes chute blocks and baffle piers
(similar to Basin III), resulting in a much shorter basin. This was considered
superior to the St Anthony Falls basin (see below) which causes rough flow and
high scour potential beyond the basin at low Froude numbers. The geometry and
dimensions are shown in Fig. 6.17.
The sill is dentated, with equal dentate widths and spacings. The distance from
the chute blocks to the baffle piers (X) is given by

X
¼ 2:65 Fr0:69 ð6:49Þ
y2 1

and the distance from the chute blocks to the sill (L1) by

L1
¼ 3:5  0:30 Fr1 ð6:50Þ
y2

The overall length of the stilling basin is the length of the hydraulic jump, which
may be assumed to be equal to 3y*2 over the applicable range of Fr1. This is
significantly less than for Basin IV (intended for the same range of Fr1) and similar
to that for Basin III. For very low Fr1, the distance L1 plus the length of the sill may
exceed L, in which case the basin should be extended to include the sill.
224 6 Energy Dissipation Structures

Another common design for small structures is the St Anthony Falls


(SAF) stilling basin (Blaisdell, 1959). It has a particularly wide range of application
conditions (1.7 < Fr1 < 17) and is very effective in shortening the jump. It also
incorporates baffle blocks, a sloping entrance and an end sill. Because the size of the
basin depends on the entry flow depth, it is sometimes advantageous to flare the
transition from the discharging structure or chute to the basin in order to increase
the width and hence reduce the depth. The diverging side walls are then extended
through the stilling basin, resulting in a trapezoidal plan shape. The geometry and
dimensions for a straight-sided basin are shown in Fig. 6.18.
The length of the basin (LB) is related to the incoming Froude number by

4:5y2
LB ¼ ð6:51Þ
Fr0:76
1

The chute blocks may be placed with either a block or a space next to the side
wall, as long as the blocks are symmetrical about the outlet centre line. The baffle

top of side wall y2* /3

TW

y1 y1 0.07 y2*

LB /3
LB

(a) Longitudinal section

wingwall

0.375 y1 min 45o pref.

0.75 y1
0.75 y1

(b) Plan

Fig. 6.18 St. Anthony Falls stilling basin geometry


6.3 Standard Stilling Basins 225

piers are placed directly downstream of the openings between the chute blocks at a
distance equal to LB/3 from the ends of the chute blocks. They should occupy
between 40% and 50% of the basin width, and no pier should be closer to the side
wall than 0.375 y1. For diverging stilling basins, the widths and spacings of the
baffle piers should be increased in proportion to the increase in the basin width at
their location. The end sill should be extended as a cutoff wall to prevent under-
mining of the basin by erosion. The floor of the basin should be set for a tailwater
(TW) height of

TW
1:4 Fr0:90
1 ð6:52Þ
y1

The side wall height should allow a freeboard of y*2/3 above the tailwater level to
accommodate the rough surface, which may extend beyond the basin at high
Froude numbers. Wingwalls should be provided for the transition from the basin to
the downstream channel. They should slope downwards at 45° from the end of the
basin, and preferably be flared at 45°.

6.4 Other Energy Dissipators

6.4.1 Bucket-Type Dissipators

If the tailwater at the bottom of a spillway is too deep to allow a free hydraulic jump
to form, energy can be dissipated effectively in a roller bucket (Fig. 6.19).

Fig. 6.19 A slotted bucket dissipator


226 6 Energy Dissipation Structures

Fig. 6.20 Bucket-type dissipator

The dissipator works on the principle of induced turbulence and mixing, which
is initiated by a sharp upward deflection of the flow (Fig. 6.20). This causes a roller
motion within the bucket, which induces a reverse roller immediately downstream.
The motion of the rollers and mixing with incoming water dissipate energy and
prevent scour.
Two basic types of submerged bucket dissipator have been developed
(Fig. 6.21), with many variations of the basic designs. The solid bucket is simply a
continuation of the spillway profile, with a circular radius. The slotted bucket is
dentated on an extended apron.
The solid and slotted configurations operate on the same principle, but the
hydraulic action of each has characteristics that impose certain limitations on its
application. The two types are therefore suited to different conditions. For the solid
bucket, the high-velocity flow is all directed upwards, creating a high boil and a
violent ground roller. The ground roller continuously draws bed material back
towards the lip of the bucket and keeps some material in constant agitation. In the
slotted bucket, the high-velocity jet leaves the lip at a much flatter angle, with only
part of it at an angle sufficient to cause a surface boil. The boil is therefore less
violent and there is better flow dispersion above the ground roller. There is less high
energy flow within the bucket and the downstream flow is smoother.
These characteristics have relative advantages and disadvantages in practice.
The solid bucket may be undesirable because of abrasion of the concrete surface by

(a) Solid bucket (b) Slotted bucket

Fig. 6.21 Basic submerged bucket designs


6.4 Other Energy Dissipators 227

the agitated bed material. Also, the higher surface turbulence generated by the
violent boil can cause damage to river banks downstream. The slotted bucket
provides more effective energy dissipation with less severe disturbance of the water
surface and the stream bed. It is, however, susceptible to sweep out at low tail-
waters, which results in high-velocity flow and potential scour problems down-
stream (Fig. 6.22a). Effective operation therefore requires the tailwater to be above
a minimum depth (Fig. 6.22b). If the tailwater is too deep for a slotted bucket, then
diving flow can occur (Fig. 6.22c). Here, the jet leaving the bucket is not deflected
to the surface but is depressed to the river bed, causing scour immediately down-
stream of the bucket and deposition of the scoured material a short distance
downstream. There is also, therefore, a maximum tailwater requirement.
By virtue of its better energy dissipation characteristics, the slotted bucket is
therefore a preferable design, provided its more stringent tailwater requirements can
be met. Full details of the design geometries are given by Peterka (1978) and USBR
(1973), including the limitations on tailwater depths for the slotted bucket. As for
the hydraulic jump stilling basins, model studies are advisable if the specified limits
of application are violated.

yt minimum

(a) Sweep-out condition

yt maximum
yt minimum

(b) Normal operating condition

yt maximum

(c) Diving flow condition

Fig. 6.22 Flow conditions for slotted buckets


228 6 Energy Dissipation Structures

6.4.2 Impact-Type Dissipators

Impact-type stilling basins have been developed for pipe or relatively narrow open
channel outlets (Peterka, 1978) (Fig. 6.23). Energy dissipation is effected by the
impact of the incoming jet on a vertical baffle spanning across the basin (Fig. 6.24)
and turbulence in eddies formed after the impact.
This basin’s capacity is generally limited by the feasibility of the structural
design, which must allow for impact loads and vibration, and to an incoming
velocity of about 15 m/s. Designs have been used for discharges up to 11 m3/s. It
has been found to be significantly more effective in dissipating energy than a
hydraulic jump on a horizontal floor. The performance is not very dependent on
tailwater conditions, although is best when the tailwater is less than about halfway
up the baffle.
A standard design geometry with dimensions for a range of discharges is pre-
sented by Peterka (1978).

Fig. 6.23 Impact-type stilling basin at an impoundment outlet

Fig. 6.24 Impact-type stilling basin action


6.4 Other Energy Dissipators 229

6.4.3 Baffled Spillways

For low weirs, chutes and drop structures, energy can be effectively dissipated by a
system of baffles over the whole slope length (e.g. Peterka, 1978; Rhone, 1977), as
illustrated in Figs. 6.25 and 6.26.

Fig. 6.25 Baffles on a low-head spillway

Fig. 6.26 Baffled spillway


layout

(a) Section

(b) Plan
230 6 Energy Dissipation Structures

The baffles limit the acceleration of flow down the slope and ensure a reasonable
terminal velocity for any slope length. Peterka (1978) has reviewed the results of
model and prototype experiments on baffled spillways and drop structures and has
formulated a procedure for their design.
The design is effective for unit width discharges up to about 5.6 m3/s/m and for
slopes between 1 V:2H and 1 V:4H. The entrance to the slope should be set above
the bottom of the approach channel to create a pool and ensure that the approach
velocity does not exceed 1.5 m/s less than the critical velocity. Recommended
baffle dimensions are shown in Fig. 6.27, in which the baffle height (H) is equal to
0.8yc. Apart from the baffle widths and spacings being equal and not less than H,
the dimensions are not critical. In narrow structures, partial baffles with widths of
0.5H–1.0H should be placed adjacent to the training walls wherever necessary to
complete rows. The bottom of the slope should be constructed below the stream bed
and backfilled, with riprap protection placed at the ends of the training walls.

Fig. 6.27 Baffle dimensions

(a) Section

1.5H 1.5H
(b) Plan
6.4 Other Energy Dissipators 231

6.4.4 Stepped Chutes and Spillways

Energy can be effectively dissipated over the length of a spillway if it constructed as


a series of vertical steps (Chanson 1993, Christodoulou 1993, Essery and Horner
1978, Rajaratnam 1990). This arrangement is particularly attractive for
roller-compacted concrete dams. The hydraulics of stepped chutes and spillways is
dealt with in Chap. 4.

6.4.5 Spillway Splitters

A flow splitter system placed on the face of a dam spillway a short distance below
the crest provides an effective way of dissipating energy (Fig. 6.28). The design,
first conceived by Roberts (1943), comprises a row of evenly spaced discrete
features (splitters) along the spillway face located above a solid, horizontal shelf,
which may be cantilevered from or inset into the spillway face (Fig. 6.29). The
splitters separate the flow into two streams. The upper stream is deflected away
from the spillway face along the upper surface of the splitters. The lower stream
passes between the splitters and is deflected by the shelf. The trajectories of the two
dispersed and aerated streams intersect, further increasing the aeration and resulting
in a spray with greatly reduced impact on the downstream apron or riverbed.
Roberts (1980) used laboratory test results, including those of Roberts (1943), to
derive geometric and performance relationships for designing a splitter arrangement

Fig. 6.28 Splitter arrangement on a gated spillway, showing operation at opened gate
232 6 Energy Dissipation Structures

H
x x
P L W S

hdam y
Ls
θ (b) Splitter details
yt
Rmin
Rmax

(a) General profile

Fig. 6.29 Spillway splitter dimensions

for spillway heads up to 3 m. These, as also reported by Calitz and Basson (2015),
are presented below.
The splitter arrangement is suitable for relatively high dams with

x þ y [ 4HD ð6:53Þ

where HD is the design head and x and y are defined in Fig. 6.29. Performance is
satisfactory for heads between a minimum (Ha) necessary for effective dispersion of
the flow streams and a maximum (Hc) above which the splitters become drowned,
ineffective and vulnerable to cavitation. These limiting heads are related to the
design head by

Hc ¼ 1:2 HD ð6:54Þ

and

Hc
Ha ¼ ð6:55Þ
4:4

The splitter dimensions are related to the number (N) installed over the length
(B) considered, which may be the total spillway length or the distance between piers
if these are installed. The splitter width (W) is then given by

3B
W¼ ð6:56Þ
7N
6.4 Other Energy Dissipators 233

and the splitter spacing (S), length (L) and surface height above the shelf level
(T) by

4B 4
S¼L¼T¼ ¼ W ð6:57Þ
7N 3

The shelf width (Ls) should be in the range

Ls
1:25   1:50 ð6:58Þ
L

The performance of the splitters depends on the level of the splitter surface
below the spillway crest (P). The splitters should be far enough below the crest for
sufficient kinetic energy to have been developed to ensure effective dispersion, but
nearer the crest than the distance to the point of aeration inception. The maximum
value of P can therefore be determined using the methods for locating the inception
point presented in Chap. 4, such as Wood et al.’s (1983) Eq. (4.51). The optimal
splitter width for effective dispersion is related to P. Roberts (1943) presented
experimental data relating the optimal W/P to Hc/P, which can be represented (with
R2 = 0.995) as
 1:24
W Hc
¼ 0:324 ð6:59Þ
P P

The design can proceed either by choosing the splitter geometry and determining
a corresponding value for P, or by choosing P and determining an appropriate
geometry. The geometry can be chosen by specifying a value for either W or N,
from which the other dimensions follow through Eqs. (6.56) to (6.58). P is then
determined from Eq. (6.59), rearranged as
 4:17
0:324 Hc1:24
P¼ ð6:60Þ
W

and checked to be less than would allow the point of aeration inception to be
reached before the splitter location.
If P is chosen first, subject to satisfying the aeration inception point limit, W is
then obtained from Eq. (6.60), and the other dimensions from Eqs. (6.56) to (6.58).
Lastly, the extent of the spray jet (Rmin to Rmax) must be determined to ensure
that it falls within the planned plunge pool or to dimension an apron if one is to be
constructed. These distances are given by Eqs. (6.61) and (6.62).
pffiffiffiffiffi
Rmax ¼ 2 f xy  y cot h ð6:61Þ
pffiffiffiffiffi
Rmin ¼ f xy  y cot h ð6:62Þ
234 6 Energy Dissipation Structures

The distances x and y are shown in Fig. 6.29 (for practical purposes
x  HD + P). Roberts (1943) presented experimental data relating f to (P + T) and
HD, which can be described (with R2 = 0.991) by

f ¼ 1:025 e0:022ððP þ T Þ=HD Þ ð6:63Þ

The design is not unique for a specified design head, as different combinations of
splitter size and P can perform satisfactorily. The combination selected must ensure
that the point of aeration inception does not occur before the splitters and that the
aerated jet clears the toe of the dam. The smaller the splitter size chosen, the greater
will be P and the smaller will be Rmin.
The relationships presented above are applicable for unaerated splitters, origi-
nally considered to be effective for design heads less than 3.0 m (corresponding to a
unit width discharge (q) of about 12 m3/s/m for a conventional ogee spillway).
Through specific model testing, designs including aeration of the splitters have been
used for considerably greater heads (Jordaan 1989). Calitz and Basson (2018)
conducted model tests for splitters designed as above for HD = 6.7 m
(q = *40 m3/s/m) and found satisfactory performance up to the design head,
above which the splitters became drowned and cavitation became inevitable. By
providing artificial aeration satisfactory performance, without cavitation risk, was
extended to at least HD = 7.6 m (q = * 50 m3/s/m).
Example 6.6
Dimension an effective spillway splitter arrangement for a 30.0 m high dam with a
150 m long ogee spillway sloped at 54°. The design head is 2.40 m and the tailwater
depth at the corresponding discharge is 4.0 m. Determine also the extent of the spray jet
at the base of the dam.
Solution Refer to Fig. 6.29 for notation.
Check the suitability of the splitter option for energy dissipation, according to
Eq. (6.53), i.e.

x þ y [ 4 HD
x þ y ¼ hdam þ HD  yt
¼ 30:0 þ 2:4  4:0 ¼ 28:8 m
4 HD ¼ 4  2:4 ¼ 9:6 m

x + y = 28.8 m > 4HD = 9.6 m. Therefore, the suitability


condition is satisfied.
Establish head limits for satisfactory performance:
Upper limit from Eq. (6.54), i.e.

Hc ¼ 1:2 HD ¼ 1:2  2:40 ¼ 2:88 m

Lower limit from Eq. (6.55), i.e.


6.4 Other Energy Dissipators 235

Hc 2:88
Ha ¼ ¼ ¼ 0:66 m
4:4 4:4
Determine Pmax to ensure splitters are located beyond the air entrainment inception
point:

Pmax  L sin h

The distance to the inception point (L) is insensitive to H, so calculate Pmax for HD.
Calculate L using Wood et al.’s (1983) Eq. (4.51), i.e.
 0:11  0:10
d x x
¼ 0:0212 ðAÞ
x Hs ks

by iteration, with x = L and d = y, the flow depth at the inception point.


Calculations with the successful iteration values are shown below:
Try L = 45.3 m.
With L = 45.3 m, try y = 0.296 m
Then
Hs ¼ H þ L sin h  y cos h

with
h ¼ 54:0o

and

H ¼ HD ¼ 2:4 m

So

Hs ¼ 2:40 þ 45:3 sin 54:0  0:296 cos 54:0 ¼ 38:9 m

and
pffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
V¼ 2gHs ¼ 2  9:8  38:9 ¼ 27:6 m/s

Check:
q

V
3=
q ¼ CD HD 2
3
¼ 2:2  2:4 =2
for a conventional ogee spillway
¼ 8:18 m3 =s/m
So
236 6 Energy Dissipation Structures

8:18
y¼ ¼ 0:296 m
27:6

which is equal to the trial value. Therefore y = 0.296 m.


Now calculate RHS of (A) with the trial values. (The result is insensitive to ks;
a value of 0.30 mm is representative.)

0:296
RHS ¼ 45:30:11   ¼ 45:2 m
45:3 0:10
0:0212 38:9 0:00030

which is approximately equal to the trial value of L.


Therefore L = 45.3 m.
Therefore

Pmax  45:3 sin 54 ¼ 36:6 m

which is greater than the dam height, so there is no effective restriction


on the location of the splitters.
Select P and determine splitter dimensions:
With no restriction on Pmax, the choice of P is arbitrary. A value of P = 7.5 m is used
for example. The splitter dimensions then follow, according to Eqs. (6.59), (6.57)
and (6.58), as follows.
W is defined by Eq. (6.59), i.e.
 1:24
W Hc
¼ 0:324
P P

Therefore
 1:24
Hc
W ¼ P 0:324
P
 
2:88 1:24
¼ 7:5  0:324 ¼ 0:74 m
7:5

Then, from Eq. (6.57)

4 4
S ¼ L ¼ T ¼ W ¼ 0:74 ¼ 0:99 m
3 3
And, from Eq. (6.58), Ls must be in the range
Ls
1:25   1:50
L

i.e.

Ls min ¼ 1:25 L ¼ 1:25  0:99 ¼ 1:24 m


6.4 Other Energy Dissipators 237

and

Ls max ¼ 1:50 L ¼ 1:50  0:99 ¼ 1:49 m

Choose

Ls ¼ 1:36 m

Determine the extent of the spray jet:


Rmin will occur at the low end of the operating range, i.e. for H = Ha = 0.66 m.
Then, from Eq. (6.62),
pffiffiffiffiffi
Rmin ¼ f xy  y cot h

with, from Eq. (6.63),

f ¼ 1:025 e0:022ððP þ T Þ=Ha Þ


¼ 1:025 e0:022ðð7:5 þ 0:99Þ=0:66Þ ¼ 0:772

and

x ¼ Ha þ P ¼ 0:66 þ 7:5 ¼ 8:16 m


y ¼ hdam  P  yt ¼ 30:0  7:5  4:0 ¼ 18:5 m

Therefore
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Rmin ¼ 0:772 8:16  18:5  18:5 cot 54 ¼ 3:96 m

So the spray jet will fall on the spillway face.


Rmax will occur at the high end of the operating range, i.e. for H = Hc = 2.88 m.
Then, from Eq. (6.61),
pffiffiffiffiffi
Rmax ¼ 2f xy  y cot h

with, from Eq. (6.64),

f ¼ 1:025 e0:022ððP þ T Þ=Hc Þ


¼ 1:025 e0:022ðð7:5 þ 0:99Þ=2:88Þ ¼ 0:961

and

x ¼ Hc þ P ¼ 2:88 þ 7:5 ¼ 10:38 m


y ¼ hdam  P  yt ¼ 18:5 m as before
238 6 Energy Dissipation Structures

Therefore
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Rmax ¼ 2  0:961 10:38  18:5  18:5 cot 54 ¼ 13:2 m

which indicates the length of apron required for protection of the river bed.

Problems

6:1 A long concrete (n = 0.013) chute has a slope of 0.10 and is designed to
convey a discharge of 25 m3/s. The chute terminates in a stilling basin with a
horizontal bed followed by an abrupt upward step to a long channel with a
slope of 0.0010 and n = 0.020. The width is 5.0 m throughout.
a. Determine an appropriate step height and basin length to keep the hy-
draulic jump within the basin and on the horizontal bed.
b. Determine the positions of the hydraulic jump if the step height is
(i) 1.20 m and (ii) 2.50 m.

For each case sketch and dimension the water surface profile and the total
energy line.
6:2 Water is to be released from a reservoir at a discharge of 20.0 m3/s and a
velocity of 12.0 m/s into a 6.0 m wide, rectangular, concrete-lined (n = 0.013)
canal on a slope of 0.00050. Energy dissipation is required before the water
enters the canal.
a. Size a stilling basin with only a smooth upward step, checking for
sweep-out.
b. Size a stilling basin with only an abrupt upward step, to create a Type A
hydraulic jump.
c. Size a stilling basin with only an abrupt upward step, to create a Type B
hydraulic jump.
d. Size a dissipator using just a sill, with the smallest satisfactory sill height.
e. Size a row of dissipating baffle blocks to be placed 4.0 m from the
reservoir outlet. Assume standard USBR block shapes with a blockage
ratio of 0.50.

6:3 A reservoir releases water onto an 8.0 m wide, horizontal apron at a discharge
of 30.0 m3/s and a velocity of 12.0 m/s. A 1.40 m high sill is placed 10.0 m
from the release point. A further 10.0 m downstream from the sill, the apron
leads to an 8.0 m wide rectangular, concrete-lined channel on a slope of
0.0010 through an abrupt step. Determine the height of the step to ensure that
the hydraulic jump forms with its toe close to the release point. (Assume the
effects of the sill and the step to be independent, and friction between the sill
and the step to be negligible.)
References 239

References

Bakhmeteff, B. A., & Matzke, A. E. (1936). The hydraulic jump in terms of dynamic similarity.
Transactions of the American Society of Civil Engineers, 101, 630–647.
Basco, D. R., & Adams, J. R. (1971). Drag forces on baffle blocks in hydraulic jumps. Journal of
the Hydraulics Division, ASCE, 97(HY12), 2023–2035.
Blaisdell, F. W. (1959). The SAF stilling basin, agriculture handbook No. 156, U S Department of
Agriculture.
Bradley, J. N., & Peterka, A. J. (1957). The hydraulic design of stilling basins. Journal of the
Hydraulics Division, ASCE, 83(HY5), 1401–1406.
Bretz, N. V. (1987). Ressaut hydraulique force par seuil, Thèse No. 699, Ecole Polytechnique
Fédérale de Lausanne, Switzerland.
Calitz, G., & Basson, G. R. (2015). The design of Roberts splitters for energy dissipation at dam
spillways. Design and Construction of Hydraulic Structures: Stellenbosch University.
Calitz, G., & Basson, G. R. (2018). The effect of aeration through an internal gallery of a dam on
the cavitation risk of Roberts splitters. Journal of the South African Institution of Civil
Engineering, 60(1), 31–43.
Carollo, F. G., Ferro, V., & Pampalone, V. (2007). Hydraulic jumps on rough beds. Journal of
Hydraulic Engineering, 133(9), 989–999.
Carollo, F. G., Ferro, V., & Pampalone, V. (2009). New solution of classical hydraulic
jump. Journal of Hydraulic Engineering, 135(6), 527–531.
Carollo, F. G., Ferro, V., & Pampalone, V. (2011). Sequent depth ratio of a B-jump. Journal of
Hydraulic Engineering, 137(6), 651–658.
Carollo, F. G., Ferro, V., & Pampalone, V. (2012). New expression of the hydraulic jump roller
length. Journal of Hydraulic Engineering, 138(11), 995–999.
Castro-Orgaz, O., & Hager, W. H. (2009). Classical hydraulic jump: basic flow features. Journal
of Hydraulic Research, 47(6), 744–754.
Chanson, H. (1993). Stepped spillway flows and air entrainment. Canadian Journal of Civil
Engineering, 20(3), 422–435.
Chow, V. T. (1959). Open-channel hydraulics, McGraw-Hill.
Christodoulou, G. C. (1993). Energy dissipation on stepped spillways. Journal of Hydraulic
Engineering, 119(5), 644–650.
Essery, I. T. S., & Horner, M. W. (1978). The hydraulic design of stepped spillways, 2nd Edition,
CIRIA Publication R33.
Fiuzat, A. A. (1986). Flow characteristics of the hydraulic jump in a stilling basin with an abrupt
bottom rise. Journal of Hydraulic Research, 24(3), 207–209.
Forster, J. W., & Skrinde, R. A. (1950). Control of hydraulic jump by sills. Transactions of the
American Society of Civil Engineers, 115, 973–987.
Fowler, N. (2019). Hydraulic jumps and steps: the relationship between jump location and
sequent depth ratio using a step, MSc(Eng) project report, University of the Witwatersrand,
South Africa.
French, R. H. (1985). Open-channel hydraulics, McGraw-Hill.
George, R. L. (1978). Low Froude number stilling basin design, Report REC-ERC-78–8,
Engineering and Research Center, United States Department of the Interior, Bureau of
Reclamation, Denver, Colorado, USA.
Hager, W. H. (1985). B-jumps at abrupt channel drops. Journal of Hydraulic Engineering, 111(5),
861–866.
Hager, W. H., & Bremen, R. (1989). Classical hydraulic jump; sequent depths. Journal of
Hydraulic Research, 27(5), 565–585.
Hager, W. H., Bremen, R., & Kawagoshi, N. (1990). Classical hydraulic jump; length of roller.
Journal of Hydraulic Research, 28(5), 591–608.
240 6 Energy Dissipation Structures

Hager, W. H., & Bretz, N. V. (1986). Hydraulic jumps at positive and negative steps. Journal of
Hydraulic Research, 24(4), 237–253.
Hager, W. H., & Li, D. (1992). Sill-controlled energy dissipator. Journal of Hydraulic Research,
30(2), 165–181.
Hager, W. H., & Sinniger, R. (1985). Flow characteristics of the hydraulic jump in a stilling basin
with an abrupt bottom rise. Journal of Hydraulic Research, 23(2), 101–113.
Henderson, F. M. (1966). Open channel flow, Macmillan.
Hsu, E. Y. (1950). Discussion of “Control of the hydraulic jump by sills” by J W Forster and R A
Skrinde. Transactions of the American Society of Civil Engineers, 115, 988–991.
Hughes, W. C., & Flack, J. E. (1984). Hydraulic jump properties over a rough bed. Journal of
Hydraulic Engineering, 110(12), 1755–1771.
Jordaan, J. M. (1989). The Roberts splitter: Fifty years on. The Civil Engineer in South Africa, 31
(10), 319–321.
Karki, K. S., & Mishra, S. K. (1986). Flow characteristics of the hydraulic jump in a stilling basin
with an abrupt bottom rise. Journal of Hydraulic Research, 24(3), 210–2015.
Kindsvater, C. E. (1944). The hydraulic jump in sloping channels. Transactions of the American
Society of Civil Engineers, 109, 1107–1154.
Mehrotra, S. C. (1976). Length of hydraulic jump. Journal of the Hydraulics Division, ASCE, 102
(HY7), 1027–1033.
Ohtsu, I. (1981). Forced hydraulic jump by a vertical sill. Transactions of the JSCE, Journal of the
Hydraulic and Sanitary Engineering Division, 13, 165–168.
Ohtsu, I., & Yasuda, Y. (1987). Discussion of “Hydraulic jumps at positive and negative steps”.
Journal of Hydraulic Research, 25(3), 407–413.
Ohtsu, I., & Yasuda, Y. (1991). Hydraulic jump in sloping channels. Journal of Hydraulic
Engineering, 117(7), 905–921.
Ohtsu, I., Yasuda, Y., & Hashiba, H. (1996). Incipient jump conditions for flow over a vertical sill.
Journal of Hydraulic Engineering, 122(8), 465–469.
Peterka, A. J. (1978). Hydraulic design of stilling basins and energy dissipators, Engineering
Monograph No. 25, revised, United States Department of the Interior, Bureau of Reclamation.
Rajaratnam, N. (1967). Hydraulic jumps, Advances in Hydroscience (Vol. 4, pp. 197–280). New
York: Academic Press.
Rajaratnam, N. (1990). Skimming flows in stepped spillways. Journal of Hydraulic Engineering,
116(4), 587–591.
Rajaratnam, N., & Murahari, V. (1971). A contribution to forced hydraulic jumps. Journal of
Hydraulic Research, 9(2), 217–240.
Rajaratnam, N., & Ortiz, N. V. (1977). Hydraulic jumps and waves at abrupt drops. Journal of the
Hydraulics Division, ASCE, 103(HY4), 381–394.
Ranga Raju, K. G., Kitaal, M. K., Verma, M. S., & Ganeshan, V. R. (1980). Analysis of flow over
baffle blocks and end sills. Journal of Hydraulic Research, 18(3), 227–241.
Rhone, T. J. (1977). Baffled apron as spillway energy dissipator. Journal of the Hydraulics
Division, ASCE, 103(HY12), 1391–1401.
Roberts, C. P. R. (1980). Hydraulic Design of Dams, Report, Division of Special Tasks,
Department of Water Affairs, Forestry and Environmental Conservation, Pretoria.
Roberts, D. F. (1943). The dissipation of energy of a flood passing over a high dam, Transactions,
South African Institution of Civil Engineers, XLI:48–92.
Schulz, H. E., Nobrega, J. D., Simoes, A. L. A., Schulz, H., & Porto, R. D. M. (2015). Details of
hydraulic jumps for design criteria of hydraulic structures, in Hydrodynamics—Concepts and
Experiments, Schulz, H. E. (Ed.), Chapter 4: 73–116, ISBN: 978-953-51-2034-6, https://doi.
org/10.5772/58963.
United States Bureau of Reclamation. (1987). Design of Small Dams (3rd ed.). Bureau of
Reclamation, United States Government Printing Office, Washington, D.C.: United States
Department of the Interior.
References 241

Valiani, A. (1997). Linear and angular momentum conservation in hydraulic jump. Journal of
Hydraulic Research, 35(3), 323–354.
Wood, I. R., Ackers, P., & Loveless, J. (1983). General method for critical point on spillways.
Journal of Hydraulic Engineering, 109(2), 308–312.

Further Reading

Chanson, H. (Ed.) (2015). Energy Dissipation in Hydraulic Structures, CRC Press.


Chapter 7
Flow-Measuring Structures

7.1 Introduction

The discharge in a river or canal can be determined either directly by measuring and
integrating the velocity distribution or indirectly by relating measured water levels
to discharge through a known rating function. The former approach is particularly
useful where only one-off measurements are needed, such as for compiling a stage–
discharge relationship, or where velocities are also required, such as for ecohy-
draulic applications. Velocities are commonly measured using propeller-type cur-
rent meters but modern instruments for acoustic Doppler current profiling have
made the approach considerably more efficient. For continuous monitoring, such as
for hydrological purposes or water supply allocations, the indirect approach is
reliable and practical and widely used. A stable rating function can be obtained by
installing a permanent structure in the channel. Measuring structures are designed,
as far as possible, to operate as true controls where the discharge can be related to
the upstream head only—often referred to as the ‘modular’ flow condition. They
can, however, still function under submerged flow conditions, but this requires
measurement of the downstream as well as the upstream flow depth. Some struc-
tures which are installed for purposes other than flow measurement can also be used
for measurement, including sluice gates, culverts, dam spillways, bridges and
diversion structures.
Many different kinds of measuring structures have been used, but most are
variations of two basic types: weirs, which induce control through vertical con-
traction of the flow, and flumes, which induce control through horizontal con-
traction of the flow. This chapter will identify the more common types, provide
their rating functions and enable the influence of measurement error on their
accuracy to be made.

© Springer Nature Switzerland AG 2020 243


C S James, Hydraulic Structures,
https://doi.org/10.1007/978-3-030-34086-5_7
244 7 Flow-Measuring Structures

7.2 Weirs

A weir is a structure which obstructs the flow, causing the water level to rise and
resulting in flow over a well defined crest. The water level in the pool upstream of
the crest is functionally related to the discharge and can, therefore, be used as a
measure of the discharge.
Many different types of weir are available for use under different conditions.
They are classified according to the shape of the crest or the opening through which
the water must pass. The broadest classification distinguishes between
broad-crested and sharp-crested (also known as thin-plate) weirs.

7.2.1 Sharp-Crested Weirs

The crest of a sharp-crested weir is formed from steel, with a 90° sharp front edge, a
flat horizontal top surface 1–2 mm wide, and a 45° chamfer on the downstream
side. A variety of crest alignment shapes have been used, the most common
being rectangular and triangular.
Rectangular Weirs
A rectangular weir has a horizontal crest and vertical sides so that a rectangular
opening is formed. There are two types of rectangular weir:
1. A contracted weir has a crest length which is less than the width of the channel,
and the vertical sides form part of the sharp crest. The flow must, therefore,
contract horizontally as well as vertically through the opening (Fig. 7.1).
2. A suppressed weir has a crest that extends across the full width of the channel,
with the channel walls forming the vertical sides (Fig. 7.2). In this case there is
no horizontal contraction.

(a) Side view (b) Front view

Fig. 7.1 Contracted weir


7.2 Weirs 245

(a) Side view (b) Front view

Fig. 7.2 Suppressed weir

The relationship between water level and discharge cannot be determined


completely by theoretical analysis. However, by making certain simplifying
assumptions, the general form of the relationship can be derived. This can be
calibrated or refined for particular geometries by empirical correction.
The analysis can be done most simply for the suppressed rectangular weir, where
the flow can be considered to be two-dimensional (Fig. 7.3). It is required to relate
the discharge over the weir to the upstream energy (H) or preferably to the upstream
flow depth (h), which is directly measurable. (Note that critical flow cannot be
assumed over the crest because the non-hydrostatic pressure distribution violates
the assumptions under which critical flow results were derived.)
The discharge over the weir shown in Fig. 7.3 can be determined by integrating
the velocity over the flow depth above the crest (this could be done at any other
location, but the velocity distribution is best known at the crest). This requires
knowledge of the thickness of the nappe above the crest and the distribution of the

total energy
Vo2/2g
H z
h p

Fig. 7.3 Two-dimensional flow over a sharp-crested weir


246 7 Flow-Measuring Structures

velocity through the nappe. These are not easy to determine, and for the purpose of
this general analysis, the following simplifying assumptions are made, with the
understanding that empirical corrections will have to be made to account for them
later.
• There is no vertical contraction over the crest.
• The pressure is atmospheric right across the nappe.
• There is no energy loss in the approach flow to the crest.
Although unrealistic, the first assumption defines the nappe thickness and hence
the limits for integration along the vertical direction of the velocity distribution, i.e.
z from V2o/2 g to H, where Vo is the approach velocity, g is gravitational acceleration
and H is the total energy relative to the weir crest. (The kinetic energy correction
factor is neglected for simplicity, and will also be accounted for in the empirical
correction to follow.)
The second assumption is also unrealistic: the pressure is indeed atmospheric at
the upper and lower surfaces, but it is greater than atmospheric in between, the
actual distribution being difficult to determine. But this assumption implies that the
distance from any point within the nappe to the energy line is the velocity head
only, and this enables the velocity distribution to be determined as
pffiffiffiffiffiffiffi
v¼ 2gz

in which v is the local velocity and z is the distance measured downwards from the
energy line. (Note that this application of the Bernoulli equation across streamlines
is valid because the flow across the nappe can be assumed to be irrotational.)
The third assumption is reasonable because the crest causes flow contraction
only, and does not introduce any energy dissipating mechanisms. This enables the
position of the energy line to be defined above the crest, which is necessary to
quantify z.
Making these assumptions, the unit width discharge over the weir crest is given
by
Z H
q¼ Vo2
vdz
2g
Z H pffiffiffiffiffiffiffi
¼ Vo2
2gzdz
2g
H
2 pffiffiffiffiffi 3=2
¼ 2gz
3 Vo2
2g

and so
7.2 Weirs 247

 2 3=2 !
2 pffiffiffiffiffi Vo
q¼ 2g H 3=2  ð7:1Þ
3 2g

It is generally more convenient in practice to relate discharge to upstream flow


depth, rather than the upstream energy level because it is the water level that is
measured. Therefore, substituting

Vo2
H ¼ hþ
2g

into Eq. (7.1) gives


 3=2  2 32 !
2 pffiffiffiffiffi Vo2 V
q¼ 2g hþ  o
3 2g 2g
 3=2  2 32 !
2 pffiffiffiffiffi 3=2 Vo2 Vo
¼ 2gh 1þ 
3 2gh 2gh

A contraction coefficient, Cc, is introduced to account for the assumption of no


contraction over the crest, i.e.
 3=2  2 32 !
2 pffiffiffiffiffi Vo2 Vo
q ¼ Cc 2gh3=2 1þ 
3 2gh 2gh

The contraction coefficient and the term in brackets both depend on the approach
flow conditions, and the two terms are generally combined and denoted the dis-
charge coefficient, Cd. Then

2 pffiffiffiffiffi
q ¼ Cd 2gh3=2 ð7:2Þ
3

Discharge coefficients are often grouped with constant terms in the equation, and
variants of Eq. (7.2) may not explicitly include the 2/3 term or both this and the
(2 g)0.5 term. These will then be incorporated in Cd, and care should always be
taken when using published empirical coefficient values to ensure that they are
appropriate for the equation forms actually used.
The discharge coefficient accounts for the assumptions in the derivation and
must be determined empirically for particular weir geometries. It is clear from the
above derivation that Cd in Eq. (7.2) depends on the approach velocity and the
upstream water level. It is, therefore, not a constant for a given weir, and depends
on both the water level and the approach geometry. Various formulations have been
proposed for Cd in terms of the geometric characteristics of the weir and the
approach flow depth (see, for example, Herschy (1995), Ackers et al. (1980) and
248 7 Flow-Measuring Structures

Wessels and Rooseboom (2009b)). Henderson (1966) presents a relationship


derived from experimental work by Rehbock in 1929 for Cd in terms of h and
W (the height of the crest above the bed), i.e.

h h
Cd ¼ 0:611 þ 0:08 for \5 ð7:3Þ
W W

Notice that for large W, Cd tends to 0.611. A large W implies a small approach
velocity, in which case Cc = Cd (according to the substitution made above). This
value of Cc (0.611) is the same as the theoretical Cc for a two-dimensional jet from
a large tank.
Equation (7.3) does not give good results for large values of h/W. As W tends to
zero, the condition approaches that for a free overfall, where critical flow occurs a
short distance upstream from the brink. For h/W > 20, Cd can be approximated by
Henderson (1966)
 3=2
W h
Cd ¼ 1:06 1 þ for [ 20 ð7:4Þ
h W

For 5 < h/W < 10 Eq. (7.3) can be used, but accuracy decreases (Cd = 1.135 for
h/W = 10). No results are available for 10 < h/W < 20.
The following factors also affect Cd.
– The thickness of the crest edge.
– Lateral contraction, if present.
– The pressure below the nappe. It was assumed in the derivation of Eq. (7.2) that
the pressure was atmospheric right across the nappe. For suppressed weirs the
edges are sealed and, unless ventilation is ensured, negative pressures will
develop and increase Cd.
– The distribution of the approach velocity. For a contracted weir with a relatively
short length, the velocity of approach of the water going directly over the crest
will be greater than the average over the channel width.
– The roughness of the upstream face of the weir. A rougher surface will reduce
the contraction effect and increase Cd.
– Adhesion and surface tension at very low heads.
For suppressed weirs the total discharge can be calculated by multiplying the
unit width discharge by the crest length, L, i.e.

2 pffiffiffiffiffi
Q ¼ qL ¼ Cd L 2gh3=2 ð7:5Þ
3

For contracted weirs there is a lateral contraction and the effective length of the
weir is less than the total length and, therefore,
7.2 Weirs 249

Q ¼ q Leff

in which Leff is the effective length. The Francis formula (Henderson 1966)
approximates the contraction as h/10 on each side, if L > 3 h, and therefore

2 pffiffiffiffiffi
Q ¼ Cd ðL  0:2hÞ 2gh3=2 ð7:6Þ
3

The effect of the end contractions can also be accounted for by adjusting the
discharge coefficient. Webber (1971) cites the Hamilton Smith formula in the
British Standard for the discharge coefficient for a weir with end contractions, i.e.
 
h
Cd ¼ 0:616 1  0:1 ð7:7Þ
L

The analysis and equations presented above apply only to modular flow, i.e.
where the weir acts as a true control and the discharge is uniquely related to the
upstream water level. This requires that the downstream water level is too low to
affect the flow over the crest. It is, therefore, advisable to ensure this free flow
condition when designing a measuring weir. Sometimes free flow cannot be
ensured over the whole range of discharges to be encountered. Measurement is still
possible when flow is submerged (Fig. 7.4), but the conventional rating equations
are obviously not applicable because the discharge depends on both the down-
stream and upstream water levels. When submerged, the weir induces an energy
loss which manifests as an afflux of the upstream water level that can be related to
the discharge.
For the submerged condition, both water levels must be measured and accounted
for in the rating equation. Webber (1971) presents the Villemonte formula for these
conditions. This provides a correction to be applied to the discharge calculated
assuming free flow, i.e.

Fig. 7.4 Submerged


rectangular weir
h
hd
250 7 Flow-Measuring Structures

 3=2 !0:385
hd
Q ¼ Q 1  ð7:8Þ
h

in which Q* is the actual discharge, Q is the free discharge associated with h, and hd
is the height of the downstream water level above the weir crest. Equation (7.8)
suggests that the downstream water level will begin to affect the flow over the weir
as soon as it exceeds the height of the weir crest. Because of the empirical nature of
this equation, it should not be applied directly to other types of weir. Generally, the
accuracy of flow measurement is decreased by submergence, so it is advisable to
ensure free flow as far as possible. This can be done in practice by setting the crest
sufficiently high or including a downward step in the canal bed to lower the
downstream water level.
Example 7.1
A sharp-crested weir is installed in a 4.0 m wide rectangular canal. The weir extends
across the full width of the channel and its crest is 1.20 m above the bed. When the
discharge is 3.20 m3/s the uniform flow depth in the channel is 1.50 m. What is the flow
depth immediately upstream of the weir at this discharge?
Solution
The uniform flow depth downstream is above the weir crest, so the weir is submerged
and the discharge is related to the upstream flow depth by Eq. (7.8), i.e.
  3=2 0:385
Q ¼ Q 1  hhd (see Fig. 7.4)
with

2 pffiffiffiffiffi
Q ¼ Cd L 2gh3=2
3

h
Cd ¼ 0:611 þ 0:08
W
h
¼ 0:611 þ 0:08
1:2

L ¼ 4:0 m

hd ¼ yo  W ¼ 1:50  1:20 ¼ 0:30 m

Therefore

      !0:385
2 h pffiffiffiffiffiffiffiffiffiffiffiffi 3=2 0:30 3=2
3:20 ¼ 0:611 þ 0:08 4:0 2  9:8h 1
3 1:20 h

from which, by trial, h = 0.62 m


and so the flow depth upstream = h + W = 0.62 + 1.20 = 1.82 m.
7.2 Weirs 251

Triangular Weirs
The triangular weir or V-notch (Fig. 7.5) is used when accuracy is important at low
discharges, but a range of higher discharges must also be measured. The effective
flow width clearly varies with h (Fig. 7.6).
Assuming a small approach velocity and applying the same principles as for the
rectangular weir, the discharge can be shown to be given by

8 a pffiffiffiffiffi 5=2
Q¼ Cd tan 2gh ð7:9Þ
15 2

and Cd = Cc if the small approach velocity condition holds.


The most commonly used triangular weirs have a = 90o, for which Cd = 0.585
and then

Q ¼ 1:382h5=2 ð7:10Þ

in SI units.

Fig. 7.5 A laboratory triangular weir

Fig. 7.6 Triangular weir


geometry α h
252 7 Flow-Measuring Structures

General
Many equations for sharp-crested weirs have been presented over a long period of
time. These show considerable discrepancies because of their inherent empiricism
and the different conditions for which they were developed. Unless these conditions
are reproduced exactly in design, the equations will be inaccurate. It is, therefore,
meaningless to provide equations without details of the weir geometry for which
they are applicable. For this reason various authorities have developed standard
weir shapes with calibrated rating formulas. Henderson (1966) presents the British
Standard Specifications, for example. If a weir is constructed which does not
conform to a standard geometry, the rating equation can be expected to be inac-
curate and it should be calibrated. This should be done using an independent
measuring technique with significantly greater accuracy than that of the weir itself,
e.g. volumetric measurement. Further details concerning discharge coefficient val-
ues and design requirements are provided by Herschy (1995) and Ackers et al.
(1980).
Other Weir Geometries
Many other different weir geometries have been proposed for particular purposes,
and standard geometries and design details with associated rating equations have
been proposed by various authorities. A design commonly used in the United States
of America is the Cipolletti weir, which has a trapezoidal shape with the sides
sloping at 1:4 (horizontal to vertical). This shape is intended to compensate for the
reducing effect of the vertical sides of contracted weirs, as accounted for by
Eqs. (7.6) and (7.7). Design details and discharge coefficient values are provided by
Herschy (1995).
Other geometries have been proposed for contriving particular discharge char-
acteristics for controlling water levels, matching simple recording devices or
establishing certain error characteristics. Some of these unusual weirs, with their
discharge relationships, are described by Troskolanski (1960).
It is sometimes useful to extend the length of a weir crest if upstream water
levels are restricted, or if water levels are required to remain within a small range
over a wide range of discharges. This can be done by orienting the crest at an angle
across the channel or using a curved or labyrinth planform. The general weir
equations are applicable to such extended weirs, but the discharge coefficients may
need calibration, especially if the flow is made to converge significantly.
For flow measurement, it is desirable for the water level to vary significantly
over the range of measurement, and this can be enhanced by using compound weir
geometries where low flows are confined to narrow portions of the structure. It is
common in river measuring weirs to use a short rectangular or triangular section for
very low flows, and increasing lengths of rectangular weir as discharge increases
(Fig. 7.7). The rating curves for such weirs are complicated and usually require
individual calibration.
7.2 Weirs 253

Fig. 7.7 A compound river measuring weir

7.2.2 Broad-Crested Weirs

A broad-crested weir is simply an extended horizontal surface with sufficient length


and height above the channel bed to induce critical flow with parallel streamlines
(Fig. 7.8).
The discharge, Q, over the weir is given, by continuity, as

Q ¼ AV

A is the flow area at the critical section, given by

A ¼ byc

Energy
H
yc

Fig. 7.8 Broad-crested weir


254 7 Flow-Measuring Structures

where b is the width of the crest and yc is the critical flow depth. The velocity at the
critical condition is related to the flow depth by
pffiffiffiffiffiffiffi
V¼ gyc

Assuming negligible energy loss over the crest, the critical flow depth can be
related to the upstream energy level above the crest, H, by
2
yc ¼ H
3

The discharge is, therefore, given by

Q ¼ AV
 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi
2 2
¼b H g H
3 3

or
pffiffiffiffiffiffiffiffiffi
Q ¼ Cd b gH 3 ð7:11Þ

with Cd = (2/3)3/2 = 0.544.


Unfortunately, the assumptions underlying this derivation are not sufficiently
realistic for the result to have the accuracy required for flow measurement. The flow
separates at the sharp front edge of the crest, causing energy loss and sloping and
curvature of the streamlines where critical flow occurs at the point where the
separation zone is thickest. Various attempts have been made to describe this
phenomenon and calibrate Cd. Separation at the leading edge can be avoided by
rounding the corner and thereby increasing Cd. Herschy (1995) provides discharge
coefficients for both sharp and rounded broad-crested weirs.
The Crump Weir
The Crump weir overcomes the difficulties associated with normal broad-crested
weirs. It has a triangular longitudinal section (Fig. 7.9) and either a horizontal or
V-shaped crest, with lateral crest slopes of 1:10, 1:20 or 1:40. This weir is com-
monly used for river gauging and is one of the main types used in South Africa
(Wessels and Rooseboom 2009a). It is simple to construct, has a stable and rela-
tively constant discharge coefficient in the modular range and is not very sensitive
to submerged conditions. A typical installation is shown in Fig. 7.10.
The discharge relationship for horizontal Crump weirs under modular conditions
is given by BSI 3680 (1986) as
rffiffiffiffiffiffi
2 2
Q ¼ Cd gbH 3=2 ð7:12Þ
3 3
7.2 Weirs 255

total energy
V 2/2g

H
h
Dv
1
1 1
n
2 5
P
≥Hmax ≥2Hmax

upstream 2Hmax
measurement
(a) Longitudinal section (b) (alternative)
V-shaped crest

Fig. 7.9 Crump weir geometry

Fig. 7.10 A Crump weir showing structure (a) and in operation (b)

with
 
0:0003 3=2
Cd ¼ 1:163 1  ð7:13Þ
h

The equation is applicable for h  0.060 m, P  0.060 m, b  0.30 m, h/


P  3.5 and b/h  2.0. Wessels and Rooseboom (2009b) point out that within
these limitations Cd varies by less than 1%, and they suggest using a constant value
of 1.163 for practical applications.
For submerged conditions, a second flow depth measurement is required. The
standard design provides for measurement through tappings just downstream of the
crest, but Wessels and Rooseboom (2009b) suggest measurement downstream of
the hydraulic jump in rivers because the standard tappings tend to block with
sediment in river structures. The discharge equation for submerged conditions is
given by Ackers et al. (1980) as
256 7 Flow-Measuring Structures

rffiffiffiffiffiffi
2 2
Q ¼ Cd f gbH 3=2 ð7:14Þ
3 3

with Cd given by Eq. (7.13). The factor f accounts for submergence according to

H2
f ¼ 1:00 for  0:75 ð7:15Þ
H
 4 !0:0647
H2 H2
f ¼ 1:035 0:817  for 0:75\  0:93 ð7:16Þ
H H
 
H2 H2
f ¼ 8:686  8:403 for 0:93\  0:985 ð7:17Þ
H H

In Eqs. (7.15)–(7.17) H2 is the downstream energy level above the crest.


The discharge relationship for V-Crump weirs under modular conditions is

4 pffiffiffi 5=2
Q ¼ Cd gnH ð7:18Þ
5

if H is within the V of the crest, and


  !
4 pffiffiffi 5=2 DV 5=2
Q ¼ Cd gnH 1 1 ð7:19Þ
5 H

if H is above the top of the triangular part of the crest profile, i.e. H > DV, where DV
is the depth of the V. The discharge coefficient in both cases is CD = 0.633 for the
specified longitudinal profile, and n is the lateral crest slope as 1:n (Wessels and
Rooseboom (2009a, b) suggest n = 10 for normal situations). This calibration holds
for h  0.06 m, DV/P  2.5, and H/P  2.5.
For submerged conditions (H2/H < 0.78), both Eqs. (7.18) and (7.19) are cor-
rected by the factor f (as for the horizontal crest case), with f given by

f ¼ F1 for Hd \0:5 ð7:20Þ

f ¼ F2 þ ðF1  F2 Þð1:5  Hd Þ for 0:5  Hd  1:5 ð7:21Þ

f ¼ F2 for Hd [ 1:5 ð7:22Þ

with Hd = H/DV and

 4 !0:40147
H2
F1 ¼ 1:2 1  ð7:23Þ
H
7.2 Weirs 257

and

 4 !0:17353
H2
F2 ¼ 1:1019 0:914  ð7:24Þ
H

7.2.3 Advantages and Disadvantages of Weirs for Flow


Measurement

Weirs have a number of advantages over other means of measuring flows. These
include the following.
– They can measure a relatively wide range of flows with reasonable accuracy.
– They are simple in form and easy to construct.
– They can be used in combination with other structures, such as division struc-
tures, turnouts, drops, etc.
– To a certain degree, they can be made adjustable and portable for small canal
applications.
They do, however, have some disadvantages, including the following.
– They require a greater operating head than other devices, and, therefore,
introduce significant head losses to a system.
– Weir pools can clog or fill with sediment and must be kept clean to ensure
consistency of the rating formula.
– Weirs disrupt the passage of migrating fish.

7.3 Flumes

7.3.1 Throated (Venturi) Flume

A flume operates in a way similar to the Venturi meter in a pipe, i.e. the flow is
contracted over a short distance and variations in the head (or water surface in this
case) over the contraction can be related to discharge. Such a contraction can be
incorporated in a channel with any cross-sectional geometry. A typical contraction
and associated water surface profiles are shown in Fig. 7.11.
An expression for discharge in terms of flow depths can be obtained by equating
specific energy upstream (section 1) and at the throat (section 2), i.e.
258 7 Flow-Measuring Structures

throat

b1 b2 b3 = b1

(a) Plan geometry

energy line
V12/2g V32/2g
V22/2g

y1 y3
y2

(b) Submerged flow condition

energy line
V1 2/2g
V22/2g V32/2g
y1 y3
y2 = yc

(c) Unsubmerged flow condition

Fig. 7.11 Throated flume

E1 ¼ E2

i.e.

V12 V2
y1 þ ¼ y2 þ 2
2g 2g

from which

V22  V12
y1  y2 ¼
2g
7.3 Flumes 259

Also, from continuity,

Q ¼ V1 b1 y1 ¼ V2 b2 y2

from which

V2 b2 y2
V1 ¼
b1 y 1

which can be substituted into the energy relationship to give


  !
V2 b2 y 2 2
y1  y2 ¼ 2 1
2g b1 y 1

This can be rearranged to obtain the following expression for the velocity at
section 2:
pffiffiffiffiffipffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2g y1  y2
V2 ¼ rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ð7:25Þ
2
1 b 2 y2
b 1 y1

A coefficient of discharge, Cd, is introduced to account for the effect of


nonuniform velocity distribution and a small energy loss. The discharge is then
given by

Q ¼ V2 Cd b2 y2

and, substituting for V2 from Eq. (7.28) gives


pffiffiffiffiffi
2gCd b2 y2 pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Q ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffi y1  y2 ð7:26Þ
1  m2

in which

b2 y 2

b1 y 1

Cd has a value in the range of 0.96–0.99. Equation (7.26) is similar to the


equation for discharge through a pipe contraction but, because of the free surface,
the flow area and m are variable.
Equation (7.26) is general and applies to both flow conditions shown in
Fig. 7.11. The submerged condition is, however, unsatisfactory for practical flow
measurement, although it is sometimes unavoidable. In this case, the use of the
260 7 Flow-Measuring Structures

equation is complicated and two water level measurements are required. Also, y2 is
not very stable, making its measurement and the small difference (y1 – y2) inac-
curate. For these reasons, throated flumes are designed to operate unsubmerged or
‘free’, as far as possible. This is achieved by selecting a throat width that will
induce critical flow in the throat, backing up upstream, and a short reach of
supercritical flow downstream, followed by a hydraulic jump. Under these condi-
tions flow is critical in the throat and so

V2
Fr ¼ 1 ¼ pffiffiffiffiffiffiffi
gy2

which implies
pffiffiffiffiffiffiffi
V2 ¼ gy2

The discharge can, therefore, be expressed as

Q ¼ b2 y2 V2
pffiffiffi 3=2
¼ b2 gy 2

It is difficult to locate the critical section exactly, so it is usually the upstream


depth (y1) that is measured. It can be shown (e.g. Linford 1961) that the ratio y2/y1 is
constant for all discharges, although it varies with the ratio b1/b2. The discharge
can, therefore, be expressed in terms of y1. Therefore,
pffiffiffi 3=2
Q ¼ b2 gy1 x3=2

in which x = y2/y1. As before, to account for nonuniform velocity and energy


losses, a discharge coefficient (Cd) is introduced, and
pffiffiffi 3=2
Q ¼ Cd b2 gy1 x3=2 ð7:27Þ

In this equation, Cd has a value of about 0.95. The application requires an


evaluation of the x3/2 factor, which depends on b1/b2.
If the ratio b1/b2 is large, the approach velocity will be small and the upstream
water level (y1) will be approximately equal to the upstream energy level which can
be equated to the (critical) energy level in the throat, i.e.

V22 gy2
y1  y2 þ ¼ y2 þ
2g 2g
3
¼ y2
2
7.3 Flumes 261

from which

y2 2
x¼ ¼
y1 3

and therefore,
 3=2
pffiffiffi 3=2 2
Q ¼ C d b2 gy 1
3 ð7:28Þ
pffiffiffi 3=2
¼ 0:544Cd b2 gy1

The value of x3/2 in Eq. (7.28) (i.e. (2/3)3/2 = 0.544) applies for small values of
b2/b1. Linford (1961) provided a graph of x3/2 for a range of values of b2/b1, which
is reproduced in Fig. 7.12. In practice, b2/b1 will not be less than about 0.70.
For Eq. (7.27) to be used in practice, it must be ensured that the flume will
operate freely and that the downstream flow will not encroach into the throat. If this
happens, then the ratio y2/y1 will not be constant and Fig. 7.12 cannot be used. Two
flow depths have to be measured and the general Eq. (7.26) applied. For rectangular
channels, it may be assumed that free flow will result if y3/y1  0.8, where y3 is the
flow depth in the channel downstream.
The design of a free-flowing Venturi flume, therefore, requires an initial cal-
culation of the depth–discharge relationship for the canal. The throat width can then
be calculated to ensure that the upstream flow depth is within the range of the
measuring instrument and greater than the downstream depth at all discharges. In
many cases free flow can be maintained at high discharges but not at low dis-
charges. Rather than narrowing the throat and thereby raising the upstream depth
excessively for high discharges, the base of the flume can be raised to induce
critical flow or the channel bed level dropped on the downstream side.

0.70

0.65
x3/2

0.60

0.55

0.50
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
b2 /b1

Fig. 7.12 Relationship between depth and width ratios for Venturi flumes under unsubmerged
conditions (adapted from Linford 1961)
262 7 Flow-Measuring Structures

As for weirs, compound arrangements may be used to measure a wide range of


discharges. For example, a number of flumes may be installed side by side with
different base levels.

7.3.2 The Parshall Flume

The most widely used standard measuring flume was developed by Ralph Parshall
in 1920 (Parshall 1926). It has the advantage of having a standard shape that can be
used over a wide range of flows, with empirically determined discharge equations
valid over the whole range. Parshall flumes have been installed with throat widths
ranging from 75 mm to 15 m, to measure discharges from 0.25 l/s up to 15 m3/s.
They can be used to measure flows using one depth measurement when operating
unsubmerged, or using two depth measurements when operating submerged.
The geometry of the Parshall flume is rather complicated (Fig. 7.13), making it
expensive and difficult to construct. The flume consists of an entrance transition, a
converging section, a throat section, a diverging section and an exit section in plan.
In the longitudinal section there is a gradual rise of the base through the entrance
transition, a level base in the converging section, a downward slope through the
throat, followed by a gradual rise through the diverging section. This geometry
induces critical conditions at the beginning of the throat section, yielding a rela-
tionship between upstream depth and discharge as for throated flumes generally.
Dimensions for the different flume sizes are tabulated in Table 7.1 and associated
measuring ranges in Table 7.2.
For unsubmerged flow, the discharge equation is

Q ¼ Khu ð7:29Þ

in which h is the water level measured at the upstream measuring point, K is a


dimensional coefficient that depends on the throat width, and u is an exponent
between 1.522 and 1.607, depending on the flume size. According to Herschy
(1995), Eq. (7.29) applies for 0.015 m < h < 1.83 m, and with h measured at the
location specified in Fig. 7.13 and Table 7.1, i.e. for the dimension c measured
along the wall, equal to (b/3 + 0.813) m or 2/3A, upstream from the crest. The
values of K and u for use in Eq. (7.29) for the different standard sizes are given in
Table 7.2.
The following steps provide an approach for sizing a Parshall flume for
unsubmerged flow conditions:
• Identify all throat width sizes (b) that can measure the required discharge range.
• Check the suitability of sizes, starting with the smallest identified.
7.3 Flumes 263

M B L G

curved h1
R
entrance
(h2)

A c A
D b C

alternative
straight
entrance converging diverging
section section
entrance throat exit
transition section section

(a) Plan

submerged flow

E h h2
N
free flow
K
Y
X

(b) Section A-A

Fig. 7.13 Parshall flume geometry

– Check encroachment on the freeboard by calculating h for the maximum


discharge and applying energy conservation to a section immediately
upstream of the flume entrance.
– Check for submergence by ensuring that h2/h is less than the limit given in
Table 7.2. h2 is the height of the downstream water surface above the flume
bed at the location of h, given by h2 = yo – M/4, where yo is the uniform flow
depth downstream (this assumes a horizontal water level at the downstream
height).
Table 7.1 Dimensions of standard Parshall flumes (adapted from Herschy 1995). Note: flumes with throat sizes 0.076–2.44 m have approach aprons with 1:4
264

slopes and rounded entrances with R = 0.4 m for 0.076–0.228 m throats, R = 0.51 m for 0.30–0.90 m throats and R = 0.60 m for 1.2–2.4 m throats
b (m) D (m) M (m) C (m) B (m) L (m) G (m) E (m) N (m) K (m) A (m) c (m) X (m) Y (m)
0.025 0.167 0.305 0.093 0.357 0.076 0.204 0.153−0.229 0.029 0.019 0.363 0.241 0.008 0.013
0.051 0.213 0.305 0.135 0.405 0.114 0.253 0.153-0.253 0.043 0.022 0.415 0.277 0.016 0.025
0.076 0.259 0.305 0.178 0.457 0.152 0.30 0.305−0.610 0.057 0.025 0.466 0.311 0.025 0.038
0.152 0.396 0.305 0.393 0.610 0.30 0.61 0.61 0.114 0.076 0.719 0.415 0.051 0.076
0.229 0.573 0.305 0.381 0.862 0.30 0.46 0.76 0.114 0.076 0.878 0.588 0.051 0.076
0.305 0.844 0.381 0.610 1.34 0.61 0.91 0.91 0.228 0.076 1.37 0.914 0.051 0.076
0.457 1.02 0.381 0.762 1.42 0.61 0.91 0.91 0.228 0.076 1.45 0.966 0.051 0.076
0.610 1.21 0.381 0.914 1.50 0.61 0.91 0.91 0.228 0.076 1.52 1.01 0.051 0.076
0.914 1.57 0.381 1.22 1.64 0.61 0.91 0.91 0.228 0.076 1.68 1.12 0.051 0.076
1.22 1.93 0.457 1.52 1.79 0.61 0.91 0.91 0.228 0.076 1.83 1.22 0.051 0.076
1.52 2.3 0.457 1.83 1.94 0.61 0.91 0.91 0.228 0.076 1.98 1.32 0.051 0.076
1.83 2.67 0.457 2.13 2.09 0.61 0.91 0.91 0.228 0.076 2.13 1.42 0.051 0.076
2.13 3.03 0.457 2.44 2.24 0.61 0.91 0.91 0.228 0.076 2.29 1.52 0.051 0.076
2.44 3.4 0.457 2.74 2.39 0.61 0.91 0.91 0.228 0.076 2.44 1.62 0.051 0.076
3.05 4.75 0.457 3.66 4.27 0.91 1.83 1.22 0.34 0.152 2.74 1.83
3.66 5.61 0.457 4.47 4.88 0.91 2.44 1.52 0.34 0.152 3.05 2.03
4.57 7.62 0.457 5.59 7.62 1.22 3.05 1.83 0.46 0.229 3.50 2.34
6.10 9.14 0.457 7.31 7.62 1.83 3.66 2.13 0.68 0.31 4.27 2.84
7.62 10.67 0.457 8.94 7.62 1.83 3.96 2.13 0.68 0.31 5.03 3.35
9.14 12.31 0.457 10.57 7.92 1.83 4.27 2.13 0.68 0.31 5.79 3.86
12.19 15.48 0.457 13.82 8.23 1.83 4.88 2.13 0.68 0.31 7.31 4.88
15.24 18.53 0.457 17.27 8.23 1.83 6.10 2.13 0.68 0.31 8.84 5.89
7 Flow-Measuring Structures
7.3 Flumes 265

Table 7.2 Discharge characteristics of Parshall flumes (adapted from Herschy 1995)
Throat Discharge range K u Head range Modular
width b Minimum Maximum Minimum Maximum limit h2/h
(m) (l/s) (l/s) (m) (m)
0.025 0.09 5.4 0.0604 1.55 0.015 0.21 0.50
0.051 0.18 13.2 0.1207 1.55 0.015 0.24 0.50
0.076 0.77 32.1 0.1771 1.55 0.03 0.33 0.50
0.152 1.50 111 0.3812 1.58 0.03 0.45 0.60
0.229 2.50 251 0.5354 1.53 0.03 0.61 0.60
0.305 3.32 457 0.6909 1.522 0.03 0.76 0.70
0.457 4.8 695 1.056 1.538 0.03 0.76 0.70
0.610 12.1 937 1.428 1.550 0.046 0.76 0.70
0.914 17.6 1427 2.184 1.566 0.046 0.76 0.70
1.219 35.8 1923 2.953 1.578 0.06 0.76 0.70
1.524 44.1 2424 3.732 1.587 0.06 0.76 0.70
1.829 74.1 2929 4.519 1.595 0.076 0.76 0.70
2.134 85.8 3438 5.312 1.601 0.076 0.76 0.70
2.438 97.2 3949 6.112 1.607 0.076 0.76 0.70
(m3/s) (m3/s)
3.048 0.16 8.28 7.463 1.60 0.09 1.07 0.80
3.658 0.19 14.68 8.859 1.60 0.09 1.37 0.80
4.572 0.23 25.04 10.96 1.60 0.09 1.67 0.80
6.096 0.31 37.97 14.45 1.60 0.09 1.83 0.80
7.620 0.38 47.14 17.94 1.60 0.09 1.83 0.80
9.144 0.46 56.33 21.44 1.60 0.09 1.83 0.80
12.192 0.60 74.70 28.43 1.60 0.09 1.83 0.80
15.240 0.75 93.04 35.41 1.60 0.09 1.83 0.80

Parshall flumes can be used to measure discharges under submerged conditions,


which requires the additional measurement of the flow depth (h2) at the downstream
measuring point. As with all measuring structures this is not a desirable flow
condition, but may be unavoidable considering the relatively low modular limits of
Parshall flumes, especially for the smaller sizes, (see Table 7.2). For submerged
flow, corrections to the free flow calculated discharge can be made using the second
flow depth measurement, as presented by Aisenbrey et al. (1974), ISO 9826 (1992)
and the United States Bureau of Reclamation (1997).
Any flume which is to operate under free flow conditions only may be modified
downstream of the throat section without affecting the discharge relationship. It is
therefore quite common to install a Parshall flume at the head of a drop structure by
continuing the throat slope into the slope of the drop.
The complicated geometry and low modular limits of Parshall flumes have led to
their declining popularity in favour of simpler types, such as the cutthroat flume and
long-throated structures.
266 7 Flow-Measuring Structures

Example 7.2
A concrete-lined (n = 0.013) canal on a slope of 0.00060 has a trapezoidal cross section
with bottom width (B) of 2.0 m and side slopes (S) of 1.2H:1V. Select a Parshall flume
size to be installed in the canal at the construction stage to measure discharges ranging
from 0.22 to 2.40 m3/s. Modular flow should be ensured at all discharges and the
maximum allowable afflux is 0.20 m above the uniform flow depth. (The uniform flow
depths for the maximum and minimum discharges are 0.72 m and 0.18 m respectively.)
Solution
From Table 7.2, the minimum throat width for measuring the maximum discharge is
1.524 m and the maximum throat width for measuring the minimum discharge is
3.658 m. The selected size should be the smallest within this range that doesn’t cause
unacceptable afflux or submergence of the throat.
Try throat width (b) = 1.524 m.
From Table 7.2, the discharge equation is

Q ¼ 3:732h1:587

and the modular limit is 0.70.


From Table 7.1, the dimensions necessary for calculating afflux and submergence
conditions are (refer to Fig. 7.13).

A ¼ 1:98 m; c ¼ 1:32 m; D ¼ 2:30 m; and M ¼ 0:457 m

Check for allowable afflux:

H
afflux
E1
Eu hu h1
huniform
1
u
M

1
Eu ¼ E1 þ M
4
V12
E1 ¼ h1 þ
2g
 1=1:587
Q
h1 ¼ h ¼
3:732
 
2:40 1=1:587
¼ ¼ 0:757 m
3:732
Q
V1 ¼
A1
 
c ð D  bÞ
A 1 ¼ h1 b þ 2
A 2
 
1:32 ð2:30  1:524Þ
¼ 0:757 1:524 þ 2
1:98 2
¼ 1:545 m2
7.3 Flumes 267

Therefore
2:40
V1 ¼ ¼ 1:553 m/s
1:545
and
1:5532
E1 ¼ 0:757 þ ¼ 0:880 m
2  9:8
So
Vu2 1
E u ¼ hu þ ¼ 0:880 þ 0:457
2g 4
Q
Vu ¼
Au

Au ¼ Bhu þ Sh2u
¼ 2:00  hu þ 1:20  h2u
So
2:40
Vu ¼
2:00  hu þ 1:20  h2u

Therefore
 2
1 2:40
hu þ ¼ 0:994
2  9:8 2:00  hu þ 1:20  h2u

from which, by trial, hu = 0.963 m


Therefore the afflux = hu – huniform = 0.96 – 0.72 = 0.24 m, which exceeds the
allowable value of 0.20 m. The throat width of 1.524 m is therefore unacceptable
and the next larger size is tested.
Try throat width (b) = 1.829 m.
From Table 7.2, the discharge equation is

Q ¼ 4:519h1:595

and the modular limit is 0.70.


From Table 7.1, the dimensions necessary for calculating encroachment and sub-
mergence conditions are (refer to Fig. 7.13)
A ¼ 2:13 m; c ¼ 1:42 m; D ¼ 2:67 m; and M ¼ 0:457 m

Check for allowable afflux: Following the same procedure as for the previous trial
throat width, the afflux is 0.14 m, which is acceptable.
Check for submergence by ensuring the modular limit is not exceeded:

h h2
hd

M
268 7 Flow-Measuring Structures

If the channel bed downstream is at the same level as upstream, and the water level is
assumed to be horizontal on the downstream side,

1
h2 ¼ hd  M
4

where hd is the uniform flow depth.


At the maximum discharge, therefore,

1
h2 ¼ 0:72  0:457 ¼ 0:606 m
4
The throat flow depth is
 
2:40 1=1:595
h¼ ¼ 0:673 m
4:519

and the submergence ratio is

h2 0:606
¼ ¼ 0:900
h 0:673

which is greater than the modular limit (0.70) and, therefore, unacceptable.
At the minimum discharge,

1
h2 ¼ 0:18  0:457 ¼ 0:067 m
4
The throat flow depth is
 
0:22 1=1:595
h¼ ¼ 0:150 m
4:519

and the submergence ratio is


h2 0:067
¼ ¼ 0:447
h 0:150
which is less than the modular limit (0.7) and, therefore, acceptable.
A flume with a throat width of 1.829 m and the same bed level upstream and
downstream is unacceptable because submergence will occur at the higher range
discharges. A larger throat width will increase the submergence ratio, and so to
ensure modular flow the channel bed downstream of the flume should be lowered
(which is feasible at the construction stage). To ensure modular flow, h2 at the
maximum discharge should be no greater than
h2 ¼ h  modular limit ¼ 0:673  0:70 ¼ 0:471 m

Again assuming a horizontal water surface on the downstream side, the level of the
downstream channel bed should be below the flume crest by at least the uniform flow
depth less h2, i.e. 0.720 – 0.471 = 0.249 m, i.e. 0.249 – 0.457/4 = 0.135 m below
the original level.
7.3 Flumes 269

7.3.3 The Cutthroat Flume

The cutthroat flume was originally developed by Skogerboe et al. (1967) and has
been widely used in irrigation canals. The geometry is very simple (Fig. 7.14): it
has a flat base and straight and vertical converging and diverging side walls.
This simple geometry enables the same forms to be used for installations with
different throat widths. The flat base enables it to be placed in a lined canal with no
excavation, making it suitable for portable construction. Flume lengths ranging
from 0.45 to 2.7 m have been used to measure discharges up to 1.4 m3/s. The flume
can operate under free flow conditions, or with submergence up to 95%.
For free flow the discharge equation is

Q ¼ Chna1 ð7:30Þ

in which Q is the discharge in m3/s, ha is the water level measured at the upstream
piezometer tapping point, C is the free flow discharge coefficient, and n1 is an
empirical exponent. C is given by

C ¼ KW 1:025 ð7:31Þ

in which W is the throat width in metres, and K is an empirical coefficient. Both n1


and K depend on the flume length (L). Skogerboe et al. (1972) presented graphical
relationships for these, which can be expressed as

n1 ¼ 1:828L0:172 for 0:45\L\2:7 ð7:32Þ

and

K ¼ 3:414L0:492 for 0:84\L\2:7 ð7:33Þ

3 6
1 1
B = W + L/4.5

ha piezometer hb piezometer
W
tap tap

2L/9 5L/9
L/3 2L/3

L
Fig. 7.14 Cutthroat flume plan geometry
270 7 Flow-Measuring Structures

For submerged flow, two flow depth measurements are required. The discharge
is given by

C1 ðha  hb Þn1
Q¼ ð7:34Þ
ð log SÞn2

in which S is the submergence, hb/ha, where hb is the water level measured at the
downstream piezometer tapping point, and C1 is the submerged flow coefficient,
given by
C1 ¼ K1 W 1:025 ð7:35Þ

K1 and n2 also depend on the flume length according to graphical relationships


provided by Skogerboe et al. (1972), which can be expressed as

K1 ¼ 2:065L0:733 for 0:45\L\2:7 ð7:36Þ

and
n2 ¼ 1:452L0:055 for 0:84\L\2:7 ð7:37Þ

It is necessary to determine whether or not the flume is submerged at a particular


discharge so that the correct discharge equation can be selected. The transition
submergence, St, can be found by equating the right hand sides of Eqs. (7.30) and
(7.34) (which must give the same result at this condition), and solving for S. The
solution for St (%) is given graphically by Skogerboe et al. (1972) and can be
expressed for 0.45 < L < 2.7 as

St ¼ 67:54L0:158 ð7:38Þ

This value can be compared with the actual submergence to select the appro-
priate equation.
Growing acceptance of the cutthroat flume has led to more detailed examination
of the discharge equations and further developments and refinements in calibration
have been presented by Keller (1984), Skogerboe et al. (1993), Weber et al. (2007),
Manekar et al. (2007) and Torres and Merkley (2008). The equation for free flow
presented by Manekar et al. (2007) is based on experiments with a wide range of
sizes and is demonstrably more accurate than Eqs. (7.30) and (7.31) with coeffi-
cients given by Eqs. (7.32) and (7.33). They propose the following single equation
in non-dimensional terms for all flume sizes.
 1:7053
Q ha
pffiffiffi 1:5 ¼ 0:9169 ð7:39Þ
gWL L
7.3 Flumes 271

ha hb
yo
Δs

Fig. 7.15 Cutthroat flume with raised bed

Weber et al. (2007) found the discharge relationship for submerged conditions to
be improved if it was expressed in terms of specific energy rather than flow depth at
the two measurement points, but their results are based on experiments with just
one flume size. Torres and Merkley (2008) found that a single equation could be
used for both submerged and free flow conditions but, again, their results are for
just one flume size and are yet to be generalized.
As for the Parshall flume, it is highly desirable for the cutthroat flume to operate
freely over the whole range of discharges to be measured. It has been suggested that
this can be ensured by raising the flume base above the channel bottom by
Ds (Fig. 7.15) so that hb  haSt at all discharges. However, this will result in an
increased afflux of approximately the same amount, even if the throat width is also
increased.
If the modular flow is not possible within the afflux constraint, then the flume
should be designed for submerged flow with two water level measurements.
Equation (7.34) should then be used for afflux and submergence calculations.
Submerged flow causes increased afflux, so a wider throat would be necessary to
meet the same afflux limit as for modular flow.
As with most hydraulic structures, the design is an iterative process of analysis
and adjustment of an initially assumed geometry. A suitable procedure would be as
follows:
• Choose a value for B to be accommodated within the canal width.
• Assume a value for W which, together with B implies a value for L (see
Fig. 7.14).
• Use this value of L to obtain values for n1, K and St from Eqs. (7.32), (7.33) and
(7.38) if Eqs. (7.30) and (7.31) are to be used or just for St if Eq. (7.39) is to be
used.
• Calculate ha for the design discharge from Eqs. (7.30) or (7.39)
• Check possible encroachment on the canal freeboard by calculating the afflux, as
follows:
– Calculate the specific energy at the position of ha, i.e.

Va2
E a ¼ ha þ
2g
272 7 Flow-Measuring Structures

– where Va = Q/Aa with Aa calculated from the geometry shown in Fig. 7.14.
– Calculate the specific energy upstream of the structure as

Eu ¼ Ea

• Calculate the flow depth upstream from the specific energy and compare it with
the allowable value (yo + afflux allowed).
• Check for submergence, i.e. whether hb < ha St. If this condition is not satisfied,
and increased afflux can be accommodated, the bed should be raised by Ds = yo
– ha St, where yo is the uniform flow depth downstream (this assumes a level
water surface downstream from the position of hb and uniform flow). If the bed
is raised the afflux should be recalculated, with Ds added to Ea and compared
with the allowed value.
• If hb < ha St cannot be obtained within the allowable afflux, then modular flow
cannot be assured and the flume will have to be operated under submerged
conditions, with both ha and hb measured.

Example 7.3
A concrete-lined (n = 0.013) canal on a slope of 0.00050 has a trapezoidal cross section
with a bottom width of 1.40 m and side slopes (S) of 1H:1 V. Select a cutthroat flume
size to be installed in the canal after construction to measure discharges ranging from
0.30 m3/s to 1.20 m3/s. The maximum allowable afflux is 0.20 m above the uniform
flow depth. (The uniform flow depths for the maximum and minimum discharges are
0.63 m and 0.29 m, respectively.)
Solution
The size selection is determined by trial to ensure acceptable afflux and modular flow if
possible. Calculations are presented only for the successful trials.
A sensible base width (B) is the bottom width of the channel, i.e. 1.40 m. The throat
width (W) determines the afflux and the submergence condition: the smaller the throat,
the greater the afflux, the larger the throat the more likely it is to be submerged.
By trial, the smallest throat width with acceptable afflux is W = 0.81 m. Calculations are
shown for this width.
For B = 1.40 m and W = 0.81 m,
L ¼ 4:5ðB  W Þ ¼ 4:5ð1:40  0:81Þ ¼ 2:655 m
Using Eq. (7.39),
 1:7053
Q ha
pffiffiffi 1:5 ¼ 0:9169
gWL L

i.e.
 
Q ha 1:7053
pffiffiffiffiffiffiffi ¼ 0:9169
9:80:81  2:4751:5 2:475
7.3 Flumes 273

Therefore

Q ¼ 1:903h1:7053
a

Check for allowable afflux:


At the maximum discharge
   
Qmax 1=1:7053 1:20 1=1:7053
ha ¼ ¼ ¼ 0:763 m
1:903 1:903

Then the upstream flow depth can be calculate from

Eu ¼ Ea

where

Va2
E a ¼ ha þ
2g

with

Qmax
Va ¼
Aa

Aa ¼ ha Wa

and, referring to Fig. 7.14

2ðB  W Þ
Wa ¼ W þ
3
2ð1:40  0:81Þ
¼ 0:81 þ
3
¼ 1:203 m

So

Aa ¼ 0:763  1:203 ¼ 0:918 m2

1:20
Va ¼ ¼ 1:307 m/s
0:918

1:3072
Ea ¼ 0:763 þ ¼ 0:850 m
2  9:8

Therefore

Vu2
Eu ¼ hu þ ¼ 0:850
2g
274 7 Flow-Measuring Structures

with
Qmax
Vu ¼
Au

and
1
Au ¼ Bhu þ 2  Sh2u ¼ 1:40hu þ 1  h2u
2
Therefore,
 2
1 1:20
hu þ ¼ 0:850
2  9:8 1:40hu þ h2u

from which, by trial, hu = 0.829 m.


Therefore, Afflux ¼ hu huniform ¼ 0:830:63 ¼ 0:20; allowable:
Afflux is satisfactory for the maximum discharge, and will be less for lower dis-
charges; it is, therefore, satisfactory for the whole range.
Check for submergence:
Transition submergence is calculated from Eq. (7.38), i.e.

St ¼ 67:54L0:158 ¼ 67:54  2:6550:158 ¼ 78:8%

huniform 0:63
At maximum discharge submergence is S ¼ ¼ ¼ 83%
ha 0:763

huniform 0:29
At minimum discharge submergence is S ¼ ¼ ¼ 85%
ha 0:339

with
 
Qmin 1=1:7053
ha ¼
1:903
 
0:30 1=1:7053
¼
1:903
¼ 0:338 m

Therefore,

huniform 0:29
S¼ ¼ ¼ 86%
ha 0:338

Submergence exceeds the transition submergence over the full discharge range and
so modular flow is not possible for W = 0.81 m on a level bed. Raising the bed can
ensure modular flow only by allowing greater afflux. The flume must, therefore, be
designed for submerged flow, with two water level measurements.
The afflux is greater for submerged flow than for modular flow, and, therefore, a larger
throat width is required. By trial, the smallest throat width for acceptable afflux is
W = 0.85 m. Calculations are shown for this width. The results of Skogerboe et al.
(1972) are used for submerged flow.
For B = 1.40 m and W = 0.85 m,
7.3 Flumes 275

L ¼ 4:5ðB  W Þ ¼ 4:5ð1:40  0:85Þ ¼ 2:475 m

The discharge is calculated by Eqs. (7.34)–(7.37), i.e.

C1 ðha  hb Þn1

ð log SÞn2

with

C1 ¼ K1 W 1:025

K1 ¼ 2:065L0:733
¼ 2:065  2:4750:733 ¼ 1:063

Therefore,

C1 ¼ 1:063  0:851:025 ¼ 0:900

and

n1 ¼ 1:828L0:172
¼ 1:828  2:4750:172 ¼ 1:564

and

n2 ¼ 1:452L0:055
¼ 1:452  2:4750:055 ¼ 1:381

and

hb huniform 0:63
S¼ ¼ ¼
ha ha ha

Check for allowable afflux:

0:900ðha  0:63Þ1:564
At Qmax 1:20 ¼  1:381
 log 0:63
ha

from which, by trial, ha = 0.767 m.


Then the upstream flow depth can be calculated from

Eu ¼ Ea

where
276 7 Flow-Measuring Structures

Va2
Ea ¼ ha þ
2g

with
Qmax
Va ¼
Aa

Aa ¼ ha Wa

and, referring to Fig. 7.14


2ðB  W Þ
Wa ¼ W þ
3
2ð1:40  0:85Þ
¼ 0:85 þ
3
¼ 1:217 m

So

Aa ¼ 0:766  1:217 ¼ 0:932 m2

1:20
Va ¼ ¼ 1:288 m/s
0:932

1:2882
Ea ¼ 0:767 þ ¼ 0:852 m
2  9:8

Therefore,

Vu2
Eu ¼ hu þ ¼ 0:852
2g

with

Qmax
Vu ¼
Au

and

1
Au ¼ Bhu þ 2  Sh2u ¼ 1:40hu þ 1  h2u
2
Therefore,
 2
1 1:20
hu þ ¼ 0:852
2  9:8 1:40hu þ hu
2

from which, by trial, hu = 0.830 m


Therefore, Afflux ¼ hu huniform ¼ 0:830:63 ¼ 0:20; allowable:
Afflux is satisfactory for the maximum discharge, and will be less for lower dis-
charges; it is, therefore, satisfactory for whole range.
7.4 Long-Throated Structures 277

7.4 Long-Throated Structures

A flow measuring structure that is gaining popularity is one with a horizontal crest
(similar to a broad-crested weir) but with an upward approach slope at between 1:2
and 1:3 (Fig. 7.16).
The longitudinal profile can be used unchanged for any cross-sectional shape,
making it a very versatile design. The geometry may include a lateral contraction as
well as the raised sill, so these structures may be similar to both weirs and flumes.
The design and establishment of a discharge relationship are described by Bos et al.
(1986). The United States Bureau of Reclamation provides computer software for
design and calibration.

7.5 Errors and Measuring Ranges

When using weirs—or other measuring structures—it is necessary to know the error
involved, which varies with discharge, and the range of discharge over which errors
are acceptable.
The discharge error, eQ = eQ/Q where eQ is the difference between the estimated
and true discharge values, is mainly the result of the indication error in the mea-
surement of the water level, eh = eh/h. The error in reading the scale on the water
level gauge, eh is about ±1.0 mm to ±0.5 mm.
The discharge error, eQ, can be related to the water level indication error by
analysing the rating equation. For a rectangular sharp-crested weir

2 pffiffiffiffiffi
Q ¼ Cd L 2gh3=2 ð7:5Þ
3

or

Q ¼ Kh3=2

Fig. 7.16 Long-throated measuring structure profile


278 7 Flow-Measuring Structures

assuming Cd is constant. Differentiating this equation with respect to h gives

dQ 3 1=2
¼ Kh
dh 2

from which

3
dQ ¼ Kh1=2 dh
2

The error in discharge can be related to the indication error as follows:

dQ 32 Kh1=2 dh
eQ ¼ ¼
Q Kh3=2
3 dh
¼
2 h

i.e.

eQ ¼ 1:5eh ð7:40Þ

This means that the discharge error is 1.5 times the indication error obtained in
measurement of the water level. It is preferable to express the discharge error in
terms of the error in the scale reading; this can be done by substituting for h in
Eq. (7.40) from the rating equation, i.e.

Q2=3

K 2=3

gives

dhK 2=3
eQ ¼ 1:5
Q2=3
ð7:41Þ
1:5K 2=3
¼ eh
Q2=3

in which eh is about ±1 mm.


Equation (7.41) shows that the discharge error is greater at lower discharges than
at higher discharges.
This analysis can be used to determine the minimum head required on a weir to
ensure a specified accuracy for the discharge measurement.
The maximum discharge is defined by the maximum head that can be accom-
modated, and the measuring range is defined by
7.5 Errors and Measuring Ranges 279

 3=2
Qmin hmin
¼ ð7:42Þ
Qmax hmax

Example 7.4
If the maximum acceptable error in discharge is ±2% and the indication error is ±1
mm, determine the minimum head and discharge to be measured

a. by a 2.0 m long, 1.20 m high suppressed sharp-crested weir, and


b. a 90o triangular weir.

Solution

a. The relationship between discharge error and indication error for a rectangular weir is
given by Eq. (7.40), i.e.

eQ ¼ 1:5eh

Therefore,

eQ max 2%
eh max ¼ ¼ ¼ 1:33%
1:5 1:5

eh 1:0 mm
The minimum head is therefore hmin ¼ ¼ ¼ 75 mm
eh max 1:33%

and the minimum discharge is given by Eq. (7.5), i.e.

2 pffiffiffiffiffi 3=2
Qmin ¼ Cd L 2ghmin
3

with

h 0:075
Cd ¼ 0:611 þ 0:08 ¼ 0:611 þ 0:08 ¼ 0:616
W 1:20
Therefore

2 pffiffiffiffiffiffiffiffiffiffiffiffi
Qmin ¼ 0:616  2:0 2  9:8  0:0753=2 ¼ 0:075 m3 =s
3

b. Following the same procedure as for the rectangular weir, it can be easily shown that for a
triangular weir

eQ ¼ 2:5eh

Therefore,
280 7 Flow-Measuring Structures

eQ max 2%
eh max ¼ ¼ ¼ 0:80%
2:5 2:5

eh 1:0 mm
The minimum head is therefore hmin ¼ ¼ ¼ 125 mm
eh max 0:80%

and the minimum discharge is given by Eq. (7.10), i.e.

5=2
Qmin ¼ 1:382hmin ¼ 1:382  0:1255=2 ¼ 0:0076 m3 =s

Note that although the triangular weir appear less accurate, a rectangular weir would
have to be quite narrow and would not be able to accommodate the same range of
discharges with comparable accuracy.

Problems
7:1 The discharge through a triangular weir is described in SI units by

Q ¼ 1:382h5=2

If the scale on the water level gauge can be read with an accuracy of ±1.0
mm, determine the lowest discharge that can be measured by this weir with an
accuracy better than ±2.0%.
7:2 A long, rectangular, concrete-lined (n = 0.014) canal has a width of 3.0 m and
a slope of 0.00030. A sharp-crested weir is installed in the canal and a vertical
sluice gate is installed a short distance downstream of the weir. Both the weir
and the gate extend across the full width of the canal. Determine the flow
depths upstream and downstream of the two structures when the discharge is
2.5 m3/s with the following structure characteristics:
a. The weir crest is 2.6 m above the bed and the gate opening is 0.20 m.
b. The weir crest is 2.0 m above the bed and the gate opening is 0.20 m.
c. The weir crest is 1.0 m above the bed and the gate opening is 0.30 m.
7:3 A concrete (n = 0.013) trapezoidal canal has a bottom width of 1.5 m, side
slopes of 1.5H:1 V and a gradient of 0.00050. The canal is designed to convey
1.10 m3/s with a freeboard of 0.25 m.
a. Select a Parshall flume to measure flows ranging from 0.11 m3/s to
1.10 m3/s under free flow conditions. Assume the bottom of the canal to be
at the same level upstream and downstream of the flume.
b. Design a cutthroat flume for measuring the same range of discharges under
free flow conditions. Compare conditions using the calibrations originally
proposed by Skogerboe et al. (1972) and by Manekar et al. (2007).
c. If the same cutthroat flume design is used in another channel with the same
geometry but on a gradient of 0.00020, calculate the upstream flow depth
7.5 Errors and Measuring Ranges 281

for a discharge of 0.11 m3/s. Modify the design to ensure modular flow for
this discharge.
7:4 A cutthroat flume in a 2.0 m wide, rectangular, concrete channel has a throat
width of 1.5 m.
a. What is the discharge if ha = 0.949 m and hb = 0.713 m?
b. What is the accuracy of the discharge measurement for the conditions in
above if the scale on the piezometers can be read with an accuracy of
±1.0 mm?
c. What is the discharge if ha = 1.618 m and hb = 1.322 m?

References

Ackers, P., White, W. R., Perkins, J. A., & Harrison, A. J. M. (1980). Weirs and flumes for flow
measurement (3rd ed.). New York: Wiley.
Aisenbrey, A. J., Hayes, R. B., Warren, H. J., Winsett, D. L., & Young, R. B. (1974). Design of
small canal structures. United States Department of the Interior, Bureau of Reclamation.
Bos, M. G., Clemmens, A. J., & Replogle, J. A. (1986). Design of long-throated structures for flow
measurement. Irrigation and Drainage Systems, 1, 75–92.
BSI 3280. (1986). Part 4B: Measurement of liquid flow in open channels—triangular profile
weirs. London: BSI.
Henderson, F. M. (1966). Open channel flow. Macmillan.
Herschy, R. W. (1995). Streamflow measurement (2nd ed.). E & F N Spon.
ISO 9826 (1992) Measurement of liquid flow in open channels—Parshall and SANIIRI flumes.
Keller, R. J. (1984). Cut-throat flume characteristics. Journal of Hydraulic Engineering, 110(9),
1248–1263.
Linford, A. (1961). Flow measurement and meters (2nd ed.). E & F N Spon.
Manekar, V. L., Porey, P. D., & Ingle, R. N. (2007). Discharge relation for cutthroat flume under
free-flow condition. Journal of Irrigation and Drainage Engineering, 133(5), 495–499.
Parshall, R. L. (1926). The improved Venturi flume, transactions. American Society of Civil
Engineers, 89, 841–880.
Skogerboe, G. V., Hyatt, M. L., Anderson, R. K., & Eggleston, K. O. (1967). Design and
calibration of submerged open channel measurement structures. Part 3—Cutthroat flumes
(pp. 1–37). Utah Water Research Laboratory, College of Engineering, Utah State University.
Skogerboe, G. V., Bennett, R. S., & Walker, W. R. (1972). Installation and field use of cutthroat
flumes for water measurement. Water Management Technical Report No 19, Colorado State
University, Colorado, USA.
Skogerboe, G. V., Ren, L., & Yang, D, (1993). Cutthroat flume discharge ratings, size selection
and installation. International Irrigation Center, Department of Biological and Irrigation
Engineering, Utah State University.
Torres, A., & Merkley, G. P. (2008). Cutthroat measurement flume calibration for free and
submerged flow using a single equation. Journal of Irrigation and Drainage Engineering, 134
(4), 521–526.
Troskolanski, A. T. (1960). Hydrometry. Oxford: Pergamon Press.
United States Bureau of Reclamation. (1997). Water measurement manual.
Webber, N. B. (1971). Fluid mechanics for civil engineers. London: Chapman and Hall.
Weber, R. C., Merkley, G. P., Skogerboe, G. V., & Torres, A. F. (2007). Improved calibration of
cutthroat flumes. Irrigation Science, 25, 361–373.
282 7 Flow-Measuring Structures

Wessels, P., & Rooseboom, A. (2009a). Flow gauging in South African rivers, Part 1: An
overview. Water SA, 35(1), 1–9.
Wessels, P., & Rooseboom, A. (2009b). Flow gauging in South African rivers, Part 2: Calibration.
Water SA, 35(1), 11–19.

Further Reading

Abt, S. R., Cook, C., Staker, K. J., & Johns, D. D. (1992). Small Parshall flume rating correction.
Journal of Hydraulic Engineering, 118(5), 798–803.
Charlton, F. G. (1978). Measuring flows in open channels: A review of methods. CIRIA Report 75.
Clemmens, A. J., Wahl, T. L., Bos, M. G., & Replogle J. A. (2010). Water measurement with
flumes and weirs. Water Resources Publications.
Gill, M. A. (1985, August). Flow measurement by triangular broad-crested weir. Water power
and dam construction (pp. 47–49).
Keller, R. J., & Mabbett, G. O. (1987). Model calibration of a prototype cutthroat flume. Journal
of Hydraulic Research, 25(3), 329–340.
Ramamurthy, A. S., Rao, M. V. J., & Auckle, D. (1985). Free flow discharge characteristics of
throatless flumes. Journal of Irrigation and Drainage Engineering, 111(1), 65–75.
Rao, N. S. L. (1975). Theory of weirs. Advances in Hydroscience, 10, 310–406.
Samani, Z., & Magallanez, H. (1993). Measuring water in trapezoidal channels. Journal of
Irrigation and Drainage Engineering, 119(1), 181–186.
Chapter 8
Intake Structures

8.1 Introduction

An intake structure is a transition device through which flow is diverted from a


source, such as a river, reservoir or the ocean, into a conduit, which may be a canal
or a pipe. Its primary functions are to admit and regulate (and possibly also to
measure) water from the source, and to ensure satisfactory quality of the admitted
water (in terms of sediment content, temperature, chemical and biological
characteristics).
An intake may be an integral part of an impoundment or diversion structure,
such as a dam or weir, or it may be entirely independent. In either case, it will
generally incorporate a conduit with protective works and screens at the entrance
section and gates or valves for regulating the flow. The location and design of
intake works must ensure reliability of operation and satisfactory water quality. The
factors affecting these considerations depend on the nature of the water body from
which the supply is taken. For example, designs would be very different for
extraction of water from shallow or deep lakes and reservoirs, and large and small
rivers.

8.1.1 Reservoir Intakes

Intakes in large lakes or reservoirs should be located in as deep water as possible to


avoid the mixed zone resulting from wind and convection. In impounded reservoirs,
intakes are, therefore, usually at or near the dam wall or embankment. This also
ensures operation at low water levels.
Intakes for deep reservoirs should be designed to enable water to be drawn off at
different depths. It is sometimes desirable to avoid drawing surface water into the
supply because of excess turbidity during high winds, high temperatures during hot

© Springer Nature Switzerland AG 2020 283


C S James, Hydraulic Structures,
https://doi.org/10.1007/978-3-030-34086-5_8
284 8 Intake Structures

summer months and the occurrence of plankton near the surface. Many microor-
ganisms that can produce objectionable odours also live near the surface because of
their requirement for light. Water at great depths can be rich in carbon dioxide, iron
and manganese, and may have excessive colour and hardness. These undesirable
characteristics can be avoided by drawing water from near the surface, although it
may sometimes be advantageous to draw iron-rich bottom water in order to reduce
the quantity of coagulant used for treatment. These considerations are important not
only for water supply systems but also for releases from large reservoirs into
downstream river reaches where environmental constraints may dictate water
quality and temperature ranges; these can often be satisfied by mixing water from
different depths through multilevel intakes.
Deep reservoir intakes are commonly constructed as towers some distance from
the shore, often with connecting bridges. Intake ports are provided at two or three
levels, depending on the maximum depth. Gates and screens should be installed
inside the tower to facilitate maintenance. The tower may also include two parts
with duplicate screens and outlets to enable uninterrupted service during mainte-
nance and improve reliability.

8.1.2 River Intakes

In large rivers intakes should be located so that the ports are always submerged to
avoid floating debris and air entrainment. The ports should also be well clear of the
bed to avoid intake of coarse sediment. River intakes are, therefore, usually located
in the deepest part of the stream, particularly if large variations of river stage occur,
and water is conveyed to the shore through a pipe along the bed or a tunnel below
the bed.
For small supplies, the intake may be simply a screened bell mouth at the end of
a pipe. For large supplies, tunnels are often required and intake towers are con-
structed which incorporate vertical shafts to connect to the tunnel. The intake ports
in the tower walls are fitted with gates and often with screens. If the river stage is
relatively constant the intake may be constructed on the shore, obviating the need
for a tunnel and resulting in a much simpler structure.
Most large river intakes require pumping and pumping stations are usually
constructed on the river bank adjacent to the intake. The intake and pump station
can often be combined if the intake can be built on the shore. The design of pump
sumps and intakes present special problems and is discussed in greater detail in
Sect. 8.3.
Intakes in small streams often requires the construction of a diversion barrage or
weir. This ensures a sufficient depth of water for abstraction and also allows some
settling of sediment, although the pools often fill quite rapidly. Intakes of this type
are often in remote locations and maintenance is intermittent, so measures must be
taken to avoid clogging by debris and sediment. Screening is, therefore, important.
8.1 Introduction 285

Bar racks and mesh screens are frequently used with small stream intakes, the
former to provide protection for the latter from heavy debris. Various types of flat
and drum screens are available.

8.2 River Intake Design for Sediment Control

Many problems associated with river intake structures are related to the sediment
being transported by the river. The structures should be designed so that they will
not clog with deposited sediment, initiate local erosion, or affect the morphological
equilibrium of the natural channel. For water diversions, they should also divert the
required quantity of water to the supply aqueduct with minimum sediment content
(diversions intended to extract sediment from a river are not considered here).
This section describes measures for minimizing the intake of sediment and
removing the bulk of diverted sediment from the water at or close to the abstraction
point. Complete removal of sediment for domestic or industrial water supply must
be performed by fine screening or filtration, which is not usually done at the intake
location. For irrigation supply it is rarely justifiable to remove all sediment from the
water, in fact, it is frequently desirable to retain fine sediment as this is beneficial in
sealing unlined canal laterals and may improve the soil texture and fertility of
croplands. Obviously, aqueducts conveying sediment-laden water must be carefully
designed to ensure that excessive deposition does not take place. This is difficult as
it is usually impossible to maintain sufficiently high velocities at all times and at all
points in the system. It must also be borne in mind that excluding sediment at an
intake and returning it to the stream can create instability problems in the natural
channel downstream, particularly if a large proportion of the stream flow is
diverted. The sediment concentration below the diversion point would be increased
and could disturb the equilibrium state of water discharge, sediment load and
channel morphology. Expensive maintenance or corrective action in the stream, or
both, may then be required.
The nature and extent of sediment control problems at intakes depend on the
type and quantity of sediment transported by the river, which in turn depends on the
rate of supply of sediment to the river and its transport capability. For rivers
transporting relatively coarse sediments, the supply rate generally exceeds the
transport capability; some deposition, therefore, occurs and the actual rate of
transport is defined by the transport capability, which can be related to local flow
conditions through sediment transport equations. For rivers transporting relatively
fine sediments the supply rate is usually less than the transport capability. The
actual transport rate is, therefore, determined by the supply rate which depends on
catchment characteristics and conditions as well as the history of
sediment-producing flows. Under these circumstances it is not possible to correlate
sediment concentration with discharge reliably, and it is extremely difficult to
establish sediment discharge rates. Site-specific measurement is advisable in all
cases.
286 8 Intake Structures

Water in aqueducts can be kept relatively sediment-free by diverting sediment at


the headworks, removing it from the aqueduct by ejectors, or inducing deposition in
settling basins from where it can be removed mechanically or by sluicing. The
structures required for diversion or ejection are very difficult to analyse, and most
designs are based on experience gained from model studies or existing prototype
performance. A basic understanding of sediment and relevant flow behaviour is,
however, essential when designing a general configuration. Of particular practical
importance are the vertical distribution of suspended sediment through the flow and
the response of near-bed sediment to flow curvature induced by bends.

8.2.1 Vertical Sediment Distribution

The total sediment load in a river can be considered to consist of two main com-
ponents, namely suspended load and bedload. (Very fine sediment, not represented
in the river bed material, is known as wash load and cannot be excluded at intakes.)
The bed load consists of the coarser size fractions which are moved along the bed or
by saltation under the influence of traction forces. The suspended load consists of
fine particles that can be maintained within the flow by the action of turbulence. The
vertical distribution of suspended material depends on both particle size and flow
characteristics.
The vertical concentration distribution of suspended material can be described
by applying the ‘diffusion analogy’. According to this analogy, if sediment con-
centration varies through the flow, then particles will tend to move from regions of
high-concentration to regions of low concentration at a rate proportional to the
concentration gradient. The proportionality between transfer rate and concentration
gradient is related to the momentum diffusivity (or eddy viscosity), which is
determined by the turbulence structure of the flow. The concentration is highest
near the bed, so particles will tend to be diffused upwards. Particles also tend to
settle under the influence of gravity at a rate equal to the concentration multiplied
by the particle settling velocity. An equilibrium distribution of suspended sediment
concentration can be derived by considering a mass balance of material in an
element of fluid under the influence of these two transfer mechanisms. Under
equilibrium conditions, the net transfer is zero and, therefore,

@C
0 ¼ wC þ es ð8:1Þ
@y

in which C is the concentration, w is the particle settling velocity (positive


downwards), es is the diffusivity for sediment, and y is the height above the bed.
The diffusivity for sediment is assumed to be similar to the momentum diffu-
sivity for small particles. Its magnitude and variation with height above the bed can
be determined through Prandtl’s mixing length theory and described by
8.2 River Intake Design for Sediment Control 287

y
es ¼ ju ðD  yÞ ð8:2Þ
D

in which j is the von Karman constant, which has a value of 0.40 for clear water,
D is the total flow depth and u* is the shear velocity, defined by
rffiffiffiffiffi
so
u ¼ ð8:3Þ
q

in which so is the bed shear stress, given by

so ¼ qgRS ð8:4Þ

in which q is the water density, g is gravitational acceleration, R is the hydraulic


radius (which can be assumed to equal the flow depth for wide channels) and S is
the energy gradient (which is equal to the bed gradient for uniform flow).
Substituting Eq. (8.2) into Eq. (8.1) and integrating yields
 w=ju
C Dy a
¼ ð8:5Þ
Ca y Da

commonly known as the ‘Rouse equation’. This equation gives the sediment con-
centration, C, at any height, y, in terms of the concentration Ca at a specified height
a. Ca is usually estimated very close to the bed, using an appropriate bed load
equation and assuming a thickness for the bed layer for a. The equation shows that
very fine particles are distributed fairly uniformly over the flow depth and that as the
particle size increases more sediment is transported close to the bed (Fig. 8.1).
In a natural river, where the total sediment load consists of various sizes of
sediment moving as bedload and suspended load, the transported sediment will be
concentrated in the lower part of the flow. This can be used to advantage when
designing sediment removal structures or devices.

Fig. 8.1 Concentration


distributions for fine and
coarse sediments

fine
y coarse D

a
C Ca
288 8 Intake Structures

Example 8.1
A long, wide channel has a slope of 0.0020. When the discharge is 2.25 m3/s/m the flow
depth is 1.0 m.
Compare the suspended sediment concentration profiles over the flow depth for sedi-
ment grain sizes of 0.25, 0.125 and 0.062 mm with corresponding settling velocities of
0.0343, 0.0118 and 0.0031 m/s. Assume the sediment concentration to be 0.25 at 1.0 cm
above the bed for all sizes.
Solution
The concentration profile is given by Eq. (8.5), i.e.
 w=ju
C Dy a
¼
Ca y Da

with
Ca ¼ 0:25
a ¼ 0:010 m
D ¼ 1:0 m
pffiffiffiffiffiffiffiffiffiffiffi
u ¼ gDSo for uniform flow in a wide channel
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
¼ 9:8  1:0  0:0020 ¼ 0:140 m/s

Then C can be calculated for 0.010 m < y < 1.0 m.


For example, for y = 0.50 m
w 0:0343
For d ¼ 0:25 mm; w ¼ 0:0343 m=s; ¼ ¼ 0:613
ju 0:40  0:140
 w=ju
Dy a
C ¼ Ca
y Da
 0:613
1:00  0:50 0:010
¼ 0:25 ¼ 0:015
0:50 1:00  0:010

w 0:0118
For d ¼ 0:125 mm; w ¼ 0:0031 m=s; ¼ ¼ 0:211
ju 0:40  0:140
 w=ju
Dy a
C ¼ Ca
y Da
 0:211
1:00  0:50 0:010
¼ 0:25 ¼ 0:095
0:50 1:00  0:010

w 0:0031
For d ¼ 0:062 mm; w ¼ 0:0118 m=s; ¼ ¼ 0:055
jux 0:40  0:140
 w=ju
Dy a
C ¼ Ca
y Da
 0:055
1:00  0:50 0:010
¼ 0:25 ¼ 0:194
0:50 1:00  0:010
8.2 River Intake Design for Sediment Control 289

Concentrations for other heights are calculated similarly and plotted and tabulated
below.
1.0

0.8

0.6
y (m)

0.4

0.2

0.0
0.00 0.05 0.10 0.15 0.20 0.25 0.30
C
d = 0.25 mm d = 0.125 mm d = 0.062 mm

y (m) d = 0.25 mm d = 0.125 mm d = 0.062 mm


0.01 0.250 0.250 0.250
0.05 0.091 0.177 0.228
0.10 0.058 0.151 0.219
0.15 0.043 0.137 0.213
0.20 0.035 0.127 0.209
0.25 0.029 0.120 0.206
0.30 0.025 0.114 0.203
0.35 0.022 0.108 0.201
0.40 0.019 0.104 0.198
0.45 0.017 0.099 0.196
0.50 0.015 0.095 0.194
0.55 0.013 0.091 0.192
0.60 0.012 0.087 0.189
0.65 0.010 0.084 0.187
0.70 0.009 0.080 0.185
0.75 0.008 0.076 0.182
0.80 0.006 0.071 0.179
0.85 0.005 0.066 0.176
0.90 0.004 0.060 0.172
0.95 0.002 0.051 0.165
0.98 0.001 0.042 0.156
1.00 0.000 0.000 0.000
290 8 Intake Structures

8.2.2 Bed Load Movement Around Bends

The behaviour of flow around bends is a very important consideration when


designing artificial canals or analysing river behaviour. In artificial canals, bends
must be designed with radii sufficiently large to avoid excessive energy losses and
to ensure satisfactory approach conditions for measuring structures and control
structures. The development of tight meander bends in natural rivers may cause
increased flood levels upstream and channel stability problems in addition to
deleterious effects on structures. Bends also induce local variations in flow velocity
which influence the location and movement of transported sediment and are,
therefore, important features to be considered in designs for sediment exclusion.
An analysis of energy component variations across a bend section shows that the
water surface rises towards the outside of the bend, and that the depth-averaged
velocity increases towards the inside of the bend. The actual flow distribution is
influenced by resistance, and the maximum velocity is near the inner bank at the
entrance to the bend but shifts towards the outer bank through the bend. Also,
because of the increasing depth (and hence hydrostatic pressure) towards the outer
bank and the low velocity near the bed, an inwardly directed pressure gradient
exists close to the bed, causing inwardly directed flow. Balancing this is an out-
wardly directed flow near the water surface. The net result is a secondary, circu-
latory flow pattern over the cross section and a helical flow through the bend
(Fig. 8.2). Raudkivi (1993) provides a more detailed and quantitative description,
with results for predicting water surface superelevation, velocity direction and
variation across the section, the elevation of the maximum velocity location (which
is below the water surface) and the angular location of the completion of secondary
flow development.
This flow pattern has a very important effect in channel forming processes in
natural rivers: sediment tends to be scoured at concave banks and deposited as bars
on convex banks. The effect is extremely important for sediment exclusion design.
As has already been demonstrated, most sediment tends to be transported close to
the bed and at a bend will, therefore, be directed towards the inner bank and away
from the outer bank. This phenomenon sometimes referred to as ‘bed load sweep’,

flow near bed

flow near surface

outer bank, inner bank,


erosion deposition

(a) Cross section (b) Plan

Fig. 8.2 Flow pattern through a bend


8.2 River Intake Design for Sediment Control 291

has been exploited for centuries. The remains of irrigation systems built by ancient
Mediterranean civilizations show that branch canals were always connected to the
main channel at straight sections or on the outsides of curves. The approach is no
less useful today and ensures that sediment concentrations in a branch channel
connected to the outside of a curve are much less than in the main channel
(Hernandez 1969).
In addition to choosing a suitable location for an intake, the angle of diversion
between the river and the diversion channel must be selected as this has a significant
effect on the amount of sediment diverted. Any diversion at an angle to the river
flow creates an effective curvature in the flow (Fig. 8.3). The water near the surface
has a higher velocity and hence more momentum than water near the bed and will,
therefore, be less easily diverted and will tend to continue with the main river flow.
The most easily diverted water will be the slower moving water near the bed, which
also has a higher sediment concentration. If the quantity of diversion is significantly
relative to the river flow, the take-off effectively creates a curve, resulting in bedload
sweep into the diversion channel. As well as causing undesirable sediment content
in the deviated water, deposition could occur in the diversion channel as an effective
point bar and this could eventually cause blocking.
The amount of sediment diverted depends on the head difference between the
main and diversion channels and the diversion angle. The flow will always enter the
diversion channel at an angle less than 90° so the ideal angle is something smaller
than 90°. It is difficult to define an optimal angle because it depends on the
diversion ratio, i.e. the ratio of discharges in the diversion and the river, and also on
the position of the diversion in a bend. The optimum angle increases as the
diversion ratio decreases. The best design approach is to conduct model tests and
choose the best angle for the dominant diversion ratio or for the condition pro-
ducing maximum bedload discharge. If a model study is not undertaken and no
sediment deviator is to be installed, an angle of diversion between 10° and 45° is
recommended (Avery 1989).

flow near surface

flow near bed

diversion angle
sediment
deposition diversion
channel
Fig. 8.3 Sedimentation at diversion
292 8 Intake Structures

Results of model studies show that, for example, for the main and diversion
channel beds at the same level and with an angle of diversion of 30° and a diversion
ratio of about l:4, about 50% of the bedload would enter the branch. For a discharge
ratio greater than about 50% practically all the bedload would be diverted
(Hernandez 1969).

8.2.3 Sediment Exclusion Structures

Three types of structures are commonly used for restricting the amount of sediment
in a supply aqueduct: Sediment diverters are intended to prevent most of the sed-
iment from entering the aqueduct; sediment ejectors remove sediment already in the
aqueduct; settling basins allow deposition of very fine sediment in the aqueduct for
subsequent removal.
The appropriate structure for a particular intake depends on the particular con-
ditions. Sediment diversions are intended mainly for the exclusion of bedload.
Rivers carrying most sediment in suspension require settling basins. Sediment
ejectors would be appropriate where the sediment is transported in suspension in the
river but becomes mainly bedload in the aqueduct because of the reduced trans-
porting capacity.
The more common approaches to sediment exclusion are described in the fol-
lowing paragraphs; further details are provided by Vanoni (1975), Avery (1989)
and Raudkivi (1993).
Sediment Diverters
A sediment diverter is a device or structure incorporated in the intake headworks for
preventing most of the sediment, especially bedload and coarser suspended frac-
tions, from entering the aqueduct. Many different arrangements have been devel-
oped either from theoretical analyses, model tests or observations of actual systems
over many years. Most take advantage of the vertical distribution of suspended

Fig. 8.4 A river diversion structure looking downstream (a) and upstream (b)
8.2 River Intake Design for Sediment Control 293

sediment and the bedload sweep phenomenon described above. The most basic
application is simply to locate the intake on the outside of a bend and as high above
the bed as practically possible. In the diversion structures shown in Fig. 8.4 the
intake (through trashracks) is on the outside of the bend with a gated sluiceway
immediately downstream and in line with the weir to direct the flow and keep the
intake clear by sluicing local bed load. More sophisticated designs applying these
principles have been developed, such as those described in concept below.
Detailed design is made difficult by the uncertainty and variability of future flows
and sediment concentrations, as well as the limitations of analytical and compu-
tational approaches. Physical model studies provide the most reliable basis for the
design. Although they are not usually able to reproduce movement and distribution
of sediment quantitatively, they are useful for providing qualitative evaluations and
comparisons between different design options.
Training Walls and Guide Banks: create an artificial curve in the flow in the
vicinity of the intake in order to take advantage of the bedload sweep phenomenon.
This approach is often used when the intake is constructed together with a weir-type
structure and where river flow is limited, with most of the excess used for sluicing,
such as illustrated in Fig. 8.5. The training walls induce curvature in the flow,
effectively placing the canal intake on the outside of a bend so that relatively clear
water is taken into the canal and high-concentration bedload proceeds to the
sluiceway. The canal intake sill is set above the elevation of the sluiceway and has a
projecting lip to further ensure that water with higher sediment concentration is
washed through the sluiceway. The spacing of the training walls must be designed
to ensure sufficient velocity to flush sediment through at low discharges. It should
be noted that diversion weirs generally fill with sediment after some time, but a low
flow channel to the diversion works can be maintained by regular sluicing.

sluiceway

canal
training walls
weir crest
A
to canal
river
A flow
to
training walls sluiceway

(a) Plan (b) Section A-A

Fig. 8.5 Training wall configuration


294 8 Intake Structures

river banks
guide bank

induced
sluiceway deposition
guide
bank

canal

Fig. 8.6 Guide bank configuration

Guide banks (Fig. 8.6) produce the same effect as training walls by forcing a
realignment of the river channel to create a convex bank at the intake location.
Where intakes are required on both ends of a weir the required curvature for both
intakes can be created by forming a suitably shaped central island upstream of the
weir.
In all applications of training walls and banks, the point of diversion must be
positioned far enough downstream along the curve for the helicoidal flow pattern
producing bedload sweep to have become established. This position depends on
size and quantity of bedload, flow velocity and channel geometry and can only be
determined reliably by model studies.
The Pocket and Divider Wall: arrangement in association with a barrage
structure induces deposition ahead of the intake and diverts flow above the
high-concentration region near the bed. A divider wall upstream of the barrage
structure creates a ‘pocket’ in front of the intake (Fig. 8.7).
The divider wall directs part of the river flows into the sluiceway (with convex
curvature approaching the intake) and the remainder over the barrage structure. The

Fig. 8.7 Pocket and divider sluiceway


wall configuration

canal

barrage crest

divider river
wall flow
pocket
8.2 River Intake Design for Sediment Control 295

pocket acts as a ponding area for deposition of sediment which can be sluiced away
either intermittently or continuously, depending on available stream flow. Whether
sluicing is continuous or intermittent, the headgate invert must be set well above the
pocket floor—at least one-third of the flow depth is recommended. This provides
temporary storage for bedload if intermittent sluicing is used and a bedload
exclusion device if continuous sluicing is used.
The length of the divider wall is determined largely by the ratio of flow in the
stream to that being diverted. If the stream flow is limited, the divider is usually
extended as a training wall to create a curved approach and the length and geometry
should be determined by modelling. If the stream flow is more than sufficient to
provide for the abstraction and sluicing, the divider wall acts merely as a splitter and
to isolate the pocket area. In such cases, the divider wall need not extend upstream
beyond the headgate.
The width of the pocket is best determined by physical modelling. The cross
section should be large enough to convey the abstraction and sluicing water, but
small enough to maintain sluicing velocities for dominant or design flow condi-
tions. A slight convergence (up to 1:10) improves scouring action, but the pocket
must not be too wide at the entrance or sluicing will be ineffective.
The diversion barrage is often a gated structure, in which case the careful
operation of the gates can greatly enhance the performance of the sediment division
works. By opening the gates furthest from the headworks the most and those nearest
the headworks the least, an effective curvature of the flow will be created, drawing
sediment towards the inside of the curve and away from the headgate. If there are
intakes on both sides of the barrage it would be advantageous to open the gates in
the centre of the barrage the most.
The relative openings of the sluice gates and barrage gates near the divider wall
also affect the amount of sediment excluded. The path of the bedload depends on
the ratio of the average flow velocities on the riverside and the pocket side of the
divider wall. If this ratio is greater than one the bedload tends to be directed to the
riverside of the wall, and vice versa.
A divider wall can be effective without the provision of a pocket for relatively
small diversion proportions (Fig. 8.8). The upstream orientation of the approach to
the intake induces flow curvature and bedload sweep towards the main river. This
effect can be enhanced by a suitably positioned island or guide walls. A sluicing
facility can also be included to reduce sediment build-up ahead of the intake, if
necessary.
Sand Screens or Skimming Weirs: are simply low barricade walls which allow
the relatively clear water near the surface to flow over to the intake, and direct the
bed load back into the flow of the stream, i.e. they provide a rudimentary
exploitation of the vertical sediment concentration distribution. The tops of the
walls are set at a level that will effectively divert the sediment for the dominant or
selected design conditions. Provision for flashboards on the tops of the walls is
often made to adapt the configuration to changes in stream flow.
296 8 Intake Structures

canal

sluiceway
divider wall

island

Fig. 8.8 Simple divider wall configuration

Sand screens are most effective where the sediment load is predominantly
bedload. To be effective they must allow sufficient flow past the point of diversion
to transport the bed load so that deposited ‘ramps’ of material cannot form, which
would make the screen ineffective.
In some cases, the efficiency of a sand screen has been increased by installing a
thin horizontal projection at the top of the wall for a short distance upstream.
Guide Vanes: are used to induce local helicoidal flow patterns in straight river
reaches in order to generate the bedload sweep phenomenon associated with the
natural helicoidal flow in river bends.
Guide vanes are used to direct the flow either close to the bed or close to the
water surface. Bottom guide vanes (Fig. 8.9a) are used to deflect the water near the
bed, with its associated bedload or highly concentrated suspended load, away from
the intake.
Surface guide vanes (Fig. 8.9b) direct the relatively clear surface water towards
the intake, which induces a counter flow near the bed away from the intake. Surface
vanes are generally supported by a raft arrangement from which the vanes project
downwards into the water far enough to influence the flow direction of surface
water.
Model tests conducted by USBR have shown that the two configurations are
about equally effective in reducing bed load intake. Surface vanes have the draw-
back that they tend to trap and accumulate floating debris, causing maintenance
problems. For a particular model study reported by Vanoni (1975) the use of bottom
vanes reduced the ratio of sediment concentration in the diverted flow to that in the
river from 2.38 to 0.10. Model tests are recommended for specific structures.
Stream Inlets: are deep inlet structures built in the bed of the stream. They tend to
trap a lot of sediment and require much maintenance unless some sort of sluiceway
is incorporated. They are, however, much less expensive than diversion barrages
with other types of sediment diverters and are, therefore, used quite commonly for
small diversions. A typical layout incorporating a sluiceway is shown in Fig. 8.10.
A steep channel gradient makes sluicing relatively easy as a short high head sluice
pipe can be used.
8.2 River Intake Design for Sediment Control 297

to canal to canal

near surface flow


near bed flow

(a) Bottom guide vanes (b) Surface guide vanes

Fig. 8.9 Guide vanes for sediment diversion

Stream inlets are usually used for relatively small diversions, which are often on
mountain streams and relatively inaccessible. They should, therefore, be designed
to ensure minimal blocking by debris. The trashrack, for example, should always
have a slope in the downstream direction about 1:12 greater than the stream
gradient.
Some more sophisticated stream inlet designs have been used for automatic
intermittent sluicing, such as that developed by Raynaud (1951). In this system,
water and sediment pass through trashracks into a triangular desilting basin where
the sediment accumulates and clear water flows over a weir crest into a collecting
channel. The desilting basin is sluiced periodically under the control of an auto-
matic gate that operates by buoyancy created when sediment has accumulated in the
desilting basin.
Tunnel Type Sediment Diverters: exploit the vertical distribution of transported
sediment by separating the high-concentration flow near the bed and directing it
through a number of tunnels back to the river downstream (Fig. 8.11). The rela-
tively clear water above the tunnels is directed to the canal. The tunnels are parallel
to the axis of the canal intake control structure. The tunnel closest to the canal
intake begins some distance upstream of the intake and the others decrease in length
298 8 Intake Structures

A
canal

sluice pipe
sluiceway
C C

A
(a) Plan

canal

sluiceway
sluice pipe
(b) Section A-A

trashrack

(c) Section C-C

Fig. 8.10 Stream inlet structure configuration

away from the intake. Tunnel diverters have proved to be very efficient for large
and small diversions.
It must be ensured that the flow separation at the tunnel entrances takes place
smoothly to reduce energy losses and to prevent induced turbulence from sus-
pending material moving as bedload into the tunnels. The tunnel dimensions must
be designed to ensure that the sediment can be easily transported and clogging does
not occur. This requires a velocity of about 3 m/s and it has been shown that the
discharge should be at least 15–20% of the canal diversion for efficient operation
(Vanoni 1975). It is extremely important to ensure that the discharge point of the
lower channel is kept clear at all times, even if this requires a channel separate from
the stream. A recommended design procedure is presented by Kothyari et al.
(1994).
8.2 River Intake Design for Sediment Control 299

canal

sluiceway
tunnels

river flow
river flow A

barrage
(a) Plan

tunnels
canal

(b) Section A-A

Fig. 8.11 Tunnel type sediment diverter

Sediment Ejectors
Sediment ejectors are designed to remove from a canal any sediment that was not
diverted at the headgates. Even if a diverter is incorporated in an intake some
sediment will enter the aqueduct, and some material transported in suspension in the
river may settle out and become bedload in the canal.
A number of different types of ejector devices have been developed for different
situations. The most common are described below (further details are given by
Vanoni (1975) and Raudkivi (1993)).
Tunnel Type Ejectors: operate on the same principle as tunnel type diverters, i.e.
by separating the relatively clear water nearer the surface from the heavily
sediment-laden water near the bed. An example configuration is shown in Fig. 8.12.
A number of tunnels extending across the width of the channel are set below the
floor level with their openings directed upstream. The tunnels converge and curve
to one side, passing through the side of the channel and a gated control structure to
an outfall. The tunnels may be divided into smaller tunnels before the convergence
in order to improve the flow pattern. At the opening, the tunnels usually have a
height equal to about 20–25% of the canal flow depth. The roof usually extends
300 8 Intake Structures

A A

canal

sediment
outfall
(a) Plan

clear water
sediment

(b) Section A-A

Fig. 8.12 Tunnel type sediment ejector

upstream of the diversion wall to prevent resuspension of material by turbulence


associated with the transition.
The tunnels should be set with a differential head large enough to induce a
velocity sufficient to move the sediment through; a velocity of 2.5–3.0 m/s is
generally sufficient for sand-sized sediments. Efficiency increases with the differ-
ential head but so does the discharged flow rate which must be allowed for in the
intake and aqueduct design. Normally the aqueduct discharge must be increased by
20–25% to operate an ejector of this type.
These ejectors can be combined with canal wasteway structures that regulate the
flow rate in the aqueduct.
Vortex Tube Ejectors: extract sediment moving along the canal bed by means of
an open tube set in the floor at an angle of about 45° to the flow direction, as shown
in Fig. 8.13.
Water flowing over the opening sets up a spiralling motion in the tube which
helps to move the bedload along the tube to an outlet. Tests have been conducted on
these structures by various researchers. Parshall (1950), for example, used a tube of
about 100 mm and found that velocity of about 0.75 m/s over the lip induced
rotation of about 200 rpm in the tube, which was sufficient to move gravel the size
of hens’ eggs.
8.2 River Intake Design for Sediment Control 301

Fig. 8.13 Vortex tube type


sediment ejector A

sluiceway
(a) Plan

(b) Section A-A

Based on theoretical analyses, confirmed by comparison with model and field


data, Atkinson (1994a, b) proposed methods for predicting the trapping efficiency
of a vortex tube extractor and for designing a tube. His results suggested that the
tube should preferably be set perpendicular to the flow direction, and the ratio of slit
width to tube diameter (he used a circular tube) should be in the range of 0.15–0.30.
Settling Basins
The diverter and ejector devices already discussed are intended to remove mainly
coarse sediment material which is transported as bedload or as suspended load close
to the bed. Finer material which is transported in suspension in the river and the
aqueduct can be removed by using a settling basin downstream of the headworks.
Underlying concepts are well described by Raudkivi (1993) and guidelines for the
design are presented by Avery (1989).
A settling basin is an expanded section of the canal, or some other feature, in
which the velocity is reduced to permit settling of the suspended particles.
Clear water only can then be directed into the aqueduct and the sediment can be
removed from the basin by sluicing or mechanical means. The basin comprises an
inlet zone, a settling zone and an outlet zone (Fig. 8.14). The inlet zone is designed
to distribute the flow and suspended sediment uniformly into the settling zone. The
expansion should, therefore, be gradual to avoid flow separation, and the inlet may
incorporate a submerged weir, guide vanes or baffle walls. The outlet zone provides
a transition back to the canal and may include measures for controlling the water
level in the settling zone such as a submerged weir or gate. As for the inlet zone, the
transition should ensure evenly distributed flow, but its length may be shorter for
the contracting rather than expanding flow.
The settling zone must be designed to allow most of the suspended sediment to
settle out and to prevent settled material from being remobilized. These objectives
302 8 Intake Structures

sluiceway

inlet zone settling zone outlet zone

(a) Plan

settled sediment
(b) Long Section

Fig. 8.14 Settling basin configuration

are achieved by appropriate sizing of the length and width of the basin, as governed
by the flow velocity and sediment settling velocity.
Settling can be described reasonably well analytically and many relationships
and mathematical models have been proposed for predicting the removal efficiency
of settling basins. For the ideal situation of particles settling in the absence of
turbulence and re-entrainment, the rate of sediment entering the basin is the con-
centration multiplied by the discharge, Q, and the rate of settling is the concen-
tration (C) multiplied by the plan area of the basin, A, and the settling velocity,
w. The settling efficiency (also referred to as the removal ratio or trap efficiency) is,
therefore, given by

CwA wL
E¼ ¼ ð8:6Þ
CQ q

in which L is the length of the settling basin and q is the unit width discharge. For
real situations, Vetter (1940) proposed the relationship

W ¼ Wo ewL=q ð8:7Þ

in which W is the weight of sediment leaving the basin and Wo is the weight of
sediment entering the basin. The settling efficiency is then given by
8.2 River Intake Design for Sediment Control 303

 
Wo  W W
E¼ ¼ 1 ¼ 1  ewL=q ð8:8Þ
Wo Wo

Equation (8.8) assumes a high degree of turbulence, with less turbulence leading
to greater settling efficiency. Camp (1946) produced a graph showing the increase
in efficiency with decreasing turbulence (Fig. 8.15). The effect of turbulence is
accounted for by the turbulent mixing coefficient, which can be expressed in terms
of the shear velocity, u*. The efficiency, therefore, depends on both the particle
settling velocity and the turbulence and is related to the parameter w/u*. The shear
velocity can be evaluated through a resistance equation, such as Manning’s, giving
the form given in Fig. 8.15 with the hydraulic radius represented by the flow depth,
D. The settling efficiencies in Fig. 8.15 agree closely with Eq. (8.8) for high tur-
bulence (w/u* = 0.01) and increase with decreasing turbulence to agree closely with
Eq. (8.6) for low turbulence (w/u* approaching 0.1 for low wA/Q and 10.0 for
higher values).

Increasing turbulence

100
wA/Q = 2.0 1.0
1.5 0.80
80
0.70
1.2
0.60
60
E (%)

0.50

0.40
40
0.30
0.20
20
0.10

0
0.01 0.10 1.0 10.0

w w D1 6
u* nV g

Fig. 8.15 Camp’s (1946) solution for settling efficiency


304 8 Intake Structures

Results of laboratory experiments by Garde et al. (1990) contradicted predictions


using the methods of Camp (1946) and Vetter (1940), especially for fine sediments.
They proposed a new equation,
 
E ¼ Eo 1  ekL=D ð8:9Þ

in which Eo is the limiting efficiency, i.e. the maximum that can be achieved for a
given sediment size and flow condition. Eo and k were both found to depend on
w/u*. Their experimental values were presented graphically and can be approxi-
mated as
 2:17
w
k ¼ 0:042 ð8:10Þ
u

and
 
w
Eo ¼ 0:434 þ 0:023 ð8:11Þ
u

Eo has a maximum value of 1.00 and Garde et al. (1990) recommend the
maximum value for k to be that corresponding to w/u* = 2.20, i.e. kmax = 0.23.
It was pointed out by Schrimpf (1991) that Camp’s (1946) graph is only valid for
a vertically uniform concentration at the stilling basin entrance, which is not always
a reasonable assumption and could account for the discrepancy found by Garde
et al. (1990).
Raudkivi (1993) describes further refinements for estimating settling efficiency.
He also introduces the more rigorous computational solutions of the sediment
diffusion equations, which are recommended for final design purposes.
Example 8.2
A canal conveying 8.0 m3/s leads into a 16.0 m wide, rectangular section settling basin
with Manning’s n = 0.013. The flow depth in the settling basin is maintained at
approximately 2.0 m. Determine the length of settling basin required to settle 75% of the
inflowing suspended sediment if its representative size is 0.10 mm with a settling
velocity of 7.82 mm/s.
Solution
The removal ratio equations include the basin length as a variable and can be rearranged
to give the length required to achieve a specified removal ratio.
Assuming quiescent flow without turbulence, the removal ratio is given by Eq. (8.6), i.e.

wA wL
E¼ ¼
Q q

Therefore
8.2 River Intake Design for Sediment Control 305

Eq

w

with

Q 8:0
q¼ ¼ ¼ 0:50 m3 =s/m
W 16:0
So

0:75  0:50
L¼ ¼ 48 m
0:00782
Assuming a high degree of turbulence, the removal ratio is given by Eq. (8.8), i.e.

E ¼ 1  ewL=q

Therefore

q
L¼ lnð1  E Þ
w
0:50
¼ lnð1  0:75Þ ¼ 89 m
0:00782
Assuming the degree of turbulence to be determined by the flow conditions, the removal
ratio is given by Fig. 8.15. The ordinate is E = 75% and the abscissa is given by

wD1=6
pffiffiffi
nV g

with
D ¼ 2:0 m
n ¼ 0:013
Q 8:0
V¼ ¼ ¼ 0:25 m/s
A 16:0  2:0
So

wD1=6 0:00782  2:01=6


pffiffiffi ¼ pffiffiffiffiffiffiffi ¼ 0:86
nV g 0:013  0:25  9:8

From Fig. 8.15

wA wL
¼ ¼ 0:80
Q q

Therefore
306 8 Intake Structures

q 0:50
L ¼ 0:80 ¼ 0:80 ¼ 51 m
w 0:00782
Figure 8.15 allows for intermediate roughness and gives a result between the extremes
given by Eqs. (8.6) and (8.8), which appears to be realistic. The position of w/u* in
Fig. 8.15 indicates only a small departure from the quiescent condition, consistent with
the solution being closer to that given by Eq. (8.6) than by Eq. (8.8).
Equations (8.9)–(8.11) also account for varying degrees of turbulence, i.e.
 
E ¼ Eo 1  ekL=D

Therefore,
 
D E
L¼ ln 1 
k Eo

with
 2:17
w
k ¼ 0:042
u

and
pffiffiffi
nV g
u ¼
D1=6 pffiffiffiffiffiffiffi
0:013  0:25  9:8
¼ ¼ 0:0091 m/s
2:01=6
So
 
0:00782 2:17
k ¼ 0:042 ¼ 0:030
0:0091

and
 
w
Eo ¼ 0:434 þ 0:023
u
 
0:00782
¼ 0:434 þ 0:023 ¼ 0:396
0:0091

Therefore,
 
2:0 0:75
L¼ ln 1 
0:030 0:396

for which there is no solution. The value of Eo = 0.396 gives the maximum possible
removal rate, so the required removal rate of 0.75 cannot be achieved with any basin
length. Using the length determined from Fig. 8.15, the removal rate according to
Eq. (8.9) is
8.2 River Intake Design for Sediment Control 307

 
E ¼ Eo 1  ekL=D
 
¼ 0:396 1  e0:03051=2:0 ¼ 0:212

which is considerably less than the required. This method suggests that the basin should
be widened to reduce V and hence u*.

The dimensions of the stilling basin must also ensure flow conditions that are not
competent to mobilize the settled sediment. The Shields criterion for scouring (see
Chap. 9) can be used for assessing the likelihood of remobilization.
The settled sediment can be removed by sluicing if there is sufficient river flow
for the additional sluicing water to be diverted. The discharge, velocity and channel
conditions must be such that sediment is transported downstream and does not
deposit and clog the river channel near the intake. Intermittent sluicing can be done
by operation of the gates controlling the sluiceway. Designs have also been
developed for continuous sluicing without temporary sediment accumulation
(Raudkivi 1993). If there is insufficient flow available for sluicing the accumulated
sediment must be removed from settling basins by mechanical means, such as
dredging or dragling, which can be very expensive.

8.3 Pump Sumps and Intakes

In many cases water cannot be extracted directly from a river and provision for
pumping must be made. The intake must then incorporate a sump to provide
sufficient volume and depth for the pumps to operate. The efficiency of a pump
intake structure is governed to a large degree by the geometry of the configuration.
As for other intake structures, design problems cannot readily be solved analytically
and most design procedures and recommendations have been developed from
model experiments and observations of existing structures. In practice, particular
situations often require unusual or complicated solutions that necessarily depart
from the general recommendations. The design solution for these problems should
be obtained—or at least confirmed—by model studies.

8.3.1 Desirable Flow Conditions

Poor flow conditions resulting from careless design can lead to unanticipated
operating restrictions and extra costs through delays in commissioning, increased
maintenance, structural alterations and retrospective model tests. Useful guidelines
for design have been produced by Hydraulic Institute (1998) and Prosser (1977),
from where much of the following content is derived and adapted.
For effective and efficient pumping the flow to the pump should be single-phase
(i.e. containing no entrained air), uniform (i.e. the flow velocity should be constant
in magnitude and direction across the approach section) and steady (i.e. not fluc-
tuating with time in magnitude and direction).
308 8 Intake Structures

Well designed configurations are usually able to achieve the single-phase ideal
but some minor departures from uniform and steady conditions are to be expected.
There will always be some nonuniformity due to boundary layer formation at the
walls and a certain amount of unsteadiness will be produced by small-scale tur-
bulence. Significant departures from these conditions, however, can lead to poor
performances or increased costs.
Departures from the single-phase requirement can have serious effects on the
pump, apart from causing possible undesirable effects of air in the water supply. High
air content can cause the pump to deprime and cease to deliver. Lesser amounts of air
can cause significant reductions in discharge and efficiency, the severity of which will
depend on the type of pump. For example, for a centrifugal pump 3% of free air has
been found to decrease efficiency by 15% and axial pumps are even more sensitive.
Further, entrained air will cause uneven loading on the pump impeller which can
result in vibration, rough running and damage to bearings.
Approach conditions should, therefore, be designed to ensure that no free air
enters the intake. There are several ways, related to intake or sump geometry, in
which air can be entrained.
If the water level in the sump is very close to the top of the intake air may be
drawn through the intake continuously or by ‘gulping’. This tendency is increased
with high intake velocities, particularly where approach velocities are relatively low
and the draw-down of the water surface near the intake is accentuated. A minimum
submergence of the intake is, therefore, necessary.
Air can also be entrained by a falling jet of water, which is common where water
enters the sump over a weir or through a culvert at a higher level than the water in
the sump. Air bubbles may be carried over the length of the sump and drawn into
the intake. If it is not possible to avoid high-level entry into the sump then the sump
should be made long enough to allow the air bubbles to rise to the surface before
reaching the intake. A minimum sump length could be calculated from the approach
velocity and water depth in the sump and the rising velocity of air bubbles (about
0.2 m/s for 2–5 mm diameter bubbles).
Intense vortex action can lead to air entrainment (Fig. 8.16). A vortex may be
stable with a diffuse or solid air core or it may be unstable and unsteady with
intermittent air entrainment only. The formation of vortices is dependent on sub-
mergence and on intake and sump geometry. High intake velocities help to initiate
air entrainment by drawing bubbles off the bottom of the surface depression
associated with the vortex.
Departures from uniform flow, such as intense swirling can cause rapid changes
in local pressure on pump impellers, which can lead to cavitation and severe
vibration with associated bearing damage. Axial type pumps are generally most
susceptible to this type of damage. Sometimes larger scale, less intensive swirling
flow may be centered on the pump axis. Depending on the direction of swirl relative
to the movement of the impeller, this could be beneficial or detrimental to pump
performance.
Intense swirling flow is associated with vortex action. In addition to the surface
vortices which also cause air entrainment, there may be submerged vortices that
8.3 Pump Sumps and Intakes 309

Fig. 8.16 A solid air-entraining vortex in a model pump sump

originate at the floor or walls of the sump rather than the water surface. Larger scale
swirl is generally caused by general circulation in the sump which is amplified as
the flow converges towards the intake. The general circulation may be initiated by a
distorted velocity profile in the approach flow and the sump geometry. This type of
nonuniformity may be steady but can help to initiate surface or submerged vortices.
Departures from a steady flow can cause variations with time of pump blade
loading, resulting in vibrations and bearing wear. The effects are worst if the pump
is very close to the intake: some of the unsteadiness becomes damped out if there is
a length of conduit, bends or changes of cross section between the intake and pump.
In addition to the small-scale turbulence which can normally be expected,
large-scale turbulence where the eddy size is of the same order of magnitude as the
intake cross section sometimes occurs. The major causes of this are unsteady flow
patterns arising from obstacles in the sump or poor inlet conditions and vortex
shedding from pillars or other pumps.
Another common cause of unsteadiness is the presence of stagnant regions of
water above or behind the intake. Boundaries between stagnant regions and the flow
tend to be unstable and the changing position causes unsteadiness in the main flow.
The presence of stagnant regions also favours vortex formation and can, therefore,
also lead to nonuniformity and air entrainment.

8.3.2 Intake and Sump Design

General Layout
Four functional zones can be identified in a typical sump-intake arrangement
(Fig. 8.17). The first zone is the inlet to the pump station from the supply source.
The entry to the pump station may be by canal or pipeline or directly from a river or
lake, and often includes a control structure such as a weir. The second zone is an
310 8 Intake Structures

screen
penstocks
weir
A A

inlet approach sump

(a) Plan

motor

pump

wet well sump bellmouth

(b) Section A-A


Fig. 8.17 General sump-intake arrangement

approach section which may contain screens for removing solid matter and gates or
division walls for directing the flow to the appropriate sump. The sump itself
constitutes the third zone. It is generally rectangular in plan with a flat floor. The
purpose of the sump is to provide storage and to damp out distorted flow patterns
which may result from the configuration of the distribution zone. The sump is
separated from the final zone by the intake, a section where the water enters a
closed conduit, i.e. a division between the free surface and internal flow. The final
zone is the short section of conduit between the intake and the pump, which should
be kept as short and straight as possible.
This is a very general description and the actual configuration may vary con-
siderably to meet specific requirements. For example, the sump may be left out
completely when water is drawn directly from a reservoir.
The pump may be installed within the sump (wet well arrangement), or in a dry
well at the end of the sump (Fig. 8.18). The wet well has the advantage of the
simplicity of design and hence relatively low-cost. It has the disadvantage that
8.3 Pump Sumps and Intakes 311

Fig. 8.18 Wet well and dry motor


well arrangements

S
S pump

(a) Wet well (b) Dry well

maintenance requires drainage of the sump or removal of the pump. The wet well
arrangement is ideal for conditions where intermittent pumping only is required,
such as storm-water pumping, and the sump will be dry for most of the time.
If reliability is important a dry well arrangement is preferable, making the pump
accessible for maintenance at all times. The intake for a dry well configuration may
be horizontal or with a turned down bellmouth in the sump, depending on minimum
water levels required in the sump. A turned down bellmouth allows a lower water
surface and is less susceptible to vortex formation.
Flow towards the intake should be uniform across the width of the sump. This
ensures a uniform approach and lessens the likelihood of swirl developing near the
intake, which could lead to the occurrence of vortices with associated air entrain-
ment and nonuniformity problems. If more than one pump is installed the approach
flow conditions are considerably improved by placing dividing walls between them,
especially if the pumps will not necessarily operate together (Fig. 8.19). Prosser
(1977) recommends a sump width of twice the bellmouth diameter and a length of
ten times the diameter. Possible stagnant flow regions should be filled, as they can
be unstable and cause fluctuations and unsteadiness in the flow. The flow velocities
should be kept low, about 0.6 m/s into the pumping station and 0.3 m/s
approaching the bellmouth. Any obstructions, such as structural elements, should
be streamlined to avoid flow separation with associated nonuniformity and
unsteadiness near the intake. Excess kinetic energy associated with changes in level
down slopes, over weirs or through control structures should be dissipated well
away from the intake. If the approach section is sloped, Prosser (1977) recommends
that the slope be less than 10° and if it expands the flare angle should not exceed
20°. Hydraulic Institute (1998) presents examples of various configurations and
their associated flow patterns.
Bellmouth Intakes
The bellmouth intake is provided to prevent flow separation which would occur
with a sharp-edged inlet. The turbulence associated with separation is thus avoided
and uniform flow through the intake with a minimum head loss is ensured. For
intake in a vertical wall, the minimum bellmouth radius required is about 0.1 of the
pipe diameter; for a suspended pipe the radius would have to be larger to allow for
312 8 Intake Structures

10D

2D D
corner
fillets

0.5D

Fig. 8.19 Multiple pump arrangement

flow from behind the bellmouth. In practice the shape is normally specified by
pump manufacturers and usually forms a quarter ellipse in section. The bellmouth
diameter is usually 1.5–1.8 times the pipe diameter. Prosser (1977) recommends a
clearance below a vertical pipe bellmouth of 0.50 of its diameter. Hydraulic
Institute (1998) gives recommendations for bellmouth diameters and entry veloci-
ties for different pumping rates.
Sump Volume
The sump volume is an important design decision. In addition to ensuring satis-
factory approach flow conditions to the pump intake, the sump provides storage to
allow for constant speed pumps to switch on and off as the sump fills and empties.
Factors affecting this decision include the number and capacities of pumps, their
operating levels and minimum cycle times. The minimum cycle time is determined
by the heat generated in an electric motor during start-up, which imposes a
restriction on the frequency of starts. This is not significant if variable-speed pumps
are used but constitutes an important constraint on sump volume for fixed-speed
pumps.
For a fixed-speed pump, the cycle time (T) for filling and emptying the active
sump volume (V) is

V V
T¼ þ ð8:12Þ
Qi Qp  Qi

in which Qi is the inflow rate and Qp is the pumping rate. This can be rearranged to
give the active volume required for a specified cycle time, i.e.
8.3 Pump Sumps and Intakes 313

Qi  
V ¼T Qp  Qi ð8:13Þ
Qp

By differentiating Eq. (8.12) it can be shown that the minimum cycle time, or the
maximum frequency of pump starts, occurs if the inflow rate is half the pumping
rate. If this condition is applied, then (from Eq. (8.13))

TQp
V¼ ð8:14Þ
4

Hydraulic Institute (1998) presents examples of the application of these rela-


tionships to multiple fixed-speed pumps with sequential operation.

Prevention of Vortices
The formation of vortices is a major source of flow nonuniformity and air
entrainment. Vortices may occur completely below the surface, attached to the floor
or a side wall, and cause considerable swirl into the intake. Surface vortices also
produce swirl and, if sufficiently intense, can entrain large quantities of air which
can lead to major pump damage.
Subsurface vortices at a vertical bellmouth can be eliminated or at least reduced
by the careful setting of the clearance below the intake and/or the installation of
swirl-preventing devices (Fig. 8.20).
Surface vortices can be prevented by ensuring uniform approach flow, providing
sufficient submergence of the intake and/or providing vortex-suppressing devices.
Uniform flow can be ensured by adhering to the general layout recommendations
above.
A high intake submergence reduces the likelihood of surface vortex formation.
However, the submergence defines the lowest point of the pumping station and,
therefore, has a significant influence on civil costs. The minimum water level is
generally defined by external conditions, so the sump floor level is determined by
the submergence and the clearance below the bellmouth. The submergence should,
therefore, be kept as small as possible to minimize costs.

(a) Cone (b) Short splitter or cross (c) Long splitter

Fig. 8.20 Sub-surface vortex prevention devices


314 8 Intake Structures

The minimum submergence to ensure vortex-free conditions depends on the size


of the bellmouth and the intake velocity; generally, the minimum submergence
increases with increasing intake velocity. Prosser (1977) recommends a minimum
of 1.5 times the bellmouth diameter for turned down intakes and 1.0 times the
diameter from the top of the bellmouth for horizontal intakes. Hydraulic Institute
(1998) recommends Eq. (8.15) for specifying the submergence required for pre-
venting strong air-entraining vortices.

S
¼ 1:0 þ 2:3 FD ð8:15Þ
D

in which S is the submergence (see Fig. 8.18), D is the diameter of the inlet (the
bellmouth if used) and FD is a ‘Froude number’ based on the inlet velocity and
diameter, i.e.

V
FD ¼ pffiffiffiffiffiffi ð8:16Þ
gD

Werth and Frizzel (2009) carried out experiments with vertical intakes to
determine the submergence required to prevent vortices weaker than able to entrain
air, but strong enough to cause swirl (as indicated by a dye core) right to the intake.
For this more restrictive condition, they proposed the minimum submergence to be
given by

S 4 2=3
¼ 2:1 þ FD ð8:17Þ
D 3

These are general recommendations and any unusual configuration should be


examined carefully—there are cases where vortex action is worst at relatively high
submergence. Submergence on its own cannot guarantee vortex-free conditions;
strong air-entraining vortices can occur at high submergence if the approach
geometry allows swirl to develop.
If surface vortices are found to occur in existing structures and modifications for
improving approach flow conditions or increasing submergence are not possible or
ineffective, then vortex-suppressing devices can be installed. These take a variety of
forms, all intended to break circulation in the vicinity of the intake. For vertical,
upward-directed intakes a perforated screen wall (Fig. 8.21a) or a vertical splitter
(Fig. 8.21b) may be effective. For a vertical downward-directed intake (a very
vulnerable arrangement) anti-rotation baffles or guide vanes have been found to be
effective (Fig. 8.21c). For horizontal intakes, the most common vortex-suppressing
devices are horizontal plates, which may be fixed or floating and solid or perforated
(Fig. 8.21d). Amiri et al. (2011) investigated different plate types and positions and
found that most satisfactory vortex suppression was obtained with a rectangular,
perforated plate positioned at the top of the intake. The optimal dimensions were
1D from the intake face in the intake flow direction and 1.5D normal to this
8.3 Pump Sumps and Intakes 315

(a) Screen wall (b) Long splitter

A A
A-A
A A A-A

(c) Baffles (d) Horizontal plate


Fig. 8.21 Surface vortex prevention devices

direction, where D here is the pipe diameter (excluding bellmouth). The optimal
area of perforations was 50% of the plate area. They did not investigate the effect of
perforation size, but Chigura et al. (2016) found the effectiveness to increase with
decreasing perforation size.

8.3.3 Model Testing for Intakes

Most guidelines for deigning pump sumps and intakes are based on limited
information derived from prototype observations and model studies. Any proposed
departures from the guidelines should, therefore, be tested by a model study of the
particular configuration. The significance of circulation and air entrainment in the
functioning of sumps and intakes presents some special challenges for their
physical modelling.
Free surface flow occurs up to the intake and model similarity can be ensured by
maintaining equality of Froude numbers in the model and prototype, provided the
scale is sufficiently large to preclude viscosity and surface tension effects. Hydraulic
Institute (1998) recommends that to safely exclude these effects the Reynolds
number (Re = VD/m, where m is the kinematic viscosity) for the model intake should
be above about 6  104 and the Weber number (We = V2D/(r/q), where r is the
316 8 Intake Structures

interface surface tension and q is the fluid density) should be above 240. It is also
recommended that to observe flow patterns reliably and to obtain sufficiently
accurate measurements, the approach bay should be at least 300 mm wide, the flow
should be at least 150 mm deep and the pump throat should be at least 80 mm in
diameter. Anwar (1968) recommends that the scale should not be less than 1:20 for
modelling vortices. Prosser (1977) suggests a minimum of 1:25, but down to 1:50
for large intakes from reservoirs.
Avoidance of surface and subsurface vortices is a crucial design objective, but
there is uncertainty regarding the validity of Froude similarity for vortex formation.
It has been observed that similarity of air-entraining vortices occurs at model
velocities greater than indicated by Froude similarity. Denny (1956) found that
similarity appeared to occur at a velocity scale equal to the linear scale rather than
the square root of the linear scale as implied by the Froude law. He suggests that the
velocity required to form a surface dimple is scaled according to the Froude number
but the local velocity required to drag air off the bottom of the dimple is unaffected
by scale. There are, therefore, good arguments for using larger velocities than
required for Froude similarity when checking for vortices, even though this might
cause the water profile to be incorrectly represented. Prosser (1977) suggests
operating the model at Froude velocities and if these tests show little or no vortex
action, increasing the flow velocity to 2 or 3 times the Froude scale velocity to get
an idea of the margin of safety with respect to air-entraining vortices and to observe
any effects which may have been missed. Hydraulic Institute (1998) recommends
that both free surface and subsurface vortices should be less severe than able to
maintain coherent dye cores for more than 10% of the time.
Air entrainment by gulping or a falling water jet will also not be accurately
represented by the model. More air will be entrained in the prototype than as
indicated by the model and air bubbles will take longer to reach the surface.
Because surface vortex formation is influenced by approach flow patterns the
whole of the sump-intake structure including the approach and inlet zones should
be modelled. If the structure is divided into similar sections it is possible to model
one section alone only if no flow between the sections could occur in the prototype.
A plane of symmetry cannot be replaced by a solid boundary in the model unless
there is actually a corresponding physical separation of flow in the prototype
structure. If the intake is to draw water from a large body of water such as a
reservoir, the model must include a sufficiently large area of the surface to enable
full flow patterns to become established. It is not necessary to include the entire
structure on the downstream side of the intake section—a length of a few pipe
diameters is sufficient.
All details such as stop-log grooves and screens should be modelled. Modelling
of screens is important if the solidity is such that the flow patterns will be affected.
Screen material need not be modelled as long as the model screen has the same
solidity as the prototype.
8.3 Pump Sumps and Intakes 317

Problems
8:1 Water is to be extracted from a wide alluvial river directly into a canal. The
river has a slope of 0.00080 and a bed composed of sediment with a median
grain size of 0.12 mm (which has a settling velocity of 0.011 m/s). When the
flow depth in the river is 1.50 m the suspended sediment concentration is
0.050 at a height of 0.10 m from the bed. At what height above the river bed
should the bottom of the offtake canal be set to ensure that the average sus-
pended sediment concentration of the diverted water is not more than 0.020?

References

Amiri, S. M., Zarrati, A. R., Roshan, R., & Sarkardeh, H. (2011). Surface vortex prevention at
power intakes by horizontal plates. Water Management, 164(WM4), 193–200.
Anwar, H. O. (1968, October). Prevention of vortices at intakes. Water Power, 393–401.
Atkinson, E. (1994a). Vortex-tube sediment extractors. I: Trapping efficiency. Journal of
Hydraulic Engineering, 120(10), 1110–1125.
Atkinson, E (1994b) Vortex-tube sediment extractors. II: Design. Journal of Hydraulic
Engineering, 120(10), 1125–1138.
Avery, P. (Ed.). (1989). Sediment control at intakes—A design guide. Bedford, England: BHRA,
The Fluid Engineering Centre.
Camp, T. R. (1946). Sedimentation and the design of settling tanks. Transactions, ASCE, 111,
895–958.
Chigura, T., Mashika, B. K., & Mpofu, S. T. (2016). The use of perforated horizontal plates as
vortex suppressors at horizontal intakes. Final Year Investigational Project, School of Civil &
Environmental Engineering, University of the Witwatersrand, Johannesburg, South Africa.
Denny, D. F. (1956). An experimental study of air-entrainment vortices in pump sumps. In
Proceedings of the Institution of Mechanical Engineers (170 pp.).
Garde, R. J., Ranga Raju, K. G., & Sujudi, A. W. R. (1990). Design of settling basins. Journal of
Hydraulic Research, 28(1), 81–91.
Hernandez, N. M. (1969). Irrigation structures. In C. V. Davis & K. E. Sorensen (Eds.), Handbook
of applied hydraulics (3rd ed.). McGraw-Hill.
Hydraulic Institute. (1998). American national standard for pump intake design. Report ANS/HI
9.8-1998, Parsippany, New Jersey, USA.
Kothyari, U. C., Pande, P. K., & Gahlot, A. K. (1994). Design of tunnel-type sediment excluders.
Journal of Irrigation and Drainage Engineering, 120(1), 36–47.
Parshall, R. L. (1950). Experiments in cooperation with Colorado Agricultural Experiment
Station. Fort Collins, Colorado.
Prosser, M. J. (1977). The hydraulic design of pump sumps and intakes. British Hydromechanics
Research Association and Construction Industry Research and Information Association.
Raudkivi, A. J. (1993). Sedimentation: Exclusion and removal of sediment from diverted water.
Hydraulic Structures Design Manual 6. Rotterdam: International Association for Hydraulic
Research, A A Balkema.
Raynaud, A. (1951). Water intakes on mountain streams, example of application to the Torrent Du
Longon. International Association for Hydraulic Research, Fourth Meeting, Bombay, India
(pp. 1–9).
Schrimpf, W. (1991). Discussion of “design of settling basins”. Journal of Hydraulic Research,
29(1), 137–143.
318 8 Intake Structures

Vanoni, V. A. (Ed.). (1975). Sedimentation engineering. Prepared by the ASCE Task Committee
for the Preparation of the Manual on Sedimentation of the Sedimentation Committee of the
Hydraulics Division. New York: American Society of Civil Engineers.
Vetter, C. P. (1940). Technical aspects of the silt problem on the Colorado River. Civil
Engineering, 10(11), 698–701.
Werth, D., & Frizzell, C. (2009). Minimum pump submergence to prevent surface vortex
formation. Journal of Hydraulic Research, 47(1), 142–144.

Further Reading

Anwar, H. O. (1966). Formation of a weak vortex. Journal of Hydraulic Research, 4(1), 1.


Anwar, H. O. (1967, November). Flow in a free vortex. Water Power, 455.
Anwar, H. O., Weller, J. A., & Amphlett, M. (1978). Similarity of free-vortex at horizontal intake.
Journal of Hydraulic Research, 16(2), 95–105.
Blaisdell, F. W. (1960). Hood inlet for closed conduit spillways. Journal of the Hydraulics
Division, ASCE, 86(HY5), 7.
Camp, T. R., & Lawler, J. C. (1969). Water supplies. In C. V. Davis & K. E. Sorensen (Eds.),
Handbook of applied hydraulics (3rd ed.). McGraw-Hill.
Dagget, L. L., & Keulegan, G. H. (1974). Similitude in free surface vortex formations. Journal of
the Hydraulics Division, ASCE, 100(HY11), 1565–1582.
Goldschmidt. (1974). Wet-well volumes for multipump systems. Journal of the Irrigation and
Drainage Division, ASCE, 100(IR3), 371–385.
Goldschmidt. (1978). Mixing fixed-speed pumps to variable flows. Journal of the Water Pollution
Control Federation, 50(7), 1733–1741.
Gordon, J. L. (1970). Vortices at intakes. Water Power, 22, 137–138.
Graf, W. H. (1971). Hydraulics of sediment transport. McGraw Hill.
Hattersley, R. T. (1965). Hydraulic design of pump intakes. Journal of the Hydraulics Division,
ASCE, 91(HY2), 223–249.
Hecker, G. E. (1981). Model-prototype comparisons of free surface vortices. Journal of the
Hydraulics Division, ASCE, 107(HY10), 1243–1259.
Henderson, F. M. (1966). Open channel flow. Macmillan.
Razvan, E. (1989). River intakes and diversion dams. In Developments in civil engineering (Vol.
25). Elsevier Science Publishers.
Vanoni, V. A. (Chmn). (1972). Chapter V: Sediment control methods: C control of sediment in
canals. Journal of the Hydraulics Division, ASCE, 98 (HY9), 1647–1689.
Chapter 9
Scour and Scour Protection

9.1 Introduction

Many hydraulic structures present the potential for scour at the interface between
flowing water and an erodible boundary. The boundary in question may be an
integral part of the structure, such as an embankment or the bed and banks of an
unlined canal, or it may be adjacent to it, such as the riverbed downstream from a
culvert or dam spillway. Scour refers to the removal or redistribution of the sedi-
ment material forming the boundary. This can be deleterious, or even catastrophic
to the functioning of the structure, and must therefore be considered in its design.
Scour may be local or general. Scour occurs at any location where the shear
stress exerted by the flow on the bed is sufficient to move the bed particles. Local
scour is often associated with the rapidly varied flow patterns induced by structures,
and frequently occurs around bridge piers and downstream of culverts, spillways
and other river structures. General scour, although a manifestation of local scour
phenomena, refers to a net bed degradation over some distance, and is associated
with an excess of sediment transport capacity over the sediment supply rate. Such
situations are common downstream of impoundments, which trap sediment and
hence reduce the supply further downstream. Similar problems occur in urban
rivers, where increased flood flows increase the sediment transport capacity.
This chapter addresses the conditions under which cohesionless sediment par-
ticles (including all sizes from fine sand to large rocks) can be moved by the flow. It
presents criteria for predicting whether or not scour can be expected, and for
designing loose protective linings (riprap) capable of resisting scour. In many
practical applications, it is required that no movement of the bed material is
acceptable, while in others (such as around bridge piers) local scour occurs but
stabilizes at some depth, with or without sediment movement, and the design
problem is to establish that the stabilized scour depth is acceptable. Some designs
for addressing general scour problems require ensuring a balance between sediment

© Springer Nature Switzerland AG 2020 319


C S James, Hydraulic Structures,
https://doi.org/10.1007/978-3-030-34086-5_9
320 9 Scour and Scour Protection

supply and transport capacity, rather than ensuring no movement of material. The
information in this chapter cannot be used directly to analyse such problems.
Much work has been done on establishing scour conditions and predicting scour
magnitudes, both theoretically and empirically, for single particles and bulk sedi-
ments. At present, no completely theoretical description of scour is possible, but a
theoretical analysis does help to gain an understanding of the phenomenon.

9.2 Theoretical Analysis

The critical flow condition at which scour can just take place can be established by
analysing the stability of a single particle in the bed (e.g. James 1990).
The stability of a particle is determined by the forces acting on it (Fig. 9.1).
These are its submerged weight (W), the reaction forces (R) and friction forces (Ff)
between it and adjacent particles, and the hydrodynamic lift (FL) and drag (FD)
forces imposed by the flow.
At the condition of incipient motion, the reaction and friction forces through all
but one or two contact points will reduce to zero, and the particle will tend to move
by sliding or pivoting over the remaining contact point or points. Pivoting is
probably more common in nature, but both mechanisms do occur and at about the
same hydraulic conditions. Either mechanism can be assumed and will lead to
similar formulations. Pivoting is the more common assumption and is followed
here.
The relationship between the relevant forces at incipient motion can be described
by considering the equilibrium of their moments about the pivot axis, i.e.

W cos a a sinðu þ aÞ ¼ W sin a a cosðu þ aÞ þ FD b þ FL c ð9:1Þ

FL

F
Ff

φ FD
R Ff pivot axis
Ff
R
R
α
W

Fig. 9.1 Forces acting on a particle


9.2 Theoretical Analysis 321

In Eq. (9.1), a, b and c are the distances from the pivot axis to the particle
centroid, the line of action of the drag force and the line of action of the lift force,
respectively, u is the pivot angle and a is the bed slope.
The variables in Eq. (9.1) can be quantified in terms of sediment, geometric and
flow parameters, and the equation rearranged to obtain a scour criterion with the
form
sc
¼ K tanðu  aÞ ð9:2Þ
cDðSs  1Þ

in which sc is the boundary shear stress at incipient motion, c is the specific weight
of water, D is the sediment particle size and Ss is the specific gravity of the
sediment. The coefficient K accounts for a variety of geometric and hydraulic
effects, which are more explicitly accounted for in the more complete analysis of
James (1990). Equation (9.2) is sufficiently rigorous for many engineering appli-
cations, but requires empirical determination of K.

9.3 Empirical Approach

It is not possible at present to develop a complete theoretical solution to the in-


cipient motion problem—some empirical content is necessary. It is possible to
address the problem completely empirically and many results have been presented
which are useful for engineering applications.
An empirical study begins with an identification of the variables which affect the
phenomenon. These are the boundary shear stress (so), the fluid density (q), the
fluid viscosity (l), the particle size (D) and the submerged specific weight of the
grain (cs − c) where cs is the sediment specific weight. The functional relationship
between these variables can be expressed as

f ðso ; q; l; D; ðcs  cÞÞ ¼ 0 ð9:3Þ

Using dimensional analysis, two dimensionless groupings can be formed from


the identified variables. Hence
 
so u D
f ; ¼0 ð9:4Þ
ðcs  cÞD m

The first term, often referred to as the Shields parameter (s*), will be recognized
as the same dimensionless shear stress which was developed theoretically. The
second is the shear Reynolds number, Re*, in which u* = (so/q)1/2 and m = l/q.
Both of these parameters have physical interpretations. The shear Reynolds number
represents the ratio of the particle size to the thickness of the viscous sublayer. It
therefore represents the degree to which grains project through this layer and are
322 9 Scour and Scour Protection

subjected to turbulent flow. The dimensionless shear stress can be interpreted as the
ratio of the shear stress (or average total drag force acting on a unit area of bed) to
the submerged weight, per unit area of bed, of a single layer of particles. It therefore
represents the ratio of disturbing to stabilizing forces on a bed of particles. The
functional relationship between these parameters requires experimental
investigation.
The classic work in this field was by Shields (1936) who produced the first
version of what has come to be known as the Shields diagram (Fig. 9.2). It is a
graphical representation of the relationship implied by Eq. (9.4), at the condition of
incipient motion (i.e. so = sc), as determined from flume experiments.
The curve in Fig. 9.2 represents the threshold of movement. If the combination
of sediment and flow characteristics represented by the two variables plots below
the curve then no movement of sediment is expected. If it plots above the curve then
movement, or erosion, is expected to occur. Figure 9.3 shows a similar curve
compiled by Yalin and Karahan (1979) including additional data; this shows a more
gradual slope for small values of Re* and a lower constant value of sc* at high
values. (A similar curve was presented by Miller et al. (1977)).
The following three zones can be identified on the Shields diagram:
Hydraulically smooth turbulent zone: Re* < *2. The particle diameter is smaller
than the thickness of the viscous sublayer and the particle is therefore enclosed by
laminar flow and particles move mainly under the influence of viscous forces. The
critical entrainment shear stress depends strongly on Re*.
Transition zone: 2  Re*  *400 (or *70 on more modern diagrams). The
particle size is about the same as the thickness of the viscous sublayer. The particle
is partly enclosed by the viscous sublayer and subjected to periodic exposure to

τo D
τ* = 0.1( S s − 1) g D
γ D ( S s − 1) ν
2 4 6 10 20 40 60 100 200 400 600 1000
0.1

0.01
0.2 1.0 10 100 1000
u D
Re* = *
ν

Fig. 9.2 The Shields diagram (adapted from Vanoni 1975)


9.3 Empirical Approach 323

τo
τ* =
γ D ( S s − 1) laminar
turbulent
0.1

0.01
0.1 1.0 10 100 1000

u D
Re* = *
ν

Fig. 9.3 The Shields diagram (adapted from Yalin and Karahan 1979)

turbulent flow. The dimensionless critical shear stress reaches its minimum of about
0.030 in this zone, at a value of Re* of about 10.
Fully developed turbulent zone: Re* > *400 (or *70 on more modern dia-
grams). In this zone, particles protrude through the viscous sublayer and are fully
exposed to turbulence bursts, which cause instantaneous shear stresses significantly
greater than the time-averaged values. The dimensionless critical shear stress is
therefore considerably less than unity, about 0.056 on old diagrams and 0.047 on
more modern versions (e.g. Miller et al. 1977; Yalin and Karahan 1979), and
independent of Re*.
There are various shortcomings of Shields’s results and many extensions and
modifications have been made since. Some of these are described below.
Because u* appears in both variables, a solution is not explicit in the hydraulically
smooth and transition zones. An iterative solution is therefore required. In the version
presented by Vanoni (1975) (Fig. 9.2), an auxiliary scale is included. If the value of
the parameter
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 
D cs
0:1  1 gD
m c

is located on the scale and projected along the direction of the parallel lines, the
intercept with the threshold curve indicates the required values of Re* and s*.
Yalin (1977) overcomes this problem by replacing Re* by the parameter
324 9 Scour and Scour Protection

rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ðSs  1ÞgD3
m2

and presenting a revised form of the diagram.


The constancy of the Shields parameter at about 0.047 in the rough turbulent
zone (Re* > *70) enables a simplified relationship to be derived. If quartz density
(Ss = 2.65) and a water viscosity of 10−6 m2/s are assumed, it can be shown that
Re* = 70 corresponds to a particle size of about 2 mm. For D > 2 mm, s*c is
therefore constant at 0.047, i.e.
so
¼ 0:047
cDðSs  1Þ

and using the general shear relationship

s0 ¼ cRS

in which R is the hydraulic radius and S is the gradient, yields

D  13RS ð9:5Þ

as an erosion criterion. It must be remembered that S is the energy gradient, which


is equal to the bed gradient only if the flow is uniform. The energy gradient can be
calculated by a resistance equation, such as the Manning equation.
Shields had very little data for low values of Re* (i.e. fine sediments) and his
original diagram exaggerated the threshold for Re* < *1.0. The curve has been
refined to account for this by White (1970), Grass (1970), Mantz (1973, 1977) who
also investigated some shape effects.
Other ways of defining erosion criteria have been proposed, using the particle
settling velocity as a characterizing variable. Collins and Rigler (1982), Komar and
Clemens (1986) proposed relationships for the critical shear stress or shear velocity
in terms of the settling velocity.
The average flow velocity is an appealing indicator of bed stability, although it
does not uniquely represent flow conditions at the bed; the flow depth should also
be accounted for in some way. One of the earliest known criteria for erosion was
proposed in 1753 by Brahms, in terms of average flow velocity, i.e.

Vc ¼ kW 1=6 ð9:6Þ

where Vc is the critical average velocity, k is an empirical constant and W is the


weight of the particle. This relationship shows that the critical velocity is only
weakly related to particle size. For example, doubling the weight of a particle
(i.e. increasing its diameter by about a quarter) implies an increase in critical
velocity of only about an eighth.
9.3 Empirical Approach 325

Yang (1973) succeeded in formulating an incipient motion criterion in terms of


the average flow velocity and particle fall velocity. He did this by expressing the
drag and lift coefficients in terms of the particle fall velocity and analysing the
stability of the particle, assuming that movement would take place by sliding. He
assumed the appropriate velocity for drag and lift was at a distance equal to the
particle diameter above the bed, and related this to the average velocity by inte-
grating the velocity profile. This analysis produced the general form of a rela-
tionship between Vc /w and Re* which is hyperbolic for Re* < 5 and constant for
Re* > 70 (Fig. 9.4). By calibrating with available data, he produced the following
relationships:

Vc 2:5
¼ þ 0:66 for Re \70 ð9:7Þ
w log Re  0:06

and

Vc
¼ 2:05 for Re [ 70 ð9:8Þ
w

(Yang’s 1973 comments on the use of shear stress or velocity to define incipient
motion make interesting reading.)

Fig. 9.4 A riprap-protected channel bank


326 9 Scour and Scour Protection

One of the major problems in the empirical approach to establishing incipient


motion conditions is the decision as to what constitutes sediment movement. Most
theoretical and empirical relationships are in terms of time-averaged conditions. In
fact, it is not a time-averaged force that moves a particle, but an instantaneous
occurrence of a fluctuating force. The fluctuations are associated with turbulence
and are randomly distributed in time. Therefore, if a single particle is observed, the
definition of whether a particular flow condition is able to move the particle
depends on the period of observation. Lavelle and Mofjeld (1987) query whether it
is reasonable to define a critical condition at all; entrainment should properly be
considered probabilistically, such as by Einstein (1950) in his bedload model.
(A comparison of Einstein’s (1950) entrainment function with the Shields criterion
(James 1988) suggests that the original Shields diagram represents a probability of
movement of about 50% in the fully turbulent region, i.e. for grain sizes greater
than about 2 mm.) Neill and Yalin (1969) and Yalin (1977) proposed criteria for
defining observation conditions to ensure consistent results. To assist in quantifying
incipient motion criteria, Shvidchenko and Pender (2000) defined the intensity of
motion as
m
I¼ ð9:9Þ
Nt

in which m is the number of particle displacements during time interval t and N is


the total number of particles over the sample area. The incipient motion condition
then represents a specified, very small, intensity of motion rather than no motion at
all.
Armitage and Rooseboom (2010) adopted this measure and related it to the
‘movability number’ = u*/w (where w is the grain settling velocity). As a criterion
for incipient motion, they selected a value of I = 2  10−5 s−1, corresponding to
visual observation of the movement of the first few grains on a bed. Using previ-
ously published and their own data, they established the following relationships
between the critical value of u*/w and Re* for different ranges of Re*:

u 2:2
¼ for Re \6:2 ð9:10Þ
w Re 1:4

and
u
¼ 0:17 for Re [ 6:2 ð9:11Þ
w

For practical engineering applications, it is sufficient to bear in mind that the


Shields and other similar criteria do not represent absolute thresholds, but a certain
probability of movement, and to exercise appropriate caution.
Any of the criteria presented can be used for predicting erosion conditions, but
specifying settling velocities for rocks makes the criteria expressed in terms of
9.3 Empirical Approach 327

particle settling velocity inappropriate for designing loose rock protection. The
critical shear stress approach, as exemplified by the Shields criterion, is therefore
recommended for design.

9.4 Design Applications

Knowledge of the flow conditions at which sediment will begin to move is of great
value in designing unlined channels through cohesionless material and loose stone
linings (riprap) for channels subject to erosion (Fig. 9.4). Both the critical shear
stress and permissible velocity approaches are commonly used, and both have been
developed further for practical application. Guidelines have also been developed for
ensuring that the material below a riprap lining is stable.

9.4.1 Critical Shear Stress Design

Channel Beds

Predicting the likelihood of erosion and designing loose rock protection on plane
surfaces are easily done through application of the critical shear stress relationships
presented in the previous section. Direct application of the Shields criterion is
particularly useful and widely accepted.
Example 9.1
A wide rectangular channel is laid on a slope of 0.00050 and has a pebble bed with a
representative size of 12 mm. The channel conveys a discharge of 1.0 m3/s/m towards a
vertical drop structure. A layer of loose stones is to be used to prevent erosion near the
drop structure.

a. Show that the pebble bed is stable under uniform flow conditions.
b. Determine a size for the protecting stones.
c. Determine the distance over which the protecting layer should extend.
Manning’s n for large bed particles can be calculated using the Strickler formula

D1=6
n¼ pffiffiffi
6:7 g

where D is the particle size in metres.


Solution
For a wide channel, the hydraulic radius, R, can be approximated by the flow depth, y.

a. For stability so \soc


where so is the bed shear stress and soc is the critical shear stress.
328 9 Scour and Scour Protection

so is given by

so ¼ c yo Sf

yo from Manning (for a wide channel, R  y):

y
q ¼ y2=3 So1=2
n

Therefore
!3=5
qn
y¼ 1=2
So

with
D1=6 0:0121=6
n¼ pffiffiffi ¼ pffiffiffiffiffiffiffi ¼ 0:023
6:7 g 6:7 9:8

So
!3=5  3=5
qn 1:0  0:023
y¼ 1=2
¼ ¼ 1:017 m
So 0:000501=2

and
so ¼ 9:8  103  1:017  0:00050 ¼ 4:98 N/m2

soc is given by
soc ¼ sc c DðSs  1Þ

with
sc ¼ 0:047 ðShields for large DÞ and Ss ¼ 2:65

Therefore
soc ¼ 0:047  9:8  103  0:012  ð2:65  1Þ ¼ 9:12 N/m2

so \soc and so the bed is stable.

b. The bed is most vulnerable near the brink, where the flow depth approaches critical, i.e.
sffiffiffiffiffi rffiffiffiffiffiffiffiffiffi
3 q
2 3 1:0
2
y ¼ yc ¼ ¼ ¼ 0:467 m
g 9:8

The protective stone size is given by the Shields critical shear relationship
9.4 Design Applications 329

c y Sf
sc ¼
cDðSs  1Þ

Therefore

y Sf

sc ðSs  1Þ

with

V 2 n2
Sf ¼ (Manning)
y4=3
q

y
D1=6
n¼ pffiffiffi
6:7 g

So

ðq=yÞ2 D1=6
Sf ¼ pffiffiffi
y4=3 6:72 g

Therefore

 3=2
q2
D¼ ðAÞ
sc ðSs  1Þy7=3 6:72 g
 3=2
1:02
¼ ¼ 0:072 m
0:047  ð2:65  1Þ  0:4677=3  6:72  9:8

A slightly larger size should be chosen, say 0.080 m.

c. The protection should extend until the flow depth is greater than the depth below which
erosion of the 12 mm material would occur. The water surface over the protected reach
follows an M2 profile from the critical depth at the drop structure.

M2
M2
ymin
yc

Equation (A) above can be rearranged to calculate the critical flow depth for a given bed
material size, i.e.
330 9 Scour and Scour Protection

 3=7
q2
ymin ¼
sc ðSs  1ÞD2=3 6:72 g
 3=7
1:02
¼ ¼ 0:78 m
0:047  ð2:65  1Þ  0:0122=3  6:72  9:8

The length of protection required is therefore the distance along the M2 profile from
yc = 0.47 m to ymin = 0.78 m. This can be computed using the Direct Step Method, as
performed in the table below:
y (m) V (m/s) E (m) DE (m) Sf Sf (ave) So − Sf x (m) x (m)
0.47 2.128 0.701 0.01191 0
0.003 0.01080 −0.01030 −0.30
0.50 2.000 0.704 0.00969 0.30
0.015 0.00837 −0.00787 −1.85
0.55 1.818 0.719 0.00705 2.16
0.023 0.00616 −0.00566 −4.07
0.60 1.667 0.742 0.00527 6.23
0.062 0.00422 −0.00372 −16.80
0.70 1.429 0.804 0.00316 23.02
0.060 0.00268 −0.00218 −27.43
0.78 1.282 0.864 0.00220 50.46
The length of protection should therefore be greater than 50 m, say 55 m

Channel Banks

The previous results must be extended for application to channel banks. Bank
particles are inherently less stable than bed particles because of their downslope
weight components. They are also exposed to boundary shear stresses different
from the cross-section average, as given by s = cRS, and the distribution of local
values over the section should be considered for efficient design. Both the applied
shear stress and the resistance to movement vary over the cross section.
A sediment particle on a channel bank may be expected to be less stable than one
on the bed because it has a downslope weight component. The shear stress required
to displace it is therefore less than for a bed particle. The critical shear stress for a
sediment particle on the bank can be expressed in terms of the critical shear stress
for a similar particle on the bed by including the downslope weight component in
an analysis of its stability.
The forces acting on a particle on the plane bank of a channel inclined at an
angle a to the horizontal include its submerged weight (W), the force applied by the
flow (F) and a force resisting movement (R) (Fig. 9.5).
9.4 Design Applications 331

Fig. 9.5 Forces on a bank R


particle (solid vectors are in
the plane of the bank, dashed
is vertical and dotted is
normal to the bank)

F
N T
W
α

The submerged weight has a component normal to the plane of the bank equal to

N ¼ W cos a

and a component down the slope of the bank equal to

T ¼ W sin a

The submerged weight, and hence N and T, can be expressed in terms of the
critical shear stress on a horizontal bed. Considering the particle to move by sliding

sc / W tan u

where / is the angle of repose of the material and tan / is the coefficient of friction.
Therefore
sc
W/
tan u

The downslope weight component can therefore be expressed as


sc
T/ sin a
tan u

and the component normal to the bank can be expressed as


sc
N/ cos a
tan u

The shear force associated with the flow is assumed to act parallel to the lon-
gitudinal slope of the channel, but may actually be inclined as a result of secondary
flow in the channel. As the incipient motion condition is being considered, this
332 9 Scour and Scour Protection

force is defined as being proportional to the critical value of shear stress on the
bank, scb, i.e.

F / scb

The force resisting motion can be expressed in terms of limiting friction, i.e.

R ¼ N tan u ¼ W cos a tan u

and, substituting the above proportionality for W,


sc
R/ cos a tan u
tan u

and so

R / sc cos a

At incipient motion, the forces on the particle in the plane of the bank are
therefore as shown in Fig. 9.6.
The proportionality constant relates force on a particle to boundary shear and can
be assumed to be the same for all forces.
At incipient motion, the forces are in equilibrium, and are therefore related by

R2 ¼ F 2 þ T 2

Making the above substitutions and cancelling the proportionality constant gives

s2c
s2c cos2 a ¼ s2cb þ sin2 a
tan2 u

whence
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
tan2 a
scb ¼ sc cos a 1  2 ð9:12Þ
tan u

Fig. 9.6 Forces on bank R ∝ τ c cos α


particle at incipient motion

F ∝ τ cb τc
T∝ sin α
tan ϕ
9.4 Design Applications 333

which can also be expressed as


sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
sin2 a
scb ¼ sc 1  2 ð9:13Þ
sin u

The above analysis was suggested in principle by Forchheimer (1924) and


introduced for design applications by Lane (1953). Equations (9.12) and (9.13) give
the critical shear stress capable of moving a particle on a bank with inclination a in
terms of the critical shear stress which would move it on the bed of the channel (as
defined by the Shields or some other criterion) and the angle of repose of the
material. The critical shear stress on the bank is always less than on the bed, but this
does not necessarily mean that bank particles are less stable than those on the bed
because the applied shear stress is less on the banks than on the bed.
In order to apply the above criterion for stability on the banks, it is necessary to
know first a value for the friction angle for the sediment, and second the actual
distribution of shear stress in terms of flow conditions.
The friction angle can be assumed to be equal to the angle of repose of the
material. This is quite easily determined experimentally and Lane (1953) has pre-
sented a graphical, empirical relationship between angle of repose, particle size and
particle shape (Fig. 9.7).
Shear stress is not uniformly distributed over the boundary (so = cRS gives the
average over the cross section). Lane (1953) determined the shear stress distribution

Fig. 9.7 Angles of repose of 42


non-cohesive sediment
(adapted from Lane 1953) 40

38

36
Angle of repose (o)

34

32

30

28

26

24

22

20
5 10 20 50 100
Particle size (mm)
334 9 Scour and Scour Protection

Table 9.1 Permissible shear Degree of sinuosity Relative limiting shear stress
stress adjustment for channel
sinuosity (Lane 1955) Straight 1.00
Slightly sinuous 0.90
Moderately sinuous 0.75
Very sinuous 0.60

experimentally for various cross-sectional shapes. He showed that for rectangular


and trapezoidal channels, the local shear stress does not exceed cyS on the bed and
about 0.76 cyS on the banks for any width–depth ratio and side slope (y is the
maximum flow depth).
It has also been found that channel sinuosity increases the likelihood of scour.
Lane (1955) proposed adjusting the permissible shear stress by the factors listed in
Table 9.1 to account for this effect. Using data from various sources, Lane (1953)
also showed that the critical shear stress is increased by a sediment load, this effect
becoming more pronounced and significant as the particle size decreases below
about 5 mm.
Example 9.2
A trapezoidal channel with a bottom width of 4.0 m and side slopes of 2.5H:1V is
excavated on a slope of 0.00080 in alluvial material with a representative particle size of
18 mm and an angle of repose of 33°. Determine the maximum discharge for which
there will be no erosion in the channel.
Manning’s n for large bed particles can be calculated using the Strickler formula

D1=6
n¼ pffiffiffi
6:7 g

where D is the particle size in metres.


Solution
The maximum discharge can be found from the Manning equation for the maximum
flow depth that would induce erosion which, in turn, can be found through the Shields
criterion. The applied and resisting shear stresses are different for the bed and banks, and
must be found separately.
For the bed:
The critical shear stress is found from the Shields criterion for large particles, i.e.

sc
¼ 0:047
c DðSs  1Þ

Therefore

sc ¼ c DðSs  1Þ0:047
¼ 9:8  103  0:018  ð2:65  1Þ  0:047 ¼ 13:7 N/m2
9.4 Design Applications 335

The maximum applied shear stress is

so ¼ c ymax S

Therefore, at incipient erosion

so ¼ sc

i.e.
c ymax S ¼ 13:7

so

13:7
ymax ¼ ¼ 1:75 m
9:8  103  0:00080
For the banks:
The critical shear stress is given by
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
tan2 h
scb ¼ sc cos h 1  2
tan /

where

1
h ¼ arctan ¼ 21:8
2:5
Therefore
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
tan2 21:8
scb ¼ 13:7  cos 21:8  1  ¼ 10:0 N/m2
tan2 33

The maximum applied shear stress is

so ¼ 0:76 c ymax S

Therefore, at incipient erosion

so ¼ sc

i.e.

0:76 k ymax S ¼ 10:0

So
10:0
ymax ¼ ¼ 1:68 m
0:76  9:8  103  0:00080
The critical flow depth is lower for erosion of the banks than for the bed, so
ymax = 1.68 m.
The discharge is given by the Manning equation combined with continuity, i.e.
336 9 Scour and Scour Protection

A 2=3 1=2
Q¼ R S
n

where
1
A ¼ Bymax þ 2  ðsymax Þymax
2
¼ 4:0  1:68 þ 2:5  1:682 ¼ 13:8m2
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P ¼ B þ 2 ðsymax Þ2 þ y2max
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
¼ 4:0 þ 2  ð2:5  1:68Þ2 þ 1:682 ¼ 13:0m
A 13:8
R¼ ¼ ¼ 1:06m
P 13:0
1=6
D 0:0181=6
n¼ pffiffiffi ¼ pffiffiffiffiffiffiffi ¼ 0:024
6:7 g 6:7  9:8

Therefore

13:8 2=3 1=2


Q¼ R S
n
13:8
¼ 1:062=3 0:000801=2 ¼ 16:9 m3 =s
0:024

9.4.2 Permissible Velocity Design

Various methods for defining bed stability in terms of a critical velocity have been
proposed, but are difficult to formulate because particle movement is not related to
the cross-section average velocity, but rather to the boundary shear stress or a
velocity near the bed. Criteria in terms of the average velocity usually require
specification of the flow depth as well. Criteria in terms of a local velocity near the
bed require estimation of this velocity through a velocity distribution equation.
However, the commonly used logarithmic distribution equation includes u*, which
depends on the flow depth. (Yang (1973) followed this approach in developing his
critical velocity criterion (Eqs. (9.7) and (9.8)), by relating the near-bed velocity to
the average velocity through integration of the velocity profile, also requiring
knowledge of u* and hence the flow depth.)

The HEC-11 Procedure for Riprap Rock Sizing

Recognizing a preference in engineering practice for using flow velocity rather than
bed shear stress as a basis for designing riprap, the United States Federal Highway
Administration reformulated the Shields criterion and expressed it as a relationship
9.4 Design Applications 337

between the minimum rock size and the average flow depth and velocity (Brown
and Clyde 1988). Although it offers little advantage over direct application of the
Shields criterion, its development is included here because of its common use.
The stability factor (SF) is defined as the ratio of the critical shear stress at
incipient motion (sc) to the applied shear stress (so),
sc
SF ¼ ð9:14Þ
so

with stability requiring a value greater than 1.2, based on observations of stable and
unstable linings. The applied shear stress is given by

so ¼ c R S ð9:15Þ

where c is the specific weight of the rock material, R is the hydraulic radius and S is
the energy gradient. The critical shear stress can be expressed in terms of the
dimensionless Shields parameter (s*c) as

sc ¼ K1 sc cD50 ðSs  1Þ ð9:16Þ

in which D50 is the median rock size and Ss is the specific gravity of the rock
material. The factor K1 accounts for a transverse slope of the surface (such as a
sloping channel bank) and is given by (from Eq. (9.13))
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
sin2 a
K1 ¼ 1  2 ð9:17Þ
sin u

where a is the side slope angle and / is the angle of repose of the riprap material.
The required value of SF can therefore be expressed as

K1 sc c D50 ðSs  1Þ


¼ 1:2 ð9:18Þ
cRS

The value of s*c = 0.047, as indicated on the Shields diagram for large particles,
was found to be supported by field observation for typical riprap rocks. For sim-
plicity, it was assumed that R = d, the average flow depth in the main part of the
channel (between flood plains). A typical value of Ss = 2.65 was assumed, but a
correction for different values allowed for subsequently. The energy gradient can be
determined from the Manning equation as

V 2 n2
S¼ ð9:19Þ
R4=3
338 9 Scour and Scour Protection

where V is the average flow velocity and n is the Manning resistance coefficient,
which can be related to the median rock size by

1=6
n ¼ 0:0482 D50 ð9:20Þ

Making these substitutions, Eq. (9.18) can be rearranged to give the minimum
stable rock size as

0:0068 V 3
D50 ¼ 3=2
ð9:21Þ
d 1=2 K1

This size can be adjusted to account for a different rock material specific gravity
(Ss) by multiplying D50 by the factor

2:12
Csg ¼ ð9:22Þ
ðSs  1Þ3=2

and for an SF different from 1.2 by multiplying by the factor


 3=2
SF
Cf ¼ ð9:23Þ
1:2

The method is intended for use with uniform or moderately gradually varied
flows. SF is used to allow for uncertainty associated with departures from these
conditions according to Table 9.2.
Note that although the method accounts for the reduced stability of rocks on side
slopes (through factor K1), it uses the average applied shear stress and does not
allow for the reduction of local shear on the banks (as by Lane (1953)). Because
erosion is related to the boundary shear, the required rock size by this method
requires specification of both flow depth and velocity (Eq. (9.21)). Within the
gradually varied flow regions associated with control structures, it is therefore

Table 9.2 Guidelines for selecting stability factor values adapted from Brown and Clyde (1988)
Condition SF
range
Uniform flow; straight or mildly curving reach (curve radius/channel width >30); 1.0–12
minimum impact of waves and floating debris; little or no uncertainty in design
parameters
Gradually varied flow; moderate bend curvature (30 > curve radius/channel 1.3–1.6
width >10); moderate impact of waves and floating debris
Approaching rapidly varied flow; sharp bend curvature (curve radius/channel 1.6–2.0
width <10); significant impact potential from floating debris; significant wind and/
or boat-generated waves (0.3–0.6 m); high flow turbulence; turbulently mixing flow
at bridge abutments; significant uncertainty in design parameters
9.4 Design Applications 339

necessary to perform gradually varied flow computations to establish an appropriate


velocity/depth combination for design.
The method for determining the D50 rock size has been presented here. The
HEC-11 report also provides guidelines for specifying the rock grading, the lining
thickness and filter layer sizes. It also makes recommendations for complete lining
design and construction, including revetment types other than riprap.

9.4.3 Protection of Underlying Material

It is not sufficient to determine a single stable stone size for a lining. If the size of
riprap is considerably larger than the underlying material, then leaching of this
material may occur by local erosion and removal by flow which penetrates the
riprap blanket. Leaching is minimized if
– the riprap layer is thick,
– interstices are closed,
– base material is cohesive or
– a protective layer is placed between the riprap and bed material.
Recommendations for thickness of riprap layers have been made and vary from 1.3
to 2 times the D50 size of the material. Other recommendations are that the
thickness of the layer should be up to 1.5 times the maximum particle size.
Protection these days is very commonly provided by a layer of geotextile. An
alternative is to place layers of stone or soil with grading characteristics that will
ensure protection. The requirements of grading characteristics for adjacent layers
are usually specified in terms of size relationships, such as the following set of
criteria (Lambe and Whitman 1969):

D15 upper
\5 ð9:24Þ
D85 lower

D15 upper
4\ \ 20 ð9:25Þ
D15 lower

D50 upper
\ 25 ð9:26Þ
D50 lower

These three equations can first be applied to the riprap and the base material to
determine if any intermediate filter layer is in fact required. If any of these criteria
are violated, then a filter layer or layers will be required and the same equations can
be used for the filter design. The criteria must hold for comparison of riprap with
the filter layer, as well as for comparison of the filter layer with the underlying
material. It may not be possible to find a material satisfying all six criteria, in which
340 9 Scour and Scour Protection

cases multiple filter layers are required and the equations must be satisfied for each
layer relative to the layers immediately above and below it.
Another approach for checking if an intermediate layer is required is to apply the
Shields criterion to the underlying material using the hydraulic radius of the rockfill
interstices. This can be estimated (Leps 1973) as

D
R¼ ð9:27Þ
8

in which D is the size of the riprap material.


For non-cohesive material larger than about 2 mm, the constant form of the
Shields criterion can be used, i.e.

sc cRS
¼ ¼ 0:047
c Ds ðSs  1Þ c Ds ðSs  1Þ

in which Ds is the grain size of the underlying material.


Substituting for R from Eq. (9.27),
D
8S
¼ 0:047
Ds ðSs  1Þ

which gives the maximum size for overlying particles as

ðSs  1ÞDs
D¼ ð9:28Þ
2:7S

If underlying particles are smaller than about 2 mm, then the constant form of
the Shields criterion cannot be used and the maximum size of overlying material
must be determined using the Shields diagram.
As in all shear stress relationships, S in Eq. (9.28) is the energy gradient. For this
application, it may be estimated from the channel flow velocity (V) by using the
Manning–Strickler equation with the product nD (where n is the layer porosity)
representing the bed roughness, i.e.
 1=6
R pffiffiffiffiffiffiffiffiffiffi
V ¼ 7:7 gRS
nD

For a wide flow area, R can be replaced by the flow depth, y, and the equation
rearranged to give

 V 2 nD1=3
7:7 y
S¼ ð9:29Þ
yg
9.4 Design Applications 341

This approach gives a relationship between sizes of adjacent layers of material to


ensure no leaching of the finer material through the coarser. It can be used suc-
cessively to design multilayer filters.

9.5 Scour Around Bridge Piers

The foregoing treatment of scour applies only to situations where the boundary
shear is applied through unidirectional flow parallel to the surface. Instream
structures or channel modifications may disrupt the uniform flow and create regions
of relatively high intensity leading to local scour. The local increase in scour
potential may result from contraction of the channel by parallel side walls, such as
for bridge crossings (Dey and Reikar 2005), the inducement of jets, such as through
outlet structures, sluice gates and culverts (Melville and Lim 2013), pressure flow
through submerged bridges (Melville 2014) or the formation of vortices around
isolated structural elements, such as bridge piers (Sheppard et al. 2014).
Under a sustained steady flow, such local scour progresses with time and sta-
bilizes at an equilibrium depth. Because the peak discharge during a design storm
usually has a much shorter duration than the time taken for equilibrium to be
reached, the maximum scour depth achieved is usually less than the equilibrium
value. Two conditions of scour hole development are recognized. Clear-water scour
occurs when the shear stress of the approaching flow is insufficient to mobilize the
bed material, and so no sediment is supplied to the hole from upstream. Live-bed
scour occurs when there is general movement of bed sediment in the approach flow
so that some of the scoured sediment is continuously replaced. Because of the
higher flow intensity associated with live-bed scour, progression is generally more
rapid than with clear-water scour, but the equilibrium depth is slightly less. The
equilibrium scour depth depends on the sediment size, and may be greater for
uniform than nonuniform sediments due to surface armouring during the entrain-
ment of particles.
Bridge piers present a common and important instance of local scour and their
design requires estimation of the anticipated scour depth. The scour results from the
action of a complex vortex system induced by the interaction between the approach
flow and the pier geometry, illustrated for a single circular pier in Fig. 9.8. The
deceleration towards the pier results in a surface bow wave or roller against its
upstream surface and a strong downward flow jet. The downward jet initiates scour
at the bed and the formation of a ‘horseshoe vortex’ that follows the leading surface
of the pier and persists for a few pier diameters downstream. Flow separation from
the sides of the pier results in alternating wake vortices from either side that interact
with the horseshoe vortex near the bed. This vortex action causes the development
of a scour hole around the pier and an accumulation of the eroded sediment further
downstream (Fig. 9.9).
342 9 Scour and Scour Protection

pier
bow wave
wake vortex

downward flow jet


original bed level

horseshoe vortex

Fig. 9.8 Vortex structure around a single circular pier

9.5.1 Scour Depth Estimation

Ettema et al. (2017) presented a structured design approach for estimating scour
depth, identifying four levels of pier form and situation complexity for which
different approaches are necessary: (i) Simple, single cylindrical piers can be
designed using semi-empirical equations for scour depth, and the Sheppard/

Fig. 9.9 A scour hole around a bridge pier (flow from left to right) (photograph by M. Skeen)
9.5 Scour Around Bridge Piers 343

Melville method (Sheppard et al. 2011, 2014) (presented below) is recommended.


(ii) Common standard pier forms, which may include components such as pile caps
and pier groups, can also be designed using semi-empirical equations provided
these include factors to account for specific geometric characteristics, but laboratory
model testing should also be considered. (iii) Common standard pier forms in
complicating situations such as debris or ice accumulation, abutment proximity or
bridge deck submergence require additional sophistication of semi-empirical
methods and usually model testing and possibly numerical (CFD) flow field
modelling. (iv) Complex or unusual pier situations require laboratory and numerical
modelling. Only the first category (simple, single cylindrical piers) is considered
further here.

Equilibrium Scour Depth

Many semi-empirical equations have been proposed for estimating equilibrium


scour depths. Sheppard et al. (2014) evaluated 23 commonly used equilibrium
scour equations using laboratory and field data from a variety of sources. (Details of
the equations are presented in Sheppard et al. (2011).) They found the Sheppard/
Melville (S/M) method to be the most accurate and recommended it for design of
simple piers in cohesionless sediment. The method is able to predict the equilibrium
scour depth for both clear-water and live-bed conditions, both of which are relevant
for design—piers in main channel sections are vulnerable to live-bed scour during
floods but piers on flood plains may be exposed only to clear-water scour during
occasional relatively low intensity flows. The method does not account for sediment
nonuniformity and therefore produces conservative predictions for sediments with
large size gradations.
The threshold between clear-water and live-bed scour occurs at an average
approach velocity (V0) equal to the critical entrainment velocity (Vc) above which
sediment movement occurs, supplying material to the scour hole (Fig. 9.10). The
critical entrainment velocity is calculated for the D50 sediment size from the Shields
criterion using Eqs. (9.30)–(9.33).
8 9
< 73:5y0 =
Vc ¼ 2:5uc ln h i
:D Re ð2:85  0:58 ln Re þ 0:0020Re Þ þ 111  6 ; ð9:30Þ
50
Re

for5  Re  70

or
 
2:21 y0
Vc ¼ 2:5uc ln for Re [ 70 ð9:31Þ
D50
344 9 Scour and Scour Protection

yse max

yse /a*

live-bed scour

clear-water scour

1.0 V0p /Vc


V0 /Vc

Fig. 9.10 Variation of equilibrium scour depth with flow velocity (adapted from Melville and
Chiew 1999)

in which y0 is the average approach flow depth and u*c is the critical shear velocity,
given by

1=2
9:09  106
uc ¼ 16:2 D50  D50 ð38:76 þ 9:6 ln D50 Þ  0:0050 ð9:32Þ
D50

Re* is the shear Reynolds number,

u k s
Re ¼ ð9:33Þ
m

with the roughness height ks given by 2.5D50 if D50 0.6 mm or 5.0D50 if


D50 < 0.6 mm.
For clear-water scour, the equilibrium scour depth (yse), normalized by the
effective pier diameter (a*), is given by

yse V0
¼ 2:5 f1 f2 f3 for 0:4   1:0 ð9:34Þ
a Vc

The effective pier diameter is given by

a ¼ K s ap ð9:35Þ

in which ap is the projected width of the pier and Ks is a shape factor with the value
9.5 Scour Around Bridge Piers 345

Ks ¼ 1:0 for circular piers ð9:36Þ

or
 p 4

Ks ¼ 0:86 þ 0:97 a  for rectangular piers ð9:37Þ
4

where a is the flow skew angle in radians. (It should be borne in mind that the local
flow direction may be rather different during flood conditions than at low flows.)
The f factors in Eq. (9.34) are calculated as
 

y0 0:4
f1 ¼ tanh ð9:38Þ
a
 
2
V0
f2 ¼ 1  1:2 ln ð9:39Þ
Vc
 
a
D50
f3 ¼  1:2  0:13 ð9:40Þ
a a
0:4 D50 þ 10:6 D50

Live-bed scour occurs for approach flow velocities greater than the critical
entrainment velocity (V/Vc > 1.0) and reaches a maximum at an approach velocity
(V0p) at which the bed planes out and bed forms disappear (Fig. 9.10). This velocity
is given by the greater of

V0p ¼ 5Vc ð9:41Þ

or
pffiffiffiffiffiffiffiffi
V0p ¼ 0:6 g y0 ð9:42Þ

The live-bed equilibrium scour depth for approach velocities between the
live-bed threshold and the maximum value is given by
" ! V0p
!#
Vc  1 Vc  Vc
V0 V0
yse V0 V0p
¼ f1 2:2 þ 2:5 f3 for 1:0   ð9:43Þ
a V0p V0p
Vc  1 Vc  1
Vc Vc

The maximum live-bed scour is given by

yse V0 V0p
¼ 2:2 f1 for [ ð9:44Þ
a Vc Vc
346 9 Scour and Scour Protection

The HEC-18 equation for predicting the maximum scour depth (Arneson et al.
2012) is widely used and performed well in the evaluation by Sheppard et al. (2011,
2014). For clear-water and live-bed scour with relatively fine sediment
(D50 < 20 mm),

yse y 0:35
0
¼ 2:0 K1 K2 K3 Fr0:43
0 ð9:45Þ
a a

in which a is the pier width (the short dimension), K1 accounts for pier shape, K2 for
the alignment angle of the approach flow (h) and K3 for the bed-form regime. For
h > 5°, the shape influence is small compared with the alignment influence and
K1 = 1.0 for all shapes; for h < 5°, K1 = 1.0 for circular and rounded piers, 1.1 for
square-nosed piers and 0.9 for sharp-nosed piers. For non-circular piers, the value
of K2 depends on the alignment and the geometry according to
 0:65
L
K2 ¼ cos a þ sin a ð9:46Þ
a

where L is the length of the pier. K3 = 1.1 for clear-water conditions and beds with
plane surfaces or antidunes or dunes smaller than about 3 m, increasing to 1.3 for
dunes larger than about 10 m. A ‘rule-of-thumb’ is recommended that
yse < 2.4a for Fr0 < 0.8 and yse < 3.0a for Fr0 > 0.8.
Arneson et al. (2012) provide a separate equation for relatively coarse
(D50 > 20 mm) and graded (r 1.5) bed sediments, under clear-water (V0/
Vc < 1) conditions, i.e.
 
yse F2d
¼ 1:1 K1 K2 tanh ð9:47Þ
a0:62 y0:38
0 1:97 r1:5

in which Fd is the densimetric particle Froude number (=V0/(g(Ss − 1)D50)0.5) and r


(=(D84/D16)0.5) is the geometric standard deviation of the sediment size distribution.
This equation was revised by Shan et al. (2016) to extend the conditions of its
applicability as
 
yse F2d
¼ 1:32K1 K2 K3 tanh ð9:48Þ
a0:62 y0:38
0 1:97r1:5

Equation (9.48) is able to predict the scour depth for clear-water scour and for
live-bed scour with V0/Vc up to 5.2 and Fr0 up to 1.7, and for sediments with
0.21 mm < D50 < 127 mm and r up to 7.5.
9.5 Scour Around Bridge Piers 347

Scour Depth Evolution

The actual scour depth depends on the duration of the scouring flow relative to the
development of the scour hole with time. Live-bed scour during floods progresses
rapidly and the equilibrium scour depth is representative (Melville and Chiew
1999). Clear-water scour develops very slowly, however, and may be intermittent
for piers on flood plains. Various methods have been proposed for estimating the
time for equilibrium scour to develop and for describing the evolution of the scour
hole towards equilibrium.
Sheppard et al. (2011) described and evaluated ten scour evolution methods.
They recommended use of the clear-water method proposed by Melville and Chiew
(1999) for circular piers, with revised coefficient values and the equilibrium scour
depth calculated by the S/M method described above. This is referred to as the
Melville/Sheppard (M/S) method and can be used for live-bed as well as clear-water
conditions. The scour depth (ys) at any time t after the commencement of scour is
given by
(   1:6 )
ys Vc t

¼ exp 0:04 ln ð9:49Þ
yse V0 te

in which te is the time to equilibrium, given by


 
a V0 y0 V0
te ðdaysÞ ¼ 200  0:40 for [ 6; [ 0:4 ð9:50Þ
V0 Vc a Vc

or
  
a V0 y0 0:25 y0 V0
te ðdaysÞ ¼ 127:8  0:40 for  6; [ 0:4 ð9:51Þ
V0 Vc a a Vc

Because scour approaches the equilibrium depth asymptotically, a time to


equilibrium has little practical meaning and Sheppard et al. (2011) also suggest the
time to 90% of equilibrium scour as a more useful measure. This is given by
 
V0
t90 ¼ exp 1:83 te ð9:52Þ
V

Oliveto and Hager (2002, 2005) proposed an equation for time-dependent


clear-water scour, also applicable for nonuniform sediments, which is independent
of the time to equilibrium, i.e.
ys
¼ 0:068 N r0:5 F1:5
d log T ð9:53Þ
yR
348 9 Scour and Scour Protection

in which yR = (y0D2)1/3 for a circular pier with diameter D or yR = (y0b2)1/3 for an


abutment with width b, N = 1 for a circular cylindrical pier or 1.25 for a vertical
abutment. Fd is the densimetric particle Froude number (=V0/(g(Ss − 1)D50)0.5) and
T = t/tr with tr = yR/(r1/3(g(Ss−1)D50)0.5). Oliveto and Hager limit the applicability
of Eq. (9.53) to clear-water scour by specifying a maximum upstream densimetric
Froude number of 1.2 for the local initiation of motion, and to sediments with D50
less than 5–10 times y0.
These equations for describing the time variation of the scour depth can be
applied to determine the pre-equilibrium depth for a specified duration of design
discharge, or to a time-varying discharge for a design hydrograph if necessary. The
scour depth resulting from a single flow episode may be considerably less than the
equilibrium value; Melville and Chiew (1999) estimate that the scour hole develops
to between 50 and 80% of the equilibrium depth after the initial 10% of the time to
equilibrium under steady flow conditions. Because of the absence of sediment
supply for clear-water scour, no infilling would occur between intermittent events
and scour would be cumulative. The total duration of scouring discharges through
the structure’s design life should therefore be considered in relation to the scour
development time. For live-bed conditions, prediction of time-dependent scour
development would require simulation including sediment transport calculations to
determine the rate of sediment supply to the scour hole.
Relationships similar to those for piers have also been developed for abutments,
for example, by Cardoso and Fael (2010), Dey and Barbhuiya (2004, 2005),
Mohammedpour et al. (2017) and Yanmaz and Kose (2009).
Example 9.3
A rectangular, sharp-edged bridge pier 20 m long and 4.0 m wide is founded in sedi-
ment with D50 = 1.0 mm. The design flow, lasting approximately 4 h has a depth of
2.0 m and a velocity of 1.5 m/s. Using the S/M and M/S methods, determine

a. the equilibrium scour depth and


b. the scour depth after the design flow duration.

Solution

a. Determine whether clear-water or live-bed scour will occur under the design conditions
by calculating the ratio of V0 to Vc. Vc is given by Eq. (9.30) or (9.31) depending on the
flow condition defined by Re*.

u ks
Re ¼
m

At the critical condition (from Eq. (9.32))


9.5 Scour Around Bridge Piers 349


1=2
9:09  106
u ¼ uc ¼ 16:2 D50  D50 ð38:76 þ 9:6 ln D50 Þ  0:0050
D50

1=2
9:09  106
¼ 16:2  0:0010  0:0010  ð38:76 þ 9:6 ln 0:0010Þ  0:0050
0:0010
¼ 0:023 m/s

For D50 0.6 mm

ks ¼ 2:5 D50 ¼ 2:5  0:0010 ¼ 0:0025 m

Therefore
0:023  0:0025
Re ¼ ¼ 57:5
1  106

which is greater than 70 and so, from Eq. (9.31)


 
2:21 y0
Vc ¼ 2:5uc ln
D50
 
2:21  2:0
¼ 2:5  0:023  ln ¼ 0:48 m/s
0:0010

Therefore

V0 1:50
¼ ¼ 3:1
Vc 0:48

and so live-bed scour will occur. The approach velocity at which maximum live-bed
scour occurs is the greater of

V0p ¼ 5Vc ¼ 5  0:48 ¼ 2:4 m/s

or
pffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
V0p ¼ 0:6 gy0 ¼ 0:6  9:8  2:0 ¼ 2:7 m/s

Therefore
V0p 2:7
V0p ¼ 2:7 m/s and ¼ ¼ 5:6
Vc 0:48

V0p > V0c, so the equilibrium scour depth is given by Eq. (9.43), i.e.
" ! V0p !#
Vc  1 Vc  Vc
V0 V0
yse
¼ f1 2:2 þ 2:5 f3
a V0p V0p
Vc  1 Vc  1
350 9 Scour and Scour Protection

with
 

y0 0:4
4m f1 ¼ tanh
a

a ¼ Ks ap
20 m 
ap ¼ 4 cos 15 þ 20 sin 15


¼ 9:04 m
 p 4

Ks ¼ 0:86 þ 0:97 a 
4
 p 4

¼ 0:86 þ 0:97 0:262 
4
15o ¼ 0:933
(= 0.262 radians)

a ¼ 0:933  9:04 ¼ 8:43 m


" 0:4 #
2:0
f1 ¼ tanh ¼ 0:51
8:43
 
a
D50
f3 ¼  1:2  0:13
a 
0:4 D50 þ 10:6 Da50
 8:43 
¼  
0:0010
 
8:43 1:2 8:43 0:13
0:4 0:0010 þ 10:6 0:0010
¼ 0:41
   

yse 3:1  1 5:6  3:1



¼ 0:51  2:2 þ 2:5  0:41 
a 5:6  1 5:6  1
¼ 0:796

Therefore

yse ¼ 0:796  8:43 ¼ 6:71 m

b. The scour depth after the design flow duration is given by Eq. (9.49) with t = 4 h or
4/24 = 0.17 days, i.e.
(   1:6 )
ys Vc t

¼ exp 0:04 ln
yse V0 te

with te given by Eq. (9.50) or (9.51) depending on the value of y0/a*. Here yo/a* =
2.0/8.43 = 0.24, indicating use of Eq. (9.51), i.e.
9.5 Scour Around Bridge Piers 351

  
a V0 y0 0:25
te ðdaysÞ ¼ 127:8  0:40
V0 Vc a
 
8:43 2:0 0:25
¼ 127:8 ð3:1  0:40Þ
1:5 8:43
¼ 1350 days

Therefore
(   )
ys 1 0:17 1:6
¼ exp 0:04 ln
yse 3:1 1350
¼ 0:80
and ys ¼ 0:80  6:71 ¼ 5:4 m

9.5.2 Bridge Scour Countermeasures

If the anticipated scour depth is unacceptable, its development needs to be pre-


vented or restricted. This can be achieved either by protecting the bed around the
pier base with a stable surface, such as riprap, or by installing features that modify
the local flow pattern to reduce its scouring action.

Design of Riprap for Pier Protection

Riprap protection has the advantage over fixed protection devices of being flexible
and able to interact and move with the bed sediments (Chiew and Lim 2000). The
design of a riprap layer requires specification of the rock size, the level of the layer
relative to the original bed, the thickness of the layer and the extent of the layer.
Many recommendations for riprap design have been proposed, all based on data
obtained from laboratory experiments. Four of the more recent and complete
methods are presented below. The results obtained by using different methods
should not be expected to be entirely consistent, as different experimental condi-
tions and definitions of failure were used in their development.
For clear-water conditions, Chiew (1995) defined three modes of riprap failure.
Shear failure is the movement of the stones by the local shear stress associated with
the flow pattern around the pier. Winnowing failure is the removal of the underlying
bed material through the voids in the riprap layer. Edge failure arises from
undermining of the stones by scour of the unprotected bed at the periphery of the
layer. The possibility of each of these should be considered in the design.
Shear failure is prevented by ensuring that the riprap stones are large enough to
resist the scouring action. Chiew (1995) showed for circular piers that the approach
velocity (V0) at which pier scour commences is equal to 0.3 times the critical
velocity for scour without the pier (Vc). This is assumed to define the critical
condition for entrainment of riprap stones, provided it is adjusted to account for the
352 9 Scour and Scour Protection

influences of the size of the riprap relative to the pier diameter and the approach
flow depth relative to the pier diameter by the introduction of adjustment factors.
The relationship is then

V0 0:3
¼   ð9:54Þ
Vc K D K y0 
Dr50 D

in which D is the pier diameter and Dr50 is the median size of the riprap stones. The
adjustment factors, derived from laboratory test data, are given by
 
D D
K ¼1 for 50 ð9:55Þ
Dr50 Dr50
     
2
D D D D
K ¼ 0:398 ln  0:034 ln for 1   50 ð9:56Þ
Dr50 Dr50 Dr50 Dr50
y  y0
0
K ¼ 1 for 3 ð9:57Þ
D D
y  y 0:322 y0
0 0
K ¼ 0:783 0:106 for 0\ \3 ð9:58Þ
D D D

Equation (9.54) can be developed as a relationship between V0 and D50 to enable


sizing of the riprap. Vc can be expressed in terms of the shear velocity through
Manning’s equation with Strickler’s relationship (n = 0.042D1/6
r50) as

 
Vc y0 1=6
¼ 7:66 ð9:59Þ
uc Dr50

At the critical entrainment condition, from the Shields criterion for large
particles,

u2c
¼ 0:056 ð9:60Þ
gDr50 ðSs  1Þ

Combining Eqs. (9.59) and (9.60) and rearranging gives

Vc3
Dr50 ¼ pffiffiffiffiffi ð9:61Þ
388 y0

which can be expressed in terms of V0 through Eq. (9.54), i.e.


9.5 Scour Around Bridge Piers 353

  3   
3
V03 K DDr50 K yD0
Dr50 ¼ pffiffiffiffiffi ð9:62Þ
10:47 y0

for the minimum riprap stone size. Note that for D/Dr50 < 50 and/or y0/D < 3, the
solution must be found iteratively because Dr50 appears in the adjustment factor.
Unger and Hager (2006) conducted laboratory tests on single layers of uniformly
sized riprap placed in concentric rows around a circular pier. Failure was assumed
to have occurred when the first stone along the pier perimeter was dislodged by
rolling, undermining or sliding. Conditions at failure are described according to the
Shields criterion, with adjustment for the effects of the size of the pier relative to the
channel width (B), the size of the riprap (Dr) relative to the underlying material and
the extent of the riprap coverage. The basic Shields criterion for rough turbulent
flow was expressed by Hager and Oliveto (2002) as
 1=6
Vc R
r1=3 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ 1:65 ð9:63Þ
gD50 ðSs  1Þ D 50

The extended form for riprap around the pier, including the adjustments derived
from the test results, is
 1=6
V0c Ro
r1=3 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ 1:65
0:40
ð1  b1=4 Þd0:20n ð9:64Þ
gD50 ðSs  1Þ D50

in which D50 refers to the underlying material, V0c is the critical approach velocity
for the riprap, R0 is the approach hydraulic radius, n is the number of rows of
stones, b = D/B and d = Dr/D50. The applicability of the equation is limited to
rough turbulent approach flow, 2 < d < 50, 1 < r < 3, 1  n  10,
0.05 < b < 0.25 and 0.60 < V0/Vc < 1.2 with Vc calculated for the approach flow
through Eq. (9.63).
Froehlich (2013) developed an equation for riprap size by defining the factor of
safety against overturning of a stone in terms of the Shields criterion and relating
the velocity adjacent to the pier to the approach velocity through a potential flow
analysis. For a horizontal riprap cover the factor of safety, defined by the ratio of
static resisting and overturning moments on a single stone, is given by

1
fs ¼ ð9:65Þ
g0 þ 1
Ss

in which η0 is the stone stability number representing the ratio of applied to critical
shear stress multiplied by the relative difference in the densities of the riprap and
water, and is given (in terms of dimensionless shear stresses) by
354 9 Scour and Scour Protection

 
1 s
g0 ¼ 1  ð9:66Þ
Ss sc

The applied shear stress can be expressed through Manning’s equation, with
n = (kn/g0.5)D1/6
r , as
 1=3
2 Dr
s ¼ q kn V2
y

where s, y and V are the local shear stress, flow depth and depth-averaged velocity.
Froehlich suggests a value for kn of 0.145. Substituting s/(qgDr(Ss–1)) in Eq. (9.66)
gives
 
1 kn2 y 2=3 2
g0 ¼ Fr ð9:67Þ
Ss sc Dr

Equating Eqs. (9.66) and (9.67) yields an expression for the riprap stone size,
0 13=2
Dr @ kn2
¼  A Fr3 ð9:68Þ
y sc Ss  1fs

The local Froude number can be expressed in terms of the approach value (Fr0)
through a relationship between the corresponding velocities. Using potential flow
theory, Froehlich showed that the maximum average velocity with a distance eDr
from the pier surface, measured normal to the approach flow direction, is given by
 
D=ðeDr Þ þ 1
V ¼ V0 ð9:69Þ
D=ð2eDr Þ þ 1

The parameter e is a multiplier of the order of unity to define the distance from
the pier to be considered, and needs to be specified. Substituting this velocity in Fr
in Eq. (9.68) and assuming the local flow depth to be the same as the approach
depth give
0 13=2
 
Dr @ kn2 A D=ðeDr Þ þ 1 3 3
¼   Fr0 ð9:70Þ
y0 sc Ss  1 D=ð2eDr Þ þ 1
fs

Froehlich presented a more complete expression for the bracketed term in


Eq. (9.68) for riprap heaped at an angle against the pier. He also provided
adjustments to account for the effect of pier width, cross-flow shear in the approach
flow, multiple equally spaced piers, different pier shapes and the alignment of the
approach flow for non-circular piers. Application of Eq. (9.70) requires
9.5 Scour Around Bridge Piers 355

specification of the safety factor, which depends on the confidence of the designer
in the calculation of the hydraulic conditions, particularly y0 and V0. Froehlich
recommends values of 1.25 for fs and 0.060 for s*c.
Winnowing failure by erosion of the underlying material can lead to the
development of a scour hole beneath the riprap layer, into which the entire intact
layer sinks (Chiew 1995). Winnowing can be prevented by providing more than
one layer of riprap. Recommendations for layer thickness (t) include t > 2Dr50 by
Neill (1973) and by Lauchlan and Melville (2001), and t > 3Dr50 by Richardson
et al. (1991).
Edge failure can be countered by providing sufficient thickness and extent of
riprap cover. A thick layer is able to maintain an armoured surface as the adjacent
fine material is eroded. The further the riprap extends beyond the pier, the less
pronounced is the scouring effect produced by the pier on the adjacent bed. Neill
(1973) recommended an extent of 1.5D in all directions from the face of the pier;
Richardson et al. (1991) recommended a lateral extent of 2D from the face of the
pier, but made no recommendation for a longitudinal extent. Lauchlan and Melville
(2001) recommend a total extent of 4D around the pier. Unger and Hager’s (2006)
Eq. (9.64) implies a relationship between the effective size and number of rows of
stones.
For live-bed conditions, an additional mode of riprap failure is recognized,
whereby the layer is destabilized by the passage of bed forms past the pier (Chiew
and Lim 2000; Lauchlan and Melville 2001). If the trough level of the bed forms is
deeper than the bottom of the riprap layer, the stones are undercut and settle into the
trough and become embedded. The destabilizing effect of the bed forms also
interacts with the modes of failure recognized for clear-water scour, but remains the
dominant failure mode.
Riprap destabilized by bed forms settles as a layer to the level of the bed-form
troughs, and remains capable of providing protection. Lauchlan and Melville (2001)
found that the effectiveness of a riprap layer against live-bed scour could be
increased by initially placing the layer some distance below the original bed sur-
face. They proposed an equation for the required riprap size, which accounts for this
effect, i.e.

Dr50
¼ KD KS Ka Kc 0:3 Fr1:2
0 ð9:71Þ
y0

Equation (9.71) implicitly assumes a riprap layer with a thickness of 2Dr50 and
an extent of 4D. KD, KS and Ka account for the pier diameter relative to the bed
material size, the pier shape and alignment, respectively, which were all assumed to
have values of 1.0. Assuming Ks = 1.0 and Ka = 1.0 would be appropriate for
circular piers and, following Chiew (1995), KD = 1.0 would be valid for D/
Dr50 > *50. The factor Kc accounts for the placement depth of the riprap layer, for
which failure of the layer is assumed to correspond to a riprap scour depth not
exceeding 20% of the maximum unprotected scour depth. From their experimental
data, Lauchlan and Melville propose
356 9 Scour and Scour Protection

 2:75
Y
Kc ¼ 1 ð9:72Þ
y0

in which Y is the placement level of the top of the riprap layer below the initial
sediment surface.
Comparison with other methods showed Eqs. (9.71) and (9.72) to be conservative,
predicting considerably larger riprap sizes, especially for Fr0 less than about 0.5.
Example 9.4
Determine the size of riprap required to protect the riverbed around a 2.5 m diameter
circular bridge pier for an approach flow with a depth of 2.0 m and a velocity of 1.8 m/s
using

a. the method of Chiew (1995),


b. the method of Froehlich (2013) and
c. the method of Lauchlan and Melville (2001) for different riprap depths.

Solution

a. The riprap size is given by Eq. (9.62), i.e.


    3   
3
V03 K DDr50 K yD0
Dr50 ¼ pffiffiffiffiffi
10:47 y0

(y0/D) = 2.0/2.5 = 0.80 < 3, so K(y0/D) is given by Eq. (9.58), i.e.


y  y 0:322
0 0
K ¼ 0:783 0:106
D D
 0:322
2:0
¼ 0:783 0:106 ¼ 0:623
2:50

K(D/Dr50) = 1.0 for D/Dr50 > 50, i.e. Dr50 < 2.5/50 = 0.050 m, which is unlikely.
Therefore, Dr50 must be found by trial, with K(D/Dr50) given by Eq. (9.56), i.e.
     
2
D D D
K ¼ 0:398 ln  0:034 ln
Dr50 Dr50 Dr50

Calculations only with the final trial value are shown.

Try Dr50 = 0.084 m.


Then
     

D 2:50 2:50 2
K ¼ 0:398 ln  0:034 ln
Dr50 0:084 0:084
¼ 0:959
9.5 Scour Around Bridge Piers 357

and
1:803  0:9593  0:6233
Dr50 ¼ pffiffiffiffiffiffiffi
10:47 2:0
¼ 0:084 m

which is equal to the trial value.

b. The riprap size is given by Eq. (9.70), i.e.


0 13=2
 
Dr @ kn2 D=ðeDr Þ þ 1 3 3
¼  A Fr0
y0 sc Ss  1 D=ð2eDr Þ þ 1
fs

with
kn ¼ 0:145
sc ¼ 0:06
Ss ¼ 2:65
fs ¼ 1:25
e ¼ 1:0
V0 1:80
Fr0 ¼ pffiffiffiffiffiffiffi ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ 0:407
gy0 9:8  2:0

Solution for Dr50 is by trial; calculations only with the final trial value are shown.

Try Dr50 = 0.158 m.


Then
!3=2  
Dr 0:1452 2:5=ð1:0  0:158Þ þ 1 3
¼   0:4073
2:0 1:25  1
0:060 2:65 2:5=ð2  1:0  0:158Þ þ 1

giving Dr = 0.158 m, which is equal to the trial value.


If fs = 1.0, Dr = 0.095 m.
If fs = 1.25 and e = 2.0, Dr = 0.141 m,
indicating quite low sensitivity to specification of e.
c. The riprap size is given by Eq. (9.71), i.e.

Dr50
¼ KD KS Ka Kc 0:3Fr1:2
0
y0

with
KD ¼ KS ¼ Ka ¼ 1:0
 
Y 2:75
Kc ¼ 1  ðequationð9:72ÞÞ
y0
358 9 Scour and Scour Protection

For Y = 0
Kc ¼ 1:0

So
Dr50
¼ 1:0  1:0  1:0  1:0  0:3  0:4071:2
2:0

giving

Dr50 = 0.204 m

For Y = 0.30 m
 2:75
Y
Kc ¼ 1
y0
 
0:30 2:75
¼ 1 ¼ 0:64
2:0

So
Dr50
¼ 1:0  1:0  1:0  0:64  0:3  0:4071:2
2:0

giving
Dr50 ¼ 0:131 m

For Y = 0.60 m
 2:75
Y
Kc ¼ 1
y0
 2:75
0:60
¼ 1 ¼ 0:375
2:0

So
Dr50
¼ 1:0  1:0  1:0  0:375  0:3  0:4071:2
2:0

giving
Dr50 ¼ 0:077 m

Scour Prevention by Flow Modification

Bridge-pier scour results primarily from the action of the downflow jet and
horseshoe vortices and measures to reduce their intensity are effective in reducing
the extent of scour. Various flow-altering countermeasures used for modifying the
flow pattern have been reviewed by Tafarojnoruz et al. (2010) and classified into
four distinct categories as summarized below.
9.5 Scour Around Bridge Piers 359

Openings through the piers reduce the strength of the downflow and horseshoe
vortex by allowing a portion of the approach flow to pass through openings within
the pier or pier group. Internal connecting tubes have entry openings at the front of
the pier and exits at locations further around the perimeter and, depending on their
geometry and the approach flow conditions, have been found to reduce the scour
depth up to 39%. A vertical slot in the pier can divert more of the flow than internal
tubes, and therefore more effectively reduce the strength of the downflow jet and up
to 88% scour reduction has been recorded. Significant flow diversion can be
achieved by replacing a large pier with a group of smaller piers.
Devices attached to the pier surface can also reduce the strength of the scouring
flow. A particularly simple measure is the threading of cables spirally around a pier.
Depending on the cable number, size and spiral angle, scour reductions of up to
46% have been achieved. A collar is a thin disc around a pier, which may be placed
singly at different levels or in multiples over the anticipated flow depth. Thin
horizontal plates may also be attached to the upstream face of the pier; various
shapes and configurations have been used and meaningful reductions in scour depth
obtained.
Bed attachments of various types are used to divert or weaken the approach flow,
and consequently reduce the strength of the downflow and horseshoe vortex.
Different forms of attachments include sacrificial piles, vanes and sills, and surface
guide panels.
The fourth category of countermeasures described by Tafarojnoruz et al.
(2010) includes modifications to the pier shape, extraction of water through holes in
the pier by a suction pump and judicious placement of the pier footing. These and
the previously mentioned measures can also be used in combination.
Problems

9:1. A plane bed of: sand is laid in a long, wide channel on a slope of 0.00050.
Determine the flow depth, velocity and unit width discharge at which move-
ment of the bed material will just occur for sand grain sizes of 1.0, 5.0 and
10 mm.
9:2 Water is to be released from a small storm-water retention pond at a discharge
of 0.90 m3/s/m and a velocity of 6.0 m/s into a long, wide channel on a slope
of 0.015 and lined with loose 100 mm stones.
a. Show that erosion is unlikely once uniform flow is attained, but likely near
the reservoir outlet.
It is proposed to provide a rigid lining near the reservoir outlet where scour
could occur. Estimate the distance from the outlet to where loose placing could
commence without scour occurring
b. if the rigid lining is made by setting the 100 mm stones in concrete and
c. if the rigid lining is plain concrete.
Gradually varied flow calculations should be done numerically using three
computational steps.
360 9 Scour and Scour Protection

Note: Manning’s n for the stones set in concrete is 0.030. For loose stone,
Manning’s n can be calculated from the Strickler relationship,
D1=6
n¼ pffiffiffi
6:7 g

where D is the stone size in metres.


9:3 A trapezoidal channel with a bottom width of 3.0 m and side slopes of 3H:1V
is excavated on a slope of 0.0020 in alluvial material with a representative
size, D, of 20 mm and an angle of repose of 35o. Manning’s n can be found
from the Strickler relationship,
D1=6
n¼ pffiffiffi
6:7 g

in which D is the representative particle size (m). Calculate the maximum


discharge for which there will be no erosion in the channel. State clearly any
assumptions made in your calculations.
9:4 A canal is required to convey 20 m3/s of water on a slope of 0.0020 through
well-rounded alluvial material with D50 = 25 mm. Recommend a suitable
cross section after considering
a. a concrete-lined canal,
b. an unlined canal with stabilized banks, with no erosion of the bed and
c. an unlined trapezoidal canal, with no erosion.

9:5 A 2.5 m circular bridge pier is found in river sand with D50 = 0.50 mm. For a
steady approach flow depth of 3.8 m with a velocity of 2.0 m/s determine
a. the equilibrium scour depth and
b. the scour depth after 8 h.
Analyse the sensitivity of equilibrium scour depth with respect to the pier
diameter, sand size and estimation of approach flow conditions.

References

Armitage, N., & Rooseboom, A. (2010). The link between movability number and incipient
motion in river sediments. Water SA, 36(1), 89–96.
Arneson, L. A., Zevenbergen, L. W., Lagasse, P. F., & Clopper, P. E. (2012). Evaluating Scour at
Bridges (5th ed. Rep. No. FHWA-HIF-12-003, HEC-18). Washington, DC: Federal Highway
Administration.
Brown, S. A., & Clyde, E. S. (1988). Design of Riprap Revetment (Rep. No. FHWA-IP-89,016,
HEC-11). USA: Federal Highway Administration.
Cardoso, A. H., & Fael, C. M. S. (2010). Time to equilibrium scour at vertical-wall bridge
abutments. Water Management, 163(10), 509–513.
References 361

Chiew, Y.-M. (1995). Mechanics of riprap failure at bridge piers. Journal of Hydraulic
Engineering, 121(9), 635–643.
Chiew, Y.-M., & Lim, F.-H. (2000). Failure behaviour of riprap layer at bridge piers under
live-bed conditions. Journal of Hydraulic Engineering, 126(1), 43–55.
Collins, M. B., & Rigler, J. K. (1982). The use of settling velocity in defining the initiation of
motion of heavy mineral grains, under unidirectional flow. Sedimentology, 29, 419–426.
Dey, S., & Barbhuiya, A. K. (2004). Clear-water scour at abutments. Water Management, 157,
77–97.
Dey, S., & Barbhuiya, A. K. (2005). Time variation of scour at abutments. Journal of Hydraulic
Engineering, 131(1), 11–23.
Dey, S., & Raikar, R. V. (2005). Scour in long contractions. Journal of Hydraulic Engineering,
131(2), 1036–1049.
Einstein, H. A. (1950). The Bed-Load Function for Sediment Transportation in Open Channel
Flows (Technical Bulletin No 1026, U S Department of Agriculture). Washington, D C: Soil
Conservation Service.
Ettema, R., Constantinescu, G., & Melville, B. W. (2017). Flow-field complexity and design
estimation of pier-scour depth: Sixty years since Laursen and Toch. Journal of Hydraulic
Engineering, 143(9), 03117006.
Forcheimer, P. (1924). Hydraulik. Leipzig and Berlin: Teubner, Verlagsgesellschaft.
Froehlich, D. C. (2013). Protecting bridge piers with loose rock riprap. Journal of Applied Water
Engineering and Research, 1(1), 39–57.
Grass, A. J. (1970). Limited instability of fine bed sand. Journal of the Hydraulics Division, ASCE,
96(HY3), 619–632.
Hager, W. H., & Oliveto, G. (2002). Shields’ entrainment criterion in bridge hydraulics. Journal of
Hydraulic Engineering, 128(5), 538–542.
James, C. S. (1988). Use of the Einstein entrainment function for predicting selective entrainment
and deposition of heavy minerals. Water S A, 14(4), 219–228.
James, C. S. (1990). Prediction of entrainment conditions for nonuniform, noncohesive sediments.
Journal of Hydraulic Research, 28(1), 25–41.
Komar, P. D., & Clemens, K. E. (1986). The relationship between a grain’s settling velocity and
threshold of motion under unidirectional currents. Journal of Sedimentary Petrology, 56(2),
258–266.
Lambe, T. W., & Whitman, R. V. (1969). Soil Mechanics. Wiley.
Lane, E. W. (1953). Progress report on studies on the design of stable channels by the Bureau of
Reclamation. In Proceedings of the American Society of Civil Engineers (No 280).
Lane, E. W. (1955). Design of stable channels. Transactions. American Society of Civil Engineers,
120, 1234–1279.
Lauchlan, C. S., & Melville, B. W. (2001). Riprap protection at bridge piers. Journal of Hydraulic
Engineering, 127(5), 412–418.
Lavelle, J. W., & Mofjeld, H. V. (1987). Do critical stresses for incipient motion and erosion really
exist? Journal of the Hydraulic Engineering, 113(3), 370–385.
Leps, T. M. (1973). Flow through rockfill. In R. C. Hirschfeld & S. J. Poulos (Eds.), Embankment
Dam Engineering (pp. 87–108). New York: Wiley.
Mantz, P. A. (1973). Cohesionless, fine graded, flaked sediment discharge by water. Nature,
Physical Science, 246, 14–16.
Mantz, P. A. (1977). Incipient transport of fine grains and flakes by fluids—extended Shields
diagram. Journal of the Hydraulics Division, ASCE, 103(HY6), 601–615.
Melville, B. W. (2014). Scour at various hydraulic structures: sluice gates, submerged bridges, low
weirs. In H. Chanson & L. Toombes (Eds.), Hydraulic Structures and Society—Engineering
Challenges and Extremes 5th IAHR International Symposium on Hydraulic Structures,
Brisbane, Australia.
Melville, B. W., & Chiew, Y.-M. (1999). Time scale for local scour at bridge piers. Journal of
Hydraulic Engineering, 125(1), 56–65.
362 9 Scour and Scour Protection

Melville, B. W., & Lim, S.-Y. (2013). Scour caused by 2D horizontal jets. Journal of Hydraulic
Engineering, 140(2), 149–155.
Miller, M. C., McCave, I. N., & Komar, P. D. (1977). Threshold of sediment motion under
unidirectional currents. Sedimentology, 24, 507–527.
Mohammedpour, R., Ghani, A. A., Zakaria, N. A., & Ali, T. A. M. (2017). Predicting scour at
bridge abutments over time. Water Management, 170(1), 15–30.
Neill, C. R. (1973). Guide to Bridge Hydraulics, Roads and Transportation Association of
Canada. Toronto, ON, Canada: University of Toronto Press.
Neill, C. R., & Yalin, M. S. (1969). Quantitative definition of beginning of movement of coarse
granular bed materials. Journal of the Hydraulics Division, ASCE, 95(HY1), 585–587.
Oliveto, G., & Hager, W. H. (2002). Temporal evolution of clear-water pier and abutment scour.
Journal of Hydraulic Engineering, 128(9), 811–820.
Oliveto, G., & Hager, W. H. (2005). Further results to time-dependent local scour at bridge
elements. Journal of Hydraulic Engineering, 131(2), 97–105.
Richardson, E. V., Harrison, L. J., & Davis, S. R. (1991). Evaluating Scour at Bridges
(Rep. No. FHWA-IP-90-017 HEC-18). Washington, DC: Federal Highway Administration.
Shan, H., Kilgore, R., Shen, J., & Kerenji, K. (2016). Updating HEC-18 Pier Scour Equations for
Noncohesive Soils (Rep. No. FHWA-HRT-16-045). McLeann, VA: Federal Highway
Administration.
Sheppard, M., Demir, H., & Melville, B. W. (2011). Scour at Wide Piers and Long Skewed Piers.
NCHRP Rep. 682, Washington, DC: National Cooperative Highway Program.
Sheppard, M., Melville, B., & Demir, H. (2014). Evaluation of existing equations for local scour at
bridge piers. Journal of Hydraulic Engineers, 140(1), 14–23.
Shields, A. (1936). Application of similarity principles and turbulence research to bed load
movement, Mitteilungen der Preussischen Versuchsanstalt fürWasserbau und Schiffbau, Heft
26, Berlin [English translation by W. P. Ott and J. C. Uchelon, California Institute of
Technology, Pasadena, California, Report No 167].
Shvidchenko, A. B., & Pender, G. (2000). Flume study of the effect of relative depth on the
incipient motion of coarse uniform sediments. Water Resources Research, 36(2), 619–628.
Tafarojnoruz, A., Gaudio, R. & Dey, S. (2010). Flow-altering countermeasures against scour at
bridge piers: a review. Journal of Hydraulic Research, 48(4), 441–452.
Unger, J., & Hager, W. H. (2006). Riprap failure at circular bridge piers. Journal of Hydraulic
Engineering, 132(4), 354–362.
Vanoni, V. A. (Ed.) (1975). Sedimentation Engineering, American Society of Civil Engineers.
White, S. J. (1970). Plane bed thresholds of fine grained sediments. Nature, 228, 152–153.
Yalin, M. S. (1977). Mechanics of Sediment Transport (2nd ed.). Pergamon.
Yalin, M. S., & Karahan, E. (1979). Inception of sediment transport. Journal of the Hydraulics
Division, ASCE, 105(HY11), 1433–1443.
Yang, C. T. (1973). Incipient motion and sediment transport. Journal of the Hydraulics Division,
ASCE, 99(HY10), 1679–1704.
Yanmaz, A. M., & Kose, O. (2009). A semi-empirical model for clear-water scour evolution at
bridge abutments. Journal of Hydraulic Research, 47(1), 110–118.

Further Reading

Kothyari, U. C. (2008). Bridge scour: Status and research challenges. ISH Journal of Hydraulic
Engineering, 14(1), 1–27.
Kothyari, U. C., & Kumar, A. (2010). Temporal variation of scour around circular bridge piers.
ISH Journal of Hydraulic Engineering, 16(Sup1), 35–48.
Lim, F.-H., & Chiew, Y.-M. (2001). Parametric study of riprap failure around bridge piers. Journal
of Hydraulic Research, 39(1), 61–72.
Further Reading 363

May, R., Ackers, J., & Kirby, A. (2002). Manual on Scour at Bridges and Other Hydraulic
Structures. London: Construction Industry Research and Information Association (CIRIA).
Melville, B. W., & Coleman, S. E. (2000). Bridge Scour. Highlands Ranch, Colorado: Water
Resources Publications.
Yanmaz, A. M. (2006). Temporal variation of clear water scour at cylindrical bridge piers.
Canadian Journal of Civil Engineering, 33(8), 1098–1102.
Postscript

Going Wider and Deeper


For the aspirant expert this book is just a start. Hydraulic structures is a wide field
and the content presented is necessarily selective in terms of both scope and detail.
The emphasis has been on presenting an exposure to some of the more common
hydraulic structures and developing an understanding of the basic concepts
underlying their performance. For practical applications it may be necessary to
obtain further details from the primary references cited or from other sources,
including those recommended for further reading at the end of each chapter. Some
general suggestions for sourcing further information are given here.
A deeper understanding of the fundamental theory can be obtained from any of
the fine books available on open channel hydraulics. These include
Castro-Orgaz, O & Hager WH (2019) Shallow Water Hydraulics, Springer
International Publishing.
Chanson, H (1999) The Hydraulics of Open Channel Flow: An Introduction,
Elsevier.
Chaudhry, MH (1993) Open Channel Flow, Prentice-Hall.
Chow, VT (1959) Open-Channel Hydraulics, McGraw-Hill.
French, RH (1985) Open-Channel Hydraulics, McGraw-Hill.
Henderson, FM (1966) Open Channel Flow, Macmillan.
Many books provide further details for practical applications. In addition to the
topic-specific books listed for further reading, the following more general books are
useful.
Davis, C V and Sorensen, K E (1969) Handbook of Applied Hydraulics, 3rd
Edition, McGraw-Hill.
Novak, P, Moffat, A I B, Nalluri, C and Narayanan, R (2001) Hydraulic
Structures, 3rd Edition, Spon.
United States Bureau of Reclamation (1987) Design of Small Dams, 3rd
Edition.

© Springer Nature Switzerland AG 2020 365


C S James, Hydraulic Structures,
https://doi.org/10.1007/978-3-030-34086-5
366 Postscript

United States Bureau of Reclamation (1978) Design of Small Canal Structures.


Design standards, manuals and guidelines covering a wider range of topics as
well as greater detail are available from various government agencies and research
organizations, many providing free downloads. Some particularly valuable sources
are
CIRIA (Construction Industry Research and Information Association).
United States Bureau of Reclamation.
United States Army Corps of Engineers.
Federal Highway Administration.
The field of hydraulic structures is continually evolving, in terms of both its
scope and its technology. The growing demand for sustainable and environmentally
acceptable solutions in infrastructure developments requires innovative design of
structures. Examples include the inclusion of fish passes in weirs and culverts,
detention and retention storage facilities in sustainable urban drainage schemes, and
low and high discharge dam outlets for environmental and channel maintenance
flow releases. Where standard procedures are unavailable it is necessary to resort to
fundamental principles in order to produce novel designs, and engagement with
experts in other disciplines such as ecology, biology, planning and landscape
architecture may be required. Continually improving instrumentation enables more
and better data to be acquired for confirming and developing design concepts and
details. Advances in analysis techniques, especially in numerical modelling capa-
bilities, allow greater reliability and application to wider ranges of conditions in
design. And, of course, there will be unexpected breakthroughs that could have
major impacts on practice. The expert practitioner must therefore keep abreast of
developments, in hydraulics as well as in allied fields. New contributions are
generally published initially in conference proceedings and academic journals;
some of the more relevant and valuable journals include the following.
Ecological Engineering.
ISH Journal of Hydraulic Engineering (Indian Society for Hydraulics).
International Journal of River Basin Management.
Journal of Hydraulic Engineering (American Society of Civil Engineers).
Journal of Hydraulic Research (International Association for Hydro-
Environment Engineering and Research).
Journal of Irrigation and Drainage Engineering (American Society of Civil
Engineers).
Urban Water Journal.
Water Management (Proceedings of the Institution of Civil Engineering).
A wealth of knowledge is there for discovering!
Index

A Cipolletti weir, 252


Aeration, 117, 134, 145, 149, 150, 153, 154, Circular culvert, 173, 174
156, 157, 159, 162–164 Colebrook–White equation, 18, 20
Air concentration, 153, 157–161, 163, 166 Conjugate depth, 33
Alternate depth, 29, 57 Conservation of energy, 8, 33
Angle of repose, 331, 333, 334, 337, 360 Conservation of mass, 7
Conservation of momentum, 12
B Continuity, 7, 14, 17, 21, 23, 27, 29, 43, 47, 48
Baffle blocks, 185, 196, 197, 209, 210, 224, Contracted weirs, 244, 248, 252
238 Contraction, 75, 77, 78, 81, 92, 94–97
Baffled spillways, 229, 230 Contraction coefficient, 63, 67
Bank stability, 333 Control, 1, 12, 35–39, 41–47, 49, 50, 54, 55,
Bed roughness, 187, 188, 190 57, 105–109, 114, 115, 117, 120, 121,
Bellmouth, 311–315 123, 124, 130, 131, 134, 136–138, 140,
Bends, 286, 290, 291, 293, 296, 309 144, 145, 150, 154, 160, 166
Bernoulli equation, 8–10, 12 Controlled hydraulic jumps, 195
Blasius equation, 19 Conveyance structures, 105, 106
Boundary layer, 115–117, 139, 140, 143, 152, Coriolis coefficient, 9
154–157 Crest gates, 113, 149
Boundary shear stress, 1, 5, 6, 15 Critical flow, 11, 30, 32, 33, 35, 45–48, 51, 52,
Bridge piers, 319, 341, 342, 348, 356, 360 58
Broad-crested weirs, 253, 254, 277 Critical shear stress, 323, 324, 327, 330, 331,
Bucket-type dissipators, 225, 226 333–335, 337, 353
Bulked flow depth, 154, 157–161 Critical velocity, 324, 336, 351
Crump weir, 254, 255
C Culverts, 169–179, 181
Cavitation, 111–113, 115, 137, 140–145, Curvilinear transitions, 87, 90, 92, 103
150–154, 162–166 Cutthroat flume, 265, 269–272, 280, 281
Cavitation index, 151, 153
Chézy equation, 17, 18, 25 D
Chutes, 108, 120, 121, 123, 124, 144–146, Darcy–Weisbach equation, 18, 23, 25, 26
148–157, 159, 160, 163, 164, 166 Design flood, 108, 131

© Springer Nature Switzerland AG 2020 367


C S James, Hydraulic Structures,
https://doi.org/10.1007/978-3-030-34086-5
368 Index

Diffusion analogy, 286 Inlet control, 170, 173, 174, 178, 179, 181
Direct Step Method, 54–56 Intakes, 283–286, 291–297, 299, 300, 307–316
Discharge coefficient, 111, 112, 114, 117, 121, Intake submergence, 313
128, 129, 133, 138, 139, 165, 247, 249, Intensity of motion, 326
252, 254, 256, 260, 269 Interference waves, 144, 145
Dual stable states, 97
L
E Labyrinth, 117, 118
Empirical approach, 321, 326 Laminar flow, 4, 9, 13, 18, 19
Energy dissipation, 183–185, 196, 197, 205, Layout, 296, 309, 313
206, 214, 217, 219, 227, 228, 234, 238 Long-throated structures, 265, 277
Energy loss, 184, 186, 188, 201, 202, 204
Entrance loss coefficient, 176 M
Errors, 243, 252, 277–279 Manning equation, 23, 25–28, 43, 48, 52
Expansion, 75, 77–82, 84, 92, 95, 102 Model testing, 315
Momentum coefficient, 12
F Momentum function, 13, 33–35, 43, 58
Filter layers, 339, 340 Moody diagram, 18–20, 25, 26, 28
Flow classification, 2, 4, 6 Movability number, 326
Flow-measuring structures, 277
Flow resistance, 3, 14, 25, 28, 49 N
Flumes, 243, 257, 258, 261–265, 268–272, Nappe flow, 146–150
274, 277, 280, 281 Nikuradse roughness, 6
Force–momentum flux equation, 12, 13, 33 Nonuniform flow, 3, 7, 16, 17, 28, 49, 50
Friction factor, 18, 19, 25, 26
Froude number, 6, 33, 39, 55 O
Outlet control, 169, 174, 178–181
G Overflow spillway, 108–110, 117, 119, 138,
General resistance equation, 15, 50 145, 150
Gradually varied flow, 3, 37, 49, 50, 52, 53, 59
Gradually varied flow computation, 52 P
Gradually varied flow equation, 51, 52 Parshall flume, 262–266, 271, 280
Gradually varied profiles, 42, 49–52 Permissible velocity, 327, 336
Piano key, 117–119
H Pump intakes, 307, 312
Hydraulically rough flow, 5, 20, 23 Pump sumps, 284, 307, 309, 315
Hydraulically smooth flow, 5, 19, 20
Hydraulic jump, 183, 184, 186–188, 190, 191, R
193–198, 200, 201, 205–207, 210, Rapidly varied flow, 3, 4, 28, 42
212–217, 219, 220, 222, 223, 225, 227, Rectangular box culvert, 170, 172
228, 238 Rectangular weirs, 244, 245, 249, 251, 252,
Hydraulic jump characteristics, 184 279, 280
Hydraulic jump length, 188, 194, 205, 220, Reservoir intakes, 283, 284
222, 223, 238 Reynolds number, 4–6, 18, 21
Hydraulic jump on slope, 200 Riprap, 319, 325, 327, 336, 337, 339, 340,
Hydraulic jump roller length, 205 351–357
Hydraulic radius, 4, 16, 50 River intakes, 284, 285
Hysteresis, 97
Hysteretic behaviour, 69–71 S
Scour, 319–321, 334, 341–349, 351, 355, 358,
I 359
Impact-type dissipators, 228 Scour depth, 319, 341–351, 355, 359, 360
Incipient motion, 320–322, 325, 326, 331, 332, Scour evolution, 347
337 Sediment control, 285
Index 369

Sediment distribution, 286 Subcritical flow transitions, 77


Sediment diversion, 292, 297 Submerged analysis, 64
Sediment ejection, 286 Submerged inlet, 174, 179
Sediment exclusion, 290, 292 Sump volume, 312
Self-aeration, 153–155, 160, 163, 166 Supercritical flow, 6, 7, 29, 30, 32–34, 37, 38,
Sequent depth, 33 42, 43, 46, 47, 54, 55
Sequent depth ratio, 186, 187, 197–199, 203, Supercritical flow transitions, 84
205–207, 210, 213, 217 Suppressed weirs, 244, 245, 248
Settling, 284, 286, 288, 292, 301–304, 307, Surface waves, 77, 84–86, 103
317
Settling efficiency, 302–304 T
Shaft spillway, 130–133, 136, 165 Tapered expansion, 79–81
Sharp-crested weirs, 244, 245, 250, 252, 277, Theoretical analysis, 320
279, 280 Throated flumes, 258, 260, 262
Shear Reynolds number, 6 Transition loss coefficient, 78, 80
Shear velocity, 5 Transitions, 75–84, 87, 89–103
Shield diagram, 322, 323, 326, 337, 340 Translatory waves, 144, 145
Shock front, 85, 87–90 Triangular weirs, 251, 279, 280
Side-channel spillway, 119, 120, 123, 127, 144 Turbulent flow, 4–6, 9, 13, 17–19, 23, 25, 26,
Side weir, 126, 128, 129, 165 50
Sills, 186, 196, 197, 205–209, 213, 218–225, 238
Sinuosity adjustment, 334 U
Siphon spillway, 134–137, 139–141, 165 Underflow gates, 62
Skimming flow, 146–150 Uniform flow, 3, 14–16, 21, 27, 36–39, 42, 43,
Specific energy, 10, 29–33, 35, 37, 47–50, 54, 47, 48, 50–52, 54, 55, 58
57, 58 Unsteady flow, 2, 3
Spillway crest shape, 109, 111 Unsubmerged analysis, 61, 67, 71
Spillways, 105, 108–117, 119–121, 123, 124, Unsubmerged inlet, 179
127, 130–141, 144–146, 150, 151, USBR stilling basins, 209, 218, 219
153–156, 158, 163–166
Spillway splitters, 231, 232, 234 V
Standard Step Method, 57 Vena contracta, 63, 65, 67
Standing waves, 84–87, 94, 95, 97 Venturi flumes, 261
St Anthony Falls stilling basin, 224 Vortex, 300, 301, 308, 309, 311, 313–316
Steady flow, 3, 14, 16
Stepped chutes, 146, 149, 150 W
Stepped spillways, 146, 150 Water surface profile, 184, 186, 190, 191, 195,
Steps, 193, 196, 197, 207, 212–217, 231, 238 202, 204, 220, 222, 238
Stilling basins, 185, 186, 188, 190, 196, 197, Wave celerity, 86
207, 209–212, 214, 217–220, 223–225, Wave suppression, 94, 97
227, 228, 238 Wave suppressor, 185, 186, 222
Straight transitions, 87, 97 Weirs, 243–257, 262, 277–280
Subcritical flow, 29–35, 38, 42, 43, 47, 50, 54

Vous aimerez peut-être aussi