0% found this document useful (0 votes)
40 views

Users Manual Edition

Uploaded by

Apurba haldar
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
40 views

Users Manual Edition

Uploaded by

Apurba haldar
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 201

PANJ-AMU RIVER BASIN (AFGHANISTAN)

USER’S MANUAL FOR THE DESIGN OF


A CROSS-REGULATOR AND A HEAD
REGULATOR ON PERMEABLE
FOUNDATIONS USING A MS EXCEL
SPREADSHEET

Prepared by Claude de Patoul PhD, with funding from the European Union

February 2017
(Second edition revised and expanded)
The information and views set out in this Manual are those of the author and do not
necessarily reflect the official opinion of the European Union. Neither the European
Union institutions and bodies nor any person acting on their behalf may be held
responsible for the use which may be made of the information contained therein.
Preamble of the Director General of Water
Affairs Management, Ministry of Energy and
Water

The Panj-Amu River Basin Programme (P-ARBP) is one of the key


programmes for proper management of water resources through the
introduction and implementation of Integrated Water Resources
Management (IWRM).
The objective of the Programme, started in 2004 and carried-out
jointly by the Ministry of Energy and Water (MEW), the European
Union (EU) and Landell Mills (LM), is poverty and unemployment
alleviation as well as food security through water security in the
basin.
The implementation of such a programme requires the
development of staff capacities, including the basin engineers and
technicians regarding survey, design and construction of hydraulic
infrastructure.
The User’s Manual for the Design of a Cross-Regulator and a Head
Regulator and the attached spreadsheet developed by Dr. Claude de
Patoul, P-ARBP Team Leader from LM is a very valuable self contained
manual. The practical approach makes it a useful guide to good
practices for all concerned with the design and operation of hydraulic
structures. It brings consistency and uniformity of approach to the
design of the most common irrigation structures in new and existing
canal systems of the Panj-Amu River Basin and the other river basins
in Afghanistan.
Therefore, I recommend this manual to the engineers working in
the design department of the River Basin Agencies and the Ministry
of Energy and Water and other experts involved in water resources
affairs of river basins of the country. The analysis, method and
process described in this manual should be adopted for the design,
rehabilitation and maintenance of hydraulic structures of irrigations
systems in our country .
I am sincerely grateful to Dr. Claude de Patoul, Design
Engineer/Team Leader of Panj Amu River Basin Programme ( P-ARBP)
for preparing this manual with the financial support given by the EU.
We wish him further success.

Eng. Sultan Mahmoud Mahmoudi


General Director Water Affairs Management
Ministry of Energy & Water
Kabul
‫تقریظ‪:‬‬

‫پروگرام حوزه دریایی پنج‪-‬آمو (‪ ) PARBP‬یکی از برنامه های مهم جهت مدیریت مناسب منابع آب‬
‫از طریق معرفی و تطبیق پروسه مدیریت همه جانبه منابع آبی (‪ )IWRM‬با توجه به روش حوزه‬
‫دریایی در حوزه دریایی پنج آمو به شمار میرود‪.‬‬
‫هدف این پروگرام که مشترکا ً توسط وزارت انرژی و آب ‪ ،‬جامعه اروپا و کمپنی لندل میلز پیش برده‬
‫میشود عبارت از کاهش فقر و بیکاری و مصئونیت غذایی به وسیله مصئونیت آب در حوزه مذکور‬
‫بوده که در سال ‪ 1384‬به فعالیت آغاز نمود‪.‬‬
‫تطبیق چنین پروگرام ایجاب می نماید تا ظرفیت کارکنان بشمول انجنیران و متخصصین فنی حوزه‬
‫مذکور در بخش سروی ‪ ،‬دیزاین و امور ساختمانی تاسیسات آبی ارتقاء یابد‪ .‬به همین سبب‪ ،‬کتاب‬
‫رهنمود دیزاین سربند تنظیم کننده که محاسبات دیزاین در برنامه اکسل ضمیمه آن بوده ‪ ،‬یک رهنمود‬
‫کامل و ارزشمند می باشد‪ .‬روش های که در آن بکار رفته‪ ،‬قابل استفاده و رهنمای خوب‪ ،‬در امور‬
‫دیزاین و امور عملیاتی ساختمان های هایدرولیکی می باشد‪ .‬این کتاب‪ ،‬زمینه روش هماهنگ و‬
‫همسان را در دیزاین ساختمان های معمولی آبیاری در سیستم های موجود و جدید حوزه دریایی پنج‬
‫– آمو و س ایر حوزه های دریایی کشور مهیا می سازد که از طرف داکتر کلود دی پتول ( ‪Dr.‬‬
‫‪ ) Claude de Patoul‬تیم لیدر پروگرام حوزه دریایی پنج آمو از کمپنی لندل میلز‪ ،‬تهیه گردیده‬
‫است‪.‬‬
‫این کتاب دارای اهمیت و ارزش خاص می باشد‪ .‬بنابرین‪ ،‬مطالعه و کاربرد این رهنمود را برای‬
‫انجنیران بخش دیزاین ادارات حوزه دریایی‪ ،‬وزارت انرژی و آب و سایر دست اندرکاران امور‬
‫مدیریت منابع آب در حوزه های دریایی کشور سفارش می نمایم‪ .‬امید است تحلیل ها‪ ،‬روش ها و‬
‫پروسه های که در این رهنمود استفاده شده است در امور دیزاین‪ ،‬احیای مجدد و حفظ و مراقبت سیستم‬
‫های آبیاری کشور در نظر گرفته شود‪.‬‬
‫بنابرین ‪ ،‬از آقای داکتر کلود دی پتول )‪ )Dr. Claude de Patoul‬انجنیر دیزاین و تیم لیدر‬
‫پروگرام حوزه دریایی پنج ‪ -‬آمو بخاطر ت هیه این رهنمود که به کمک مالی اتحادیه اروپا صورت‬
‫گرفته‪ ،‬قدردانی نموده؛ موفقیت هر چه بیشتر شان را در امور مربوطه خواهانم‪.‬‬

‫با احترام‬

‫انجنیر سلطان محمود محمودی‬


‫رئیس عمومی تنظیم امور آب‬
‫وزارت انرژی و آ‬
Foreword

The double purpose of this USER’S MANUAL FOR THE DESIGN


OF A CROSS-REGULATOR AND A HEAD REGULATOR ON
PERMEABLE FOUNDATIONS USING A MS EXCEL SPREADSHEET
is to serve as a guide to good practices for all concerned with the
design and operation of structures. It brings consistency and
uniformity of approach to the design of the most common irrigation
structures in the new and existing canal systems of the Panj-Amu
River Basin and other river basins in Afghanistan.
This Manual presents instructions, standards and procedures for
the selection and design of a cross-regulator and a gated head
regulator. It is fairly self-contained and based on the publications
listed in the references. It does not include the calculation of
structure stability (shear etc.) of the structure which is dealt with
in other publications.
To practically implement the technique, the author developed a
stand-alone MS Excel spreadsheet computer programme that allows
one to design and compute the flow rate through a complete check
structure. This spreadsheet is an interactive computer program that
generates all the necessary data for the design of a cross-regulator,
a gated head regulator and its discharge calibration (vertical slide
gates). A cross-regulator is used for passive control of the water
level in the river or parent canal. The structure is normally applied
for free flow only. The User’s Manual considers only rectangular
cross-sections to facilitate the analysis. In this manner, the author
can develop the design procedure more fully for rectangular cross-
sections, thereby providing a better basis for later investigations by
others regarding other geometric sections (e.g., trapezoidal,
triangular, or circular).
This application also provides simple graphs to help define
dimensions and hydraulic properties of the structures. A worked out
example of the design of a standard irrigation structure of the basin
is presented in the attached MS Excel sheet. Never blindly enter
data in the spreadsheet and accept the result at face value; one may
get incorrect results and not even realise it.
The User’s Manual is not intended to cover sophisticated or
peculiar irrigation structures linked to particular technical problems
and high discharges. In cases such as these, a different approach
may be used with the help of specialists. Furthermore, the designer
should not overlook the use of alternative methods, if seen to be
more appropriate in peculiar situations.
This second revised and expanded edition has taken into account
additional field experience and the remarks received during the
training sessions and classroom courses given to participants and
students. Although this spreadsheet has been extensively tested to
eliminate errors and inaccuracies, the author cannot guarantee its
suitability for any purpose. The river basin and sub-basin agencies or
company’s designers must ensure that safe and adequate designs are
properly prepared when using the User’s Manual. The use of this
material in any manner whatsoever shall only be done with competent
professional assistance. The author provides no expressed or implied
warranty that this material is suitable for any specific purpose or
project and shall not be liable for any damages including but not
limited to direct, indirect, incidental, punitive and consequential
damaged alleged from the use of this Manual.
Any comments which may lead to improve the next edition of this
Manual are welcomed and can be addressed to [email protected].

Claude de Patoul, PhD


Team leader
Panj-Amu River Basin Programme
CdP User’s Manual February 2017 (second edition revised and expanded)

Contents

1. FLOW CLASSIFICATION 2

1.1. INTRODUCTION 2

1.2. JARGON AND CRITERIA CONSIDERED 2


1.2.1. Normal flow 2
1.2.2. Steady and unsteady flow 2
1.2.3. Uniform and non uniform flow 2
1.2.4. Laminar, transitional and turbulent flow 4
1.2.5. Critical, sub and super-critical flow 4

1.3. EQUATIONS 4
1.3.1. Basic equations 4
1.3.2. Flow classification 5

2. HYDRAULIC STRUCTURE FOR FLOW CONTROL AND


DISTRIBUTION 7

3. HYDRAULIC AND STRUCTURAL DESIGN 10

3.1. LONGITUDINAL AND CROSS-SECTION 10

3.2. HYDRAULIC DESIGN 10


3.2.1. Hydraulic design for sub-surface flow 10
3.2.2. Hydraulic design for surface flow 10

3.3. STRUCTURAL DESIGN 10

3.4. DESIGN CONDITIONS FOR GATES 11

3.5. RIVER TRAINING WORKS 12

4. GENERAL FIELD DATA 13

4.1. SITE INVESTIGATION 13

I
CdP User’s Manual February 2017 (second edition revised and expanded)

4.2. MAIN FIELD DATA 13

5. CROSS-REGULATOR (WEIR) 16

5.1. BROAD CRESTED WEIR 16

5.2. WEIR ELEVATION 16


5.2.1. In the river or parent canal cross-regulator structure 16
5.2.2. In the head regulator of the branching canal 17

6. DISCHARGE AND ENERGY OVER A CROSS-REGULATOR


18

6.1. DISCHARGE RATE 18


6.1.1. Flow rate equation 18
6.1.2. Effective discharge coefficient and streamlines at control section
19
6.1.3. Velocity coefficient 20

6.2. UPSTREAM FLOW VELOCITY AND KINETIC ENERGY HEAD 21


6.2.1. Approach flow velocity 21
6.2.2. Kinetic energy head 22
6.2.3. Upstream total energy elevation HE0 22

6.3. DOWNSTREAM FLOW VELOCITY AND KINETIC ENERGY HEAD 23


6.3.1. D/s elevation HE2 23
6.3.2. Kinetic energy head V32/2g 23

6.4. DISCHAGE EVALUATION AND FLOW STATE 23


6.4.1. Discharge measurement station 23
6.4.2. Flow state of gauging weir 24

6.5. FLUMING OF WATERWAY 25

7. MODULAR FLOW AND SUBMERGENCE RATIO 27

7.1. MODULAR FLOW 27

II
CdP User’s Manual February 2017 (second edition revised and expanded)

7.2. MODULAR LIMIT 27

7.3. DISCHARGE RATE CONSIDERATIONS 28

7.4. HEAD LOSS AND TRADEOFFS 29

8. FLOW VELOCITY AND SEDIMENTATION 30

9. ENERGY DISSIPATION 32

9.1. TOTAL AND SPECIFIC ENERGY 32


9.1.1. Total energy 32
9.1.2. Specific energy 32

9.2. HYDRAULIC JUMP 33

9.3. HYDRAULIC JUMP VARIABLES 34

9.4. JUMP FORMATION 37

10. ANALYTICAL DESIGN OF THE HYDRAULIC JUMP 38

10.1. HYDRAULICS OF THE STILLING BASIN 38

10.2. SPECIFIC ENERGY EQUATION 38

10.3. COMPUTATION OF THE CRITICAL AND SEQUENT DEPTHS OF FLOW


38
10.3.1. Determination of critical water depth yc 38
10.3.2. Determination of super-critical water depth y1 39
10.3.3. Determination of the u/s Froude number 40
10.3.4. Relationship between the sequent depths 40

10.4. CHARACTERISATION OF THE HYDRAULIC JUMP 41


10.4.1. Concentration and degradation 41
10.4.2. Base point and surface point elevation of jump formation 41
10.4.3. Hydraulic jump efficiency 42
10.4.4. Hydraulic jump energy head loss 42

III
CdP User’s Manual February 2017 (second edition revised and expanded)

10.4.5. Hydraulic jump height 43


10.4.6. Hydraulic jump length 43
10.4.7. Characteristic curves of the hydraulic jump 44

10.5. STABILITY AND CONTROL OF THE HYDRAULIC JUMP 45


10.5.1. Pattern of jump positions 45
10.5.2. Tailwater consideration 46
10.5.3. Control of the position of the hydraulic jump 48
10.5.4. Glacis slope 50

11. HYDRAULIC JUMP DESIGN WITH DESIGN CHART 51

11.1. DESIGN CHART 51

11.2. ESTIMATION OF THE SPECIFIC ENERGY 51

11.3. SEQUENT DEPTHS ESTIMATION WITH DESIGN CHART 51

11.4. POSITION OF THE POINTS OF JUMP FORMATION 54


11.4.1. Elevation of the base point of jump formation 54
11.4.2. Elevation of the surface point of jump formation 54

11.5. CHARACTERISATION OF THE HYDRAULIC JUMP 54

12. STANDARD ENERGY DISSIPATER STRUCTURES 56

12.1. INTRODUCTION 56

12.2. SELECTION OF THE ENERGY DISSIPATER 56


12.2.1. Froude number 56
12.2.2. Topographical drop 57

12.3. CARACTERISTICS OF THE ENERGY DISSIPATER 57


12.3.1. Classification of energy dissipater 57
12.3.2. Total length of dissipater 57

12.4. STRAIGHT DROP 63


12.4.1. Introduction 63
12.4.2. Aeration of nappe 64

IV
CdP User’s Manual February 2017 (second edition revised and expanded)

12.4.3. Drop number 65


12.4.4. Basin elevation 65
12.4.5. Straight drop with hydraulic jump 65
12.4.6. Straight drop with impact blocks 65

12.5. END SILL 66

13. HYDRAULIC JUMP WATER SURFACE ELEVATION IN


THE BASIN 71

13.1. INTRODUCCION 71

13.2. JUMP WATER SURFACE ELEVATION IN THE SUB-CRITICAL REACH 71

13.3. JUMP WATER SURFACE ELEVATION IN THE SUPER-CRITICAL REACH


72

14. WATER REQUIREMENT AND CONTROL AT HEAD


REGULATOR 74

14.1. WATER SUPPLY 74

14.2. IRRIGATION REQUIREMENT AND DESIGN DISCHARGE 74

14.3. GATE CHAMBER 74


14.3.1. Pier 74
14.3.2. Under-flow gate 74
14.3.3. Approach flow velocity 75

14.4. EFFECTIVE OR CLEAR WATERWAY 75


14.4.1. Pier(s) contraction 75
14.4.2. Pier length and critical depth 77

14.5. GATE OPENING, ENERGY AND MOMENTUM 77

14.6. WATER SURFACE PROFILES WITH GATE 79


14.6.1. Introduction 79
14.6.2. Surface flow profiles encountered with a gate 80

V
CdP User’s Manual February 2017 (second edition revised and expanded)

14.7. FLOW BEHAVIOURS THROUGH A GATE 82


14.7.1. Regulation and flow types 82
14.7.2. Pivot table 82

15. GATE(S) OPENING WITH NON ORIFICE FLOW


CONDITIONS 87

15.1. CASE Nº1: GATE LOWERED INTO A RECTANGULAR HORIZONTAL


CANAL ABOVE FLOW 87
15.1.1. Outflow conditions with h0 < w 87
15.1.2. Outflow discharge equation for non orifice conditions 87
15.1.3. Effective discharge coefficient 89
15.1.4. Contraction coefficient 89
15.1.5. Velocity coefficient 89

16. GATE(S) OPENING WITH ORIFICE FLOW


CONDITIONS 90

16.1. CASE Nº2: GATE(S) LOWERED INTO A RECTANGULAR HORIZONTAL


CANAL TO A HEIGHT BELOW CRITICAL DEPTH WITH FREE FLOW TYPE D/S OF
THE GATE(S) 90
16.1.1. Flow description downstream of the gate(s) 90
16.1.2. Free flow discharge equation 91
16.1.3. Contraction coefficient 91
16.1.4. Discharge coefficient 92
16.1.5. Velocity coefficient 93

16.2. CASE Nº3: GATE(S) LOWERED INTO A RECTANGULAR HORIZONTAL


CANAL WITH SUBMERGED FLOW D/S OF THE GATE(S) 93
16.2.1. Flow description downstream of the gate(s) 93
16.2.2. Modular limit 93
16.2.3. Submerged flow discharge equation 94
16.2.4. Contraction coefficient 95
16.2.5. Discharge coefficient 95
16.2.6. Velocity coefficient 95
16.2.7. Froude number 95

16.3. CASE Nº4: GATE(S) LOWERED INTO A RECTANGULAR HORIZONTAL


CANAL TO A HEIGHT CREATING A TRANSITIONAL TYPE OF FLOW 95
VI
CdP User’s Manual February 2017 (second edition revised and expanded)

16.4. GATE SILL 96

16.5. Forces AND MOMENT on gate(s) 96


16.5.1. Forces on closed gate(s) 96
16.5.2. Forces on closed gate(s) 97

16.6. FLOW RECAPITULATION 98

17. RATING CURVE 101

17.1. Discharge MEASUREMENT AND CONTROL 101


17.1.1. Discharge measurement 101
17.1.2. Discharge control 101

17.2. Discharge rate AND STRUCTURE DESIGN 101

17.3. GATE SIZING AND NUMBER OF GATES 102

17.4. FLOW CONDITIONS 102


17.4.1. Flow conditions 102
17.4.2. Gate opening and position of the hydraulic jump 102

17.5. RATING CURVE 103


17.5.1. Introduction 103
17.5.2. Rating curve for natural channel 103
17.5.3. Rating curve of non orifice flow 104
17.5.4. Rating curve of orifice flow 106
17.5.5. Hydrograph and gate(s) openings 108

17.6. ACCURACY OF THE RATING CURVE 110


17.6.1. Introduction 110
17.6.2. Source of errors 110
17.6.3. Submerged flow type 110

17.7. CONCLUSIONS 111

18. WATERWAY AND REGIME SCOUR DEPTH 114

18.1. WATERWAY DETERMINATION 114

VII
CdP User’s Manual February 2017 (second edition revised and expanded)

18.1.1. Wetted perimeter 114


18.1.2. Looseness factor 114

18.2. SCOUR DEPTH 115

18.3. SOIL CONDUCTIVITY 115

18.4. SILT FACTOR 117

18.5. CALCULATION OF REGIME SCOUR DEPTH IN CHANNEL 117


18.5.1. Mean scour depth 118
18.5.2. Normal scour depth 118
18.5.3. Maximum discharge intensity per unit width of weir 118

19. CUT-OFF WALL 120

19.1. CUT-OFF WALL ROLE 120

19.2. CUT-OFF WALL DEPTH 120

20. FLOW PATH IN PERMEABLE SOILS 124

20.1. SUB-SOIL DATA 124

20.2. STREAM LINES 124

20.3. SUB-SOIL PRESSURE 124

21. PRINCIPAL CAUSES OF INESTABILITY 125

21.1. PRINCIPAL FAILURE OF A STRUCTURE 125


21.1.1. Failure due to sub-surface flow 125
21.1.2. Failure due to surface flow 125

21.2. ADOPTED SOLUTIONS AGAINST FAILURE 126

22. EXIT GRADIENT AND STRUCTURE LENGTH 127

VIII
CdP User’s Manual February 2017 (second edition revised and expanded)

22.1. HYDRAULIC GRADIENT 127

22.2. EQUATION OF THE EXIT GRADIENT 127

22.3. SAFE EXIT GRADIENT 127

22.4. DETERMINATION OF THE EXIT GRADIENT 128


22.4.1. Estimated value of λ 128
22.4.2. Calculated value of λ 128

22.5. EXIT GRADIENT CONTROL 129


22.5.1. Parameters of control 129
22.5.2. Length of the solid floor 129
22.5.3. Depth of cut-off wall 129

23. UPLIFT SUB-SOIL PRESSURE AND FLOOR THICKNESS


130

23.1. FLOOR THICKNESS 130

23.2. STANDARD PROFILES 130

23.3. UPLIFT PRESSURE 131


23.3.1. Key points 131
23.3.2. Upstream pile 132
23.3.3. Downstream pile 132

23.4. CORRECTIONS OF UPLIFT PRESSURE 132


23.4.1. Correction for mutual interference of cut-off walls 132
23.4.2. Correction for the floor thickness 133
23.4.3. Correction for floor slope 134
23.4.4. Correction in the jump trough 134

23.5. HEADS AND GRADIENTS 136

24. UPSTREAM AND DOWNSTREAM PROTECTIVE


WORKS 138

24.1. LOCATION 138


IX
CdP User’s Manual February 2017 (second edition revised and expanded)

24.2. TYPE OF PROTECTION WORKS 138

24.3. INVERTED GRANULAR FILTER 138


24.3.1. Description 138
24.3.2. Concrete blocks revetment 140
24.3.3. Riprap revetment 141
24.3.4. Stone piching revetment 141
24.3.5. Boulders revetment 142

24.4. LAUNCHING APRON 143

24.5. LENGTH OF PROTECTIVE WORKS 143

24.6. ROCK SIZING EQUATIONS FOR REVETMENT 143

25. TRANSITION STRUCTURES AND CHANNEL SIDE


SLOPE 145

25.1. INTRODUCCION 145

25.2. CHANGE OF SECTION AND SPECIFIC ENERGY 145


25.2.1. Change in bed level 145
25.2.2. Change in channel width 147

25.3. TYPES OF TRANSITIONS 149


25.3.1. Transitions and wing walls 149
25.3.2. Head loss 150

25.4. CHANNEL SIDE SLOPE 151

25.5. STRUCTURE FREEBOARD 151

26. DESIGN PRINCIPLES OF GRAVITY WALLS 153

26.1. DESIGN PRINCIPLES 153

26.2. PRESSURE 153

26.3. VERTICAL PRESSURE FORCES 154


X
CdP User’s Manual February 2017 (second edition revised and expanded)

26.4. LATERAL PRESSURE FORCES 154


26.4.1. Active pressure forces 154
26.4.2. Passive pressure forces 155

26.5. BASIC INSTABILITY MODES 155

26.6. SLIDING 155

26.7. OVERTURNING 156


26.7.1. Overturning moments 156
26.7.2. Stabilizing or resistance moments 156

26.8. SOIL BEARING 157

26.9. DRAINAGE 158

26.10. SAFETY FACTORS 158

LIST OF PRINCIPAL SYMBOLS 171

GLOSSARY 175

REFERENCES 181

XI
CdP User’s Manual February 2017 (second edition revised and expanded)

XII
PART I:

WATER FLOW AND DESIGN APPROACH

1
CdP User’s Manual February 2017 (second edition revised and expanded)

1. FLOW CLASSIFICATION
1.1. INTRODUCTION

In the User’s Manual, a channel is a natural conduit with a free surface open
to the atmosphere and a canal is a water human-made or artificial conduit with
a free surface open to the atmosphere.
However, in engineering practice, the words are interchangeable.
Depending on the circumstances, they also receives different name such as
channel, chute, aqueduct etc. These names are used fairly loosely and can be
defined only in a very general manner.
1.2. JARGON AND CRITERIA CONSIDERED

The flows are classified according to the variation in the parameters of flow
in relation to space and time. The state of flow is classified according to the
range of the invariants of flow in relation to (dynamic) viscosity (friction force
which exists inside a fluid as it flows) and gravity in the direction of flow.
1.2.1. Normal flow
The normal flow is the flow where it wants to be with friction forces exactly
balancing gravitational acceleration.
1.2.2. Steady and unsteady flow
The criterion considered is time. The flow is said to be steady if the fluid
properties (depth of flow, velocity, density and discharge) don’t change during
the time interval under consideration. The flow is unsteady if one of the
properties (the depth of flow etc.) changes with time.
1.2.3. Uniform and non uniform flow
The criterion considered is space (or depth) (δy/δx = 0). The flow is said to
be uniform if the depth of flow is the
same at every section of the canal or
channel1. In other words, there is a
balance between the frictional loss and
drop in elevation of the channel: the
friction slope Sf is equal to the bottom
Figure 1.1: uniform flow slope S0 when the head loss hL is equal
to the elevation drop, that is Sf = S0
when hL = Z1 – Z2.
For a given roughness coefficient (resistance to flow), discharge and slope,
there is only one possible depth for maintaining a uniform flow (empirical
Manning´s equation). This flow depth is the normal depth which corresponds

1 A uniform flow may be steady or unsteady, depending on whether or not the depth changes with time.

2
CdP User’s Manual February 2017 (second edition revised and expanded)

to the minimum energy level. The water will always to try to flow at normal
depth.
The basic requirement of a non uniform flow is the slope of the water
surface and the slope of the bed are different.

Figure 1.2: non uniform flow

 Gradually varied flow (δy/δx < 1): uniform flow whose water depth
varies gradually along a long distance in space with stream wise
distance because of an imbalance between gravitational forces and
friction forces. This may occur as the result of a change in channel
conditions (slope, cross-section or roughness) or as an adjustment
brought about by u/s or d/s disturbances1. Because the variation is
gradual, the flow can still be treated as one-dimensional (varying only
with space) and the pressure2 as hydrostatic Furthermore, the u/s and
d/s specific energy are still assumed as equal.
 Rapidly varied flow (δy/δx > 1): the rapidly varied flow has a very
pronounced curvature of the streamlines and all the parameters are
modified in a very short distance. The change of curvature may be so
pronounced that the flow profile is virtually broken (jump). The u/s
and d/s specific energy are no more assumed as equal (E1 ≠ E2) but the
u/s and d/s specific force are the same (M1 = M2) (refer to figures 9.1a
and 9.1b).

1 Backwater curve if depth of flow increases in the direction of flow; drawdown curve if decreases in the direction
of flow.
2
Engineers often refer to pressure in terms of metres of water rather than as a pressure in kN/m 2. They can do
this because of the unique relationship between pressure p and water depth h (p=ρ*g*h). It is called the pressure
head or just head and is measured in metres using the pressure-head equation.

3
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 1.3: non uniform and uniform flow forces

1.2.4. Laminar, transitional and turbulent flow


The criterion considered is the viscous forces and the gravity relative to the
interne inertia of the flow (Reynolds number)1.
This type of flow rarely exists in nature and so is not of great practical
concern in hydraulics; only turbulent flow is considered in this Manual.
1.2.5. Critical, sub and super-critical flow
The criterion is gravity whose effect is represented by a ratio of inertial
forces to gravitational force and given by the number Fr (refer to article 10.3.3).
A small surface wave occurs when the flow in the open channel is disturbed.
Sub-critical2 flow occurs if the small surface wave can propagate u/s as well as
d/s. The velocity of flow is less than that of the wave propagation. It will lead to
a backwater zone of great distances before the obstacles3.
If the wave can only propagate d/s and cannot propagate against the flow,
this flow is called super-critical4 flow, in which the flow velocity is greater than
that of the wave propagation (wave celerity). The backwater zone occurs only
around the obstacles
If the velocity of the small surface wave propagating d/s is zero, it is just the
critical situation to distinct the super-critical flow and the sub-critical flow. This
flow is called the critical flow.
1.3. EQUATIONS

1.3.1. Basic equations

1 The laminar flow is not considered in this manual due to the low water kinematic viscosity.
2 Sub because yc is below the flow depth considered.
3 Backwater and draw-down curves are formed where flow is influenced by cross-regulators, drop structures, gate

and measuring flumes. The water surface gradually rises when flow is backed-up with at the same time a
diminishing of the flow velocity. A draw-down profile gradually reduces water depth while the flow velocity
gradually increases.
4 Super because y is above the flow depth considered.
c

4
CdP User’s Manual February 2017 (second edition revised and expanded)

The hydraulic theory is based on three basic equations:


 Continuity equation: links u/s and d/s discharges which means also
areas ad velocity.
 Energy equation: links pressure force (head, kinetic and potential) to
velocity.
 Momentum equation: links force (mass) and velocity.
1.3.2. Flow classification
The empirical Manning’s equation is used for free flow in open channels. It
cannot be used where obstacles such as gates, weirs etc. are present.
In order to analyse situations where obstructions are to be considered,
Bernoulli’s energy equation must instead be used in certain conditions1. For
Bernoulli’s energy equation to hold, the streamlines must be parallel (usually
applicable to steady flow). In certain flow phenomena, however, we simply can
no longer ignore the mechanical energy losses due to the canal banks or such
as in a hydraulic jump (where streamlines are curved) and we must look to the
momentum principle as an alternative ways of describing the flow 2
(momentum (kgm/s) = mass (kg) * velocity (m/s)).
Some described flow classifications are shown in the following figures:

1 In the Bernoulli’s energy equation, the solutions depend upon the assumption that the total energy head at the
u/s section should be equal to the total energy head at the d/s section plus the loss of energy between the two
sections: the mechanical energy is conserved (principle of mechanical energy conservation). In fluid mechanics,
it is found convenient to separate mechanical energy from thermal energy and to consider the conversion of
mechanical energy to thermal energy as a result of frictional effects as mechanical energy loss. Then the energy
equation becomes the mechanical energy: the sum of the gravitational (elevation Z), potential and kinetic
energies. If energy actually leaks from the system via frictional head loss, the Bernoulli equation will overstate
the energy available to the flow and the related predictions of velocity and depth will proportionately be in error.
2 The momentum principle is an alternative way of describing the flow (principle of u/s and d/s forces

corresponding to mass flow rate times velocity conservation). The momentum equation is brought to bear on
problems involving high internal energy changes such as the task of describing the hydraulic jump where the flow
is far less than perfectly energy conservative. If the energy equation is applied to such problems, the unknown
internal energy loss ΔE is indeterminate and the omission of this term would result in considerable errors (refer
to article 9.3). Indeed, all real flows in nature dissipate energy in overcoming frictional resistance. If we want to
use Bernoulli’s equation, we minimize this error by considering only short reaches of channel and only gradual
transitions in order to ignore these energy losses. However, the energy equation is similar to the momentum
equation when applied to certain flow phenomena such as the gradually varied flow and the two principles will
produce practically identical results.

5
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 1.4: flows classification downstream a sluice gate

Figure 1.5: flows classification with the presence of a weir

If the sill level of the weir rises sufficiently, then, as the specific energy
cannot be less than the critical specific energy Ec, the u/s flow must back up,
increasing the depth and the energy in the sub-critical flow immediately u/s the
weir, as pictured in the above figure (refer to table 14.3 and figures 25.1a &
25.1b).
Once critical conditions are established over the weir, it can be used as a
flow-metering devises (refer to article 10.3.1).

6
CdP User’s Manual February 2017 (second edition revised and expanded)

2. HYDRAULIC STRUCTURE FOR FLOW CONTROL AND


DISTRIBUTION
In the User’s Manual, the head work consists of two modules: a passive
water level regulator in the parent canal or in the river (usually an ungated
broad crested weir) and a manual gated head regulator in the canal (off-take or
intake).
The head regulator is of the gated breast wall type while the structure is of
the open channel type. Cross-regulator (usually ungated) is of the open channel
type. The head regulator performs two functions at the same time: flow
discharge measurement function and water level control function.

Figure 2.1: plan of head work with cross-regulator and head-regulator (off-
take) modules

Figure 2.2: isometric view of head work with cross-regulator and head
regulator (off-take) modules

7
CdP User’s Manual February 2017 (second edition revised and expanded)

From u/s to d/s, the different parts of the hydraulic structures are as
follows:
 U/s transitional structure to guide the flow from the channel into
the structure.
 Scour protective works u/s of the structure.
 U/s protected transition to guide the flow from the channel into
the structure.
 Water level (head) control section (broad crested weir).
 D/s conveyance structure (sloping glacis).
 Energy dissipation reach (USBR type of stilling basin).
 Scour protective works d/s of the structure.
 D/s transitional structure to guide the flow from the structure
into the channel.
The locating the main head work along a stream has to be carefully planed.
The best location is on the outside of a river bend as shown in the following
figure. If located on the inside of a band, it will be continually silted up as a
result of the actions of secondary flows in the river.

Figure 2.3: view of head work location along a river

8
CdP User’s Manual February 2017 (second edition revised and expanded)

Weir rating curve with broad and / or short crested regime

Depth of water yn downstream of structure (mj


1.80

Depth of water ho upstream of weir crest (m)


1.60

1.40

1.20

1.00

0.80

0.60

0.40

0.20

0.00
0.00 10.00 20.00 30.00 40.00 50.00 60.00 70.00 80.00 90.00
Weir Discharge (m3/s)

Stilling basin
Cross-regulator

Stilling basin Secondary canal

Head regulator
Panj-Amu river basin
1.5 85
Gate opening height w and depth of flow ho (m)

1.4 80
Flow in structure (m3/s)

75
1.3
70
1.2

Main canal
65
1.1
60
1.0
55

0.9
50

0.8 45

0.7 40

35
0.6

30
0.5
25
0.4
20
0.3
15
0.2
10

0.1 5

0.0 0
March 2 March 3 April 1 April 2 April 3 May 1 May 2 May 3 Jun. 1 Jun. 2 Jun. 3 Jul. 1 Jul. 2 Jul. 3 Aug. 1 Aug. 2 Aug. 3 Sep. 1 Sep. 2 Sep. 3
(decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade)
Decadal average river or parent canal depth of flow hₒ (m) with crest = 0.40 m Decade (day)
Decadal gate opening height w (m) with max. crop water requirements (obtained from orifice and non orifice rating curves)
Decadal gate opening height w with decadal calculated crop water requirements (m)

Figure 2.4: bird view of Selaba head work with ungated cross-regulator and rating curve on the right and gated head
regulator with canal hydrograph and corresponding gate(s) openings on the left (Yatim Tepa canal)

9
CdP User’s Manual February 2017 (second edition revised and expanded)

3. HYDRAULIC AND STRUCTURAL DESIGN


3.1. LONGITUDINAL AND CROSS-SECTION

The sections of a cross-regulator structure are given below (refer to figure


14.6):
1. Section A: converging flow.
2. Section C-C: water level (head) control.
3. Section 1-1: super-critical flow.
4. Section 2-2: energy dissipation.
5. Section 3-3: super-critical flow.
In this User’s Manual, we consider the structure with a rectangular cross-
section.
3.2. HYDRAULIC DESIGN

The hydraulic design comprises of fixing the overall dimensions and profiles
of the structure through analytical equations and available empirical formulas 1.
The hydraulic design determines the (rectangular) cross-sections of the
structure, the total mechanical energy of water (expressed as the total head in
metre of water), the permissible velocities, the cut-off wall (refer to glossary)
considered u/s and d/s of the structure, the glacis, the type of stilling basin and
the u/s and d/s protected transition works.
The hydraulic design of the structure involves two sets of hydraulic
conditions for sub-surface flow and for surface flow.
3.2.1. Hydraulic design for sub-surface flow
The sub-surface or seepage flow occurs due to the hydrostatic head Hs
difference from the u/s to the d/s sections of the structure. It includes piping
and the associated sand boil phenomenon.
In this User’s Manual, the effects of the sub-surface flow on the stability of
the hydraulic structure are computed with the Khosla’s empirical method of
independent variables which determines the strength of the up-thrust (uplift)
pressure.
3.2.2. Hydraulic design for surface flow
In certain conditions, the dynamic action of the water flow may create the
formation of the hydraulic jump causing an uplift unbalanced head in the jump
basin which may be larger than the sub-soil seepage.
3.3. STRUCTURAL DESIGN

1
If necessary, the dimensions set by the hydraulic design can be further refined by testing a scale model of the
structure in an appropriate laboratory.

10
CdP User’s Manual February 2017 (second edition revised and expanded)

The structural design is the dimensioning of the super-structures and sub-


structures and the manner in which the load bearing (structural) members of a
hydraulic structure support each other in sharing the load (stress). The various
parts of the structure are analysed for stresses under different loads and
reinforcement1 or other structural details are worked out. The stability analysis
is not carried out in details in this Manual for the type of structures considered
and can be found in most hydraulic books.
In the User’s Manual, larger concrete head regulator (also called discharge
regulator) are divided in 3 parts to allow movement due to settlement 2.
3.4. DESIGN CONDITIONS FOR GATES

In the design of the hydraulic structures, the following criteria have to be


taken into account for the gates in closed position:
 The sliding gates for the head regulators (usually breast wall type) are
designed to withstand and operate against the full unbalanced head,
with the u/s water level above the gate crest level with no water d/s.
 The sliding gates for the structures (open channel type) are designed
to withstand and operate against the full unbalanced head, with the
u/s water level at the top of the gate with no water d/s.
 The cross-regulator sliding gates (open channel type) are designed to
withstand and operate against a head of water 0.30 m above the top
of the gate assuming no water downstream. The gates are ether the
open channel type or breast wall type. Since the open channel type is
subject to overtopping in times of flood they shall be designed, for
safety reasons, as it were of the breast wall type, thus giving the worst
case loading condition.

1
The amount of rebar used in typical structures is a small percentage of the amount of concrete; mostly, as a rule
of thumb, we use about 1% to 0.5 % rebar for carrying the tension forces in bending. Columns may use up to 6%
rebar, partly because the rebar carries both tension and axial forces. Since rebar costs much more than concrete,
efficient engineering design minimizes rebar use. In order for reinforcing bars to be in the required location in
reinforced concrete, the bars must often be fabricated to special shapes. Shop drawings take the schematic
information from the structural drawing and show the actual bar lengths, bends, clearances, etc. For a rough
estimation of the reinforcement steel in construction, the following thumb rules may be adopted:
Construction Unit Quantity
Slab kg/m3 of concrete 50 to 80
Lintel kg/m3 of concrete 80
Beam kg/m3 of concrete 100 to 150
Column kg/m3 of concrete 150 to 225
Footing slab kg/m3 of concrete 80

2An expansion joint of 3 cm between the parts is made watertight by a 3-bulb rubber or PVC water stop cast in
both concrete ends.

11
CdP User’s Manual February 2017 (second edition revised and expanded)

3.5. RIVER TRAINING WORKS

River training works are necessary


to achieve as follows:
 Prevent out flanking of the
structure.
 Prevent u/s land flooding by the
river.
 Provide favorable curvature of
Figure 3.1: braided river flow u/s of the structure
 Guide the river to flow axially
through a structure.
Channel migration may occur naturally or as a result of human activity, and
may be associated with any of the causes that give rise to degradation and
aggradation. Migration of the entire river channel as part of the process of
meander progression, or movement of the deep-water channel within the
same overall channel banks, can affect the scour exposure of a structure whose
foundations may have been set in relation to an earlier channel position. In
some cases, migration may occur rapidly in response to a particular flood event,
but in other cases it may be gradual. In a braided river, the channel positions
are continuously changing, as shown in the opposite picture.
As a general rule, if there is potential for channel migration, the foundations
should be designed or assessed on the basis of any credible shifts of the deep-
water channel or channels. Alternatively, training works may be carried out to
limit the possible movement of the deep-water channel.
At many instances, it is necessary to guide and/or restrict the course of the
river through the hydraulic structure and it is achieved by the use of river
training works.

12
CdP User’s Manual February 2017 (second edition revised and expanded)

4. GENERAL FIELD DATA


4.1. SITE INVESTIGATION

The general layout of the structure is mainly influenced by the topographical


features, the approach conditions of the river or the parent canal and
environmental considerations.
The topographic study of the flood plain allows defining an average cross-
section of the river or parent canal and obtaining several field data essential for
the design of the hydraulic works, such as the bottom slope of the river, the
height of the banks, the type of material from the river bed etc.
The irregular river cross section should be simplified to a representative
trapezoidal section for simplifying calculations as shown in the following figure:

Figure 4.1: irregular river cross section and trapezoidal cross-section

The d/s reach should be investigated if the hydraulic structure causes flood
risks (backwater curve) which would means additional embankments.
4.2. MAIN FIELD DATA

The main field data to be collected from the site for the design of the
structure are listed below.
1. Index map of the site.
2. Contour plan of the area.
3. Command area of canal (ha).
4. Discharge rates of:
a. Flood or maximum discharge of river or parent canal Qmax1.
b. Minimum discharge in the river or parent canal Qmin of river.
1
The flood discharge rate or design volumetric flow rate considered in the river has an annual exceedance
probability (AEP) of 1 in 100.

13
CdP User’s Manual February 2017 (second edition revised and expanded)

c. Irrigation water need or water duty Qcrop. (irrigation water


need) as shown in the following figure:

Figure 4.2: climate to irrigation water need

5. Cross-sections of the river or parent canal and off-taking canal.


6. Side slope of river or parent canal bank (H: V).
7. Upstream investigation for additional flood risk by backflow.
8. Test pits log chart and bed materials (d50 mean diameter).
9. Permeability coefficients and soil mechanical parameters.
10. Rainfall data of the location.
11. Location and accessibility of quarry areas for coarse and fine
aggregates.

14
CdP User’s Manual February 2017 (second edition revised and expanded)

PART II:

HYDRAULICS OF THE CROSS-REGULATOR


STRUCTURE

15
CdP User’s Manual February 2017 (second edition revised and expanded)

5. CROSS-REGULATOR (WEIR)
5.1. BROAD CRESTED WEIR

A cross-regulator or weir is defined as a barrier over which the water flows


in an open channel. The edge or surface over which the water flows is called
the crest or sill. The overflowing sheet of water is the nappe. If the nappe
discharges into the air, the weir has a free discharge. If the discharge is partially
under water, the weir is submerged or drowned.
In the User’s Manual, a broad crested weir with a rectangular control
section is usually adopted for its greater structural stability. The calculation of
the rate of flow relies on the assumption that the fluid will pass through the
critical depth somewhere near to its d/s edge, as shown in the following figure.
Furthermore, the crest is sufficiently long for parallel streamlines to be
established (hence broad crested) to obtain parallel streamlines (refer to figure
6.2) but insufficiently long for significant frictional losses1. Consequently this
makes the calculation of the discharge relatively straightforward.
In the User’s Manual, a standard shape is used, then there is a large body of
literature available relating to their design, operation and coefficient of
discharge according to known relationships 2 (refer to article 6.4).

Figure 5.1: illustration of broad crested weir terminology


5.2. WEIR ELEVATION

5.2.1. In the river or parent canal cross-regulator structure

1
The disadvantage of a short crest is that it does not ensure a constant coefficient of discharge for varying heads,
hence, the fall cannot be used as a meter; but it provides a higher coefficient of discharge and thus a greater
discharging capacity.
2 The ability to calibrate the structure using equations instead of measurements is based on the existence of

parallel streamlines in the control section over the weir crest; nowhere above the short/sharp crest can the
curvature of the streamlines be neglected.

16
CdP User’s Manual February 2017 (second edition revised and expanded)

One of the most important design parameters is the height of the crest
above the channel bed. This height should be sufficient to provide modular flow
for the entire range of discharges that the broad crested weir is intended to
measure. In order to ensure critical flow over the cross-regulator for all flow
conditions, the maximum anticipated flow rate through the channel should be
used to calculate the required cross-regulator height. Furthermore, the crest
height should at least cover the head losses at the intake and in the approach
to the intake.
Afflux and discharge are
related to the crest level. The
operation of cross-regulators
is often misunderstood and it
is believed that they cause the
flow to always back up and so
raise water level u/s as shown
with the presence of a bridge
in the opposite figure. This
Figure 5.2: afflux only happens once critical
conditions are achieved on the
sill (refer to article 25.2). A lower crest level results in lesser afflux, in an
increased height of gates and a decrease in floor thickness. By providing a high
afflux, the width of the structure can be narrowed but the cost of training works
will go up and the risk of failure by out flanking will increase. The crest level is
also function of the required water surface elevation in the branching canal.
5.2.2. In the head regulator of the branching canal
In the User’s Manual, the crest elevation is based on the lower water
discharge rate in the river or parent canal (Qmin) and generally kept 0.1 m to 0.2
m higher than the crest level in the river or parent canal; it is also function of
the size available and number of gates to be provided in the branching canal.
Furthermore, the crest elevation should at least cover the head losses at the
intake and in the approach to the intake.

17
CdP User’s Manual February 2017 (second edition revised and expanded)

6. DISCHARGE AND ENERGY OVER A CROSS-REGULATOR


6.1. DISCHARGE RATE

The main cross-regulators sections and plans are summarized in the below
sections.
Sections Plans

Broad
crested

Sharp
crested

Ogee

Straight
drop

Figure 6.1a: sections and plans of cross-regulators

In the User’s Manual, a rectangular broad crested weir is usually used for
passive control of the water levels in the river or parent canal.
6.1.1. Flow rate equation
The presence of the weir in the open channel introduces an area
constriction to an otherwise uniform flow in a prismatic channel of mild slope
(refer to article 25.2 and table 25.1). The ratio H0/L (refer to article 6.4.2)
indicates if the weir behaves as a broad crested weir or a short crested weir.
The head-discharge equations based on ideal flow must be corrected for
energy losses, velocity distributions and streamline curvature by the
introduction of a discharge coefficient Ce. The resulting head-discharge
equation for a rectangular channel is as follows:

Q = Ce * Bweir * H03/2 (6.1a)


Where:
Q = flow (discharge) rate (m3/s).
Ce = effective discharge coefficient or weir coefficient.
Refer to list of symbol.

18
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 6.1b: illustration of broad crested weir terminology

In an open channel, it is not possible to measure the energy head H0 directly


and it is therefore common practice to relate the flow rate to the u/s sill or crest
referenced water head h0 by introducing the velocity coefficient Cv.

Q = Ce * Bweir * h03/2 (6.1b)


Where:
Q = flow (discharge) rate (m3/s).
Ce = effective discharge coefficient or weir coefficient.
Bweir = overall width of the weir across the flow (m).
Refer to list of symbol.
In the case of an ungated suppressed weir, the overall width of the weir
across the flow Bweir has the same value as the overall width of the structure
between abutments Bstructure or as the clear waterway Bcl (refer to article 14.4
and table 14.2).
6.1.2. Effective discharge coefficient and streamlines at control
section
When using the velocity coefficient Cv, the effective discharge coefficient is
given by the following equation and depicted in the following figure:

19
CdP User’s Manual February 2017 (second edition revised and expanded)

Ce = Cd * Cv * 2/3 * (2/3 *g)1/2 (6.1c)


Where:
Refer to list of symbols.
The discharge coefficient Cd, hence Ce, is affected by the friction on the
channel wall and bottom between the gauging and the control section, by the
velocity profile in the approach channel and control section and by the
changes in pressure distribution caused by streamline curvature as shown in
the below figure1.

Figure 6.2: illustration of conditions of streamlines at control section affecting


the discharge coefficient

In equation 6.1b, the semi-empirical effective discharge coefficient (or weir


coefficient) Ce equals in practice:
 1.7 for rectangular broad crested weir state; Cd remains fairly
constant2 with a smooth and sufficiently long broad crest creating
parallel streamlines at the water level control section (refer to figure
6.2 and article 6.4.2 on how to obtain it).
 1.84 to 1.9 for short crested weir state; the streamlines (flow lines) at
the control section are curved as shown in figure 6.2. Since the
discharge coefficient Cd increases if the streamlines curvature at the
control section increases, Cd should be multiplied by a corrector factor
which is always greater than unity (varying between 1.1 and 1.3). This
phenomenon explains the change of value of the effective discharge
coefficient to approximately 1.9 (1.84 in the User’s Manual) for short
crested weir state (refer to article 6.4.2).
6.1.3. Velocity coefficient

1 A minimum distance for the approach channel is necessary for the development of uniform and symmetric flow
conditions and the establishment of a stable water surface (refer to article 6.4.1).
2
Changes of Cd as a function of H0 are usually insignificant.

20
CdP User’s Manual February 2017 (second edition revised and expanded)

In the field, it is not possible to measure the energy head H 0 directly and it
is common practice to relate the discharge to the u/s water head h0 as shown
in equation 6.1.b. A positive velocity coefficient correction is therefore
necessary for neglecting the velocity head Va2/2g in the approach channel in
order to have the true magnitude of Ce for the above equation 6.1b. Cv may be
approximated from the following equation for a rectangular cross-section:

Cv = (H0 / h0)1.5 (6.1d)


Where:
Refer to list of symbols.
Finally, Cv varies from 1.01 up to 1.30 and usually is equal to 1 with low
approach velocity1. The cases given above show that traditional discharge
equations are often a mixture of rational analysis and experimental coefficient
evaluation.
6.2. UPSTREAM FLOW VELOCITY AND KINETIC ENERGY HEAD

6.2.1. Approach flow velocity


The velocity to be considered is the approach (or accelerated) velocity Va at
the entry of the structure (Manning’s equation). For solving the problems in
hydraulic engineering, the velocity used is the average velocity of flow over a
section.
With a rigid section, the permissible limits of velocity are given in the
following table:
Table 6.1: Maximum permissible average velocities of water in canals
Max. permissible
Type of soil or lining average velocity
(m/s)
Unlined canal
Soft clay or very fine clay 0.2
Very fine or very light pure sand 0.3
Very light loose sand or silt 0.4
Coarse sand or light sandy soil 0.5
Average sandy soil and good loam 0.7
Sandy loam, small gravel 0.8
Average loam or alluvial soil 0.9
Firm loam, clay loam 1.0
Firm gravel or clay 1.1

1
During flood stages, the velocity varies greatly; hence the velocity head is usually included in the total energy
head.

21
CdP User’s Manual February 2017 (second edition revised and expanded)
Stiff clay soil, ordinary gravel soil, or clay & gravel 1.4
Broken stone and clay 1.5
Coarse gravel, cobbles, shale 1.8
Conglomerates, cemented gravel, soft slate 2.0
Soft rock, rocks in layers, tough hardpan 2.4
Hard rock 4.0
Lined canal
Cast-in-place cement concrete 2.5
Precast cement concrete 2.0
Stones 1.6-1.8
Cement blocks 1.6

However, the conveyance reach u/s of the weir should have a sub-critical
flow as to limit the head loss (Froude number Fr < 1). The Froude number Fr can
be calculated with the formula 10.3a or 10.3b (V1 becomes the approach
velocity Va at the entrance of the structure; y1 is the u/s water depth within the
structure and B is the open-water width of the structure). In the User’s Manual,
it is often Iimited to < 0.5 (refer to article 6.4.1) to avoid standing waves in the
conveyance reach. Hence, the recommended maximum approach velocity Va at
the entrance of the structure is given by the following equation:

Va max = 0.5 (g * y)0.5 (6.2a)


Where:
Va max = maximum approach velocity at entrance of structure (m/s).
y = u/s water depth in the structure (m).
6.2.2. Kinetic energy head
When all parts of a system move with the same velocity, the kinetic energy
per unit mass is expressed as:

Va2 / 2g = (Q / A)2 / 2g (6.2b)


Where:
Va2/2g = kinetic energy head due to u/s velocity (m).
Q = low water discharge rate (m3/s).
Refer to list of symbols.
6.2.3. Upstream total energy elevation HE0
The elevation of the total u/s mechanical energy HE0 is the sum of the u/s
water depth in relation to the weir sill, the kinetic energy head due to u/s flow
velocity and the crest elevation.

22
CdP User’s Manual February 2017 (second edition revised and expanded)

HE0 = crest sill height + Va2 / 2g + h0 (6.3)


Where:
Refer to list of symbols.
6.3. DOWNSTREAM FLOW VELOCITY AND KINETIC ENERGY HEAD

6.3.1. D/s elevation HE2

HE2 = HE0 – ΔE (6.4)


Where:
Refer to list of symbols.
6.3.2. Kinetic energy head V32/2g

V32 / 2g = (Q / A)2 / 2g (6.5)


Where:
Q = low water flow rate (m3/s).
Refer to list of symbols.
6.4. DISCHAGE EVALUATION AND FLOW STATE

Once critical conditions are established over the weir, it can be used as a
flow-metering device.
6.4.1. Discharge measurement station
The approach of the User’s Manual is to combine the water level control
and the flow measurement functions by
establishing the corresponding rating curves
(refer to article 17).
The energy head H0 measurement (staff
gauge or ruler) should be located at a
distance between 2 or 3 times H0 max u/s of
the weir toe (refer also to figures 5.1 and
6.1). In the case of gates with piers (open Figure 6.3: ruler location
channel or breast wall type), the distance is
considered from the weir or the u/s nose of the pier whichever is located more
upstream.
The Froude number should not exceed 0.5 at the staff gauge location at
maximum flow. This is a primary design criterion intended to ensure that the
water level can be measured with reasonable accuracy at the gauge and to
avoid standing waves as the water surface and flow to become critical due to

23
CdP User’s Manual February 2017 (second edition revised and expanded)
possible decreased canal roughness from the river or parent channel into the
canal1.
6.4.2. Flow state of gauging weir
The discharge coefficient Cd corrects for such phenomena as the energy loss
between the gauging and the water level control section, the non uniformity of
the velocity distribution and the streamline curvature in these two sections.
These phenomena are closely related to the value of the ratio H0/L.
In terms of the ratio H0/L (refer to figures 6.1 & 6.2), four different flow
states over the weir may be distinguished:
H0/L < 0.08: over this range, the weir cannot be used as a measuring device.
 The depth of flow over the weir crest is such that sub-critical flow
occurs above the crest.
 The water level control section is situated near the d/s edge of the weir
crest.
 The energy lost through friction above the sill is a relatively large part
of H0 (the thin layer of water above the sill is very close to the rough
boundary).
 Cd is determined by the resistance to flow characteristics of the crest
surface (roughness coefficient).
0.08 ≤ H0/L ≤ 0.33: over this range, the weir may be described as broad crested.
 A region of parallel flow occurs somewhere midway above the crest.
 The control section is located at the end of the section where parallel
flow occurs.
 Cd has a constant value over this H0/L-range, provided that the
approach velocity Va has no significant influence on the shape of the
separation bubble
0.33 < H0/L < 1.5 to 1.8: over this range, the weir should not be termed as broad
crested but as short-crested2.
 The two downward slopes of the water surface merge and parallel
flow will not occur above the crest.
 Streamline curvature at the water level control section has a
significant positive effect on the discharge, resulting in higher Cd
values.
H0/L > 1.5: over this range, the weir acts as a short crested weir
 The ratio H0/L has such a high value that the nappe may separate
completely from the crest.

1 To eliminate this error, increase the size of the approach channel, reduce the control section width, or increase
the height of the crest relative to the invert of the approach channel.
2 Frequently, in this range of H /L values, the weir is also classified as broad crested.
0

24
CdP User’s Manual February 2017 (second edition revised and expanded)

6.5. FLUMING OF WATERWAY

The contraction of the waterway of the channel (i.e. fluming of the


waterway) will reduce the open water width of the cross-regulator, which is
likely to produce economy in many cases.
The maximum fluming is generally governed by the extent that the velocity
in the trough that should remain sub-critical (refer to article 6.4.1).
However, the greater is the fluming, the greater is the length and height of
the d/s transition wings and the length of the d/s transition wings.

25
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 6.4: illustration of a typical ungated cross-regulator structure with glacis

26
CdP User’s Manual February 2017 (second edition revised and expanded)

7. MODULAR FLOW AND SUBMERGENCE RATIO


7.1. MODULAR FLOW

The following figure shows the different flows over a cross-regulator:

Figure 7.1: modular and drowned (submerged) flow conditions

The flow over a channel feature is modular when it is independent of


variations in tailwater surface elevation. The flow is usually contracted and
passes through a critical depth yc. To maintain a modular flow through a
structure (i.e., critical depth at the control section), there must be some head
loss through the structure due to the energy losses caused by friction and the
expansion of the flow in the sudden or gradual protected transition d/s of the
throat (weir)1.
7.2. MODULAR LIMIT

The modular limit is often expressed as the submergence ratio h2/h0 (or
H2/H0). In the User’s Manual, the adopted value of the modular limit for a broad
crested weir to operate satisfactorily in free flow conditions is 0.75 (or 75 %),
that is when h2 = 0.75*h02. The submergence ratio is given by the following
equation:

1 When the d/s water surface elevation rises above a critical point, the resistance to flow in the d/s channel
becomes sufficient to reduce the d/s velocity, increases the d/s flow depth, and causes a backwater effect in the
structure; the broad crested weir will cease to operate in modular conditions. The head discharge relationship
can no more be applied once the head of water above the crest of the tailwater (h2) exceeds the critical depth yc
on the crest (the flow no longer passes through a critical depth). Further increases in d/s depth cause the d/s
depth to increase without a change in discharge.
2 For sharp crested weirs, the head-discharge relationship becomes inaccurate at a submergence ratio of around

0.22, so the broad crested type has a wider operating range.

27
S = 100 * h2 / h0 (7.1)
Where:
S = submergence (%).
Refer to list of symbols.
The u/s water head h0 in relation to weir sill is computed with the following
equation:
h0 = H0 – Va2 / 2g (7.2)
Where:
Refer to list of symbols.
The d/s water head h2 in relation to weir sill may also be taken as the
difference between the tailwater surface elevation and the weir sill elevation.
If the tailwater channel or canal is relatively wide or if the tailwater surface
elevation is affected by a d/s structure, it may occur that the weir as a
measuring structure is modular at its maximum design capacity, but non
modular with lesser discharges. Under such circumstances, a decrease in the
u/s head means an increase in the submergence ratio h2/h0. The crest of the
control section should then be raised so that h 2, and thus the ratio h2/h0,
decrease to below the modular limit 1.
7.3. DISCHARGE RATE CONSIDERATIONS

With modular flow, the water user with his own canal inlet cannot increase
the discharge by lowering the tailwater. On the other hand, in case of a
submerged flow, the water user can increase the discharge by lowering
artificially the tailwater level.
The decrease in the u/s velocity due to the submergence may lead to
aggravate sedimentation problems, in addition to the obvious effects of
increasing the complexity of determining the reduced discharge rate and
raising the d/s flow depth (which could lead to overtopping the bank channel).
It is important to remember that d/s water levels can change with changes
in d/s flow resistance, which frequently varies with sediment deposits, debris,
canal checking operations etc. Increased d/s flow resistance can result in
structures originally designed for free flow to experience submergence.

1 A common mistake done by the designer is to calculate the head that will occur over a weir at a particular
discharge without considering at all the heights of weir required to obtain a modular flow. The approach taken in
the User’s Manual is to allow the user to analyze alternatives to reach a more economic design by determining
the minimum crest height for which modular flow can be obtained.

28
CdP User’s Manual February 2017 (second edition revised and expanded)

7.4. HEAD LOSS AND TRADEOFFS

Rectangular weirs have proven to be an effective option since they can be


built easily with a simple design process. Some important design criteria have
been considered in article 6 (Froude number, submergence etc.). Some options
are given in the below table to satisfy the three primary design criteria for a
gauged weir:
1. The Froude number must be less than 0.5 at the gauge for Qmax (refer
to article 6.4.1).
2. The gauged weir must not be submerged at Qmin.
3. The gauged weir must not be submerged at Qmax.
Table 7.1: options for satisfying design requirements for gauged weirs
Design requirement Options
At Qmax
Raise the crest
Narrow the control section
Modular flow (avoid
Add or modify d/s slope ramp to regain potential energy
submergence)
Increase d/s expansion ratio (refer to article 25.3)
Choose a location where more drop is available
Increase approach channel top width
Low Froude number
Deepen the approach channel
At Qmin
Raise the crest
Narrow the control section
Modular flow (avoid
Add or modify d/s slope ramp to regain potential energy
submergence)
Increase d/s expansion ratio (refer to article 25.3)
Choose a location where more drop is available

29
8. FLOW VELOCITY AND SEDIMENTATION
The hydraulic structures should be designed to prevent sedimentation as
much as possible.
The faster the water travels, the larger the particles of silt it can carry with
it. The following figure shows the relationship between the size of sediment
and the velocity required to erode (lift it), transport it and deposit it (Hjulström
curves). The critical deposition curve shows the maximum velocity at which a
river or parent canal can be flowing before a particle of a certain size is
deposited. The zone in-between is the zone of transportation in suspension1.

Figure 8.1: erosion susceptibility chart (Hjulström curve)

For example, in the above figure, it can be seen that, when the flow velocity
reaches 50 cm/s, the current would pick up particles of 0.02 mm diameter from
the natural bed of the channel and erosion commences.
For preventing sedimentation (deposition) from taking place, the water
around the structure should no slow down enough to drop its sediment. If this
is not possible, a mechanism for flushing the area subjected to sedimentation
should be provided (desilting or flushing).

1Velocities for transportation are lower than that for erosion, because it takes much more energy to lift sediment
than to maintain it. The other strange pattern is that it takes more energy to erode some of the smallest particles.
This is because they are clay particles which are clogged or bonded together, therefore require a lot of energy to
be eroded.

30
CdP User’s Manual February 2017 (second edition revised and expanded)

PART III:

FLOW ENERGY DISSIPATION

31
CdP User’s Manual February 2017 (second edition revised and expanded)

9. ENERGY DISSIPATION
9.1. TOTAL AND SPECIFIC ENERGY

9.1.1. Total energy


The total mechanical energy (Bernoulli) is measured from some fixed datum
(refer to below figure) and its value can only reduce as energy is lost through
friction1.

Figure 9.1: datum, total and specific energy

9.1.2. Specific energy


The specific energy (Bakhmeteff) is the average energy per unit weight of
water as expressed in relation to the channel bottom (datum); it is the sum of
the water depth and the velocity head and measured from the bed of a
channel2 (refer to above figure).
When the bed level changes, the specific energy also changes. It also means
that specific energy can rise as well as fall, depending on what is happening to
the channel bed: when the flow moves from the channel over a cross-regulator,
the specific energy falls; when it comes off the cross-regulator it rises again.
The total mechanical energy of a channel flow refers to a datum; if the
datum coincides with the channel bed elevation at the section (datum = 0 on
horizontal floor), the resulting expression is known as specific energy.
The difference between total and specific energy can be illustrated by the
uniform flow. The total mechanical energy falls gradually as energy is lost
through friction. However, the specific energy remains constant along the
channel because there are no changes in water velocity and depth (no raised
bed, hump etc.).

1 When there is a change in the bed level of a channel (e.g. when water flows over a weir), there are also changes
in the energy components but the total mechanical energy remains the same.
2 Simply stated the specific energy is the energy in a channel measured from the bed of a channel.

32
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 9.2a: specific energy curve and alternate depth

For a given unit flow q, there are two possible depths in the channel for the
same specific energy, which are called alternate depths y1 and y2.
The critical flow condition (yc) is considered the most efficient, although not
necessarily the most desirable 1. To the critical depth yc corresponds the critical
energy or minimum energy Emin. Unique critical values exist for any discharge.
For a given flow rate per unit width q0 in a rectangular channel, to the critical
depth corresponds the minimum specific energy with:

Emin = yc * 3 / 2 (9.1a) or yc = Emin * 2 / 3 (9.1b)


Where:
Refer to list of symbols.
The Froude number = 1 (refer to article 10.3.3).
9.2. HYDRAULIC JUMP

A hydraulic jump is to quickly reduce the excess energy (velocity) of the


flow, passing from a predominant kinetic energy term v2/2g on a paved apron
to a point where the flow becomes incapable of scouring the d/s channel bed
with a predominant heat and potential energy. Other practical applications of
the hydraulic jump are many such as:
 Prevention or containment of the scour d/s of the structure.

1 Critical flow is unstable and, generally, it cannot be maintained over a long distance; it is rather a local
phenomenon. Due to the shape of the specific energy curve close to the critical point, a small change in energy E
(possibly as a result of small channel irregularities) can result in significant fluctuations in water depth as the flow
oscillates between sub- and super-critical flow. In other words, especially in super-critical flow, a slight change of
water depth may correspond to a great change of specific energy dissipated.

33
CdP User’s Manual February 2017 (second edition revised and expanded)

 Increase in the discharge of a sluice gate by holding back the


tailwater1.
 To recover head or raise the water level in the channel for
irrigation or other water distribution purposes.
 To increase weight on an apron and thus reduce direct uplift
pressure under a solid floor2.
9.3. HYDRAULIC JUMP VARIABLES

The design of a hydraulic jump involves the following 8 variables (also


shown in the following figure):
 E1 or Ef1: calculated or estimated specific energy head
(Bakhmeteff) in super-critical (subscript 1) or in sub-critical
(subscript 2) flow range of jump in relation to canal bottom (m).
 V1: super-critical velocity (m/s).
 y1: super-critical initial depth of flow (m).
 V2: sub-critical velocity (m/s).
 y2: sub-critical sequent depth of flow 3 (m).
 q: unit discharge (m3/s.m).
 ΔE: specific energy loss or dissipation in the standing wave (m).
The following figure shows a hydraulic jump on horizontal bed interpreted
by the specific energy and the specific force (due to fluid pressure and
gravitation acting on the fluid particle) curves for a constant unit discharge q.

1 The effective head is reduced if the tailwater is allowed to drown the hydraulic jump.
2 The raising water depth on the apron reduces the uplift pressure.
3 In fluid dynamics, the conjugate depths or sequent depths refer to the depth (y ) u/s and the depth (y ) d/s of
1 2
the hydraulic jump whose momentum fluxes are equal for a given unit discharge (volume flux) q but specific
energy E1 and E2 are different due to the energy losses in the jump. It is important to note that the conjugate
depths are different from the alternate depths that have the same specific energy (no losses in the system due
to friction and slope absolutely flat).

34
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 9.2b: jump on horizontal floor (or very mild slope), water depths and
specific energies for a constant flow rate

The depth before the jump is called initial depth y 1 and that after the jump
is called sequent or conjugate depth y2; energy loss is involved 1. The values
shown are for a unit discharge q1 only. Different q values are not plotted on the
specific energy diagram above2.

Figure 9.2c: hydraulic jump on horizontal floor interpreted by specific


energy and specific force curves

1 Sequent depths are depths in which specific energy loss is involved. Alternate depths have the same specific
energy. If there is no energy losses, the initial and sequent depths would be identical with the alternate depths
in a prismatic channel.
2 On the diagram, there is a distinct asymptotic relationship as the top part of the curve approaches the E = y line

which means that v22/2g decreases (refer to equation 10. 1 in terms of unit flow rate q and y) and the bottom
part of the curve tends toward the x axis.

35
CdP User’s Manual February 2017 (second edition revised and expanded)

The u/s and d/s specific energy are not equal (E1 ≠ E2) (Bernoulli energy
equation not applicable) but the u/s and d/s momenta are the same
(momentum equation applicable with M1 = M2).
The following figure shows side by side the specific energy curve (plotted
for a constant unit discharge q) and the specific discharge curve (plotted for a
constant specific energy) in a rectangular section.

Figure 9.3: graphs of specific energy and specific discharge curves

It must be pointed out that the last two variables (q and ΔE) are variables of
the hydraulic design:
1. For a fixed (constant) unit discharge q:
 E minimum at yc.
 For all other values of E, there are two possible depths (sub-critical
and super-critical depths).
2. For a fixed (constant) specific energy E:
 q maximum at critical depth yc.
 For all other values of q, there are two possible depths (sub-
critical and super-critical depths).
The above diagrams also indicate another difference between super-critical
and sub-critical flow. For a particular value of discharge, a decrease in E head in
sub-critical flow (Fr < 1) results in a decrease in the water depth1. On the
contrary, with super-critical flow (Fr > 1), the figure indicates that the opposite
happens: a decrease in E results in an increase in the water depth2.

1 For instance, the water surface is drawn down as the flow passes an obstruction in the channel such as over a
weir or through a gate chamber or a bridge opening.
2 Due to this difference of behaviour, it is indispensible for the design engineer to know which type he is being

dealt with.

36
CdP User’s Manual February 2017 (second edition revised and expanded)

9.4. JUMP FORMATION

When the flow changes from super-critical state (where the kinetic energy
predominates) to sub-critical state
(where the thermal and potential
energy predominate), the jump
takes place. The change of state is
due to an abrupt break in the
bottom slope where the glacis
slope suddenly turns flat when
reaching the basin1 or due to the
presence of a gate, as shown in the
opposite figure.
Figure 9.4: hydraulic jump Once there is super-critical flow in
a channel, it is the d/s depth of
flow that determines if a jump will occur. To create a jump the d/s depth must
be just right. If the depth is too shallow a jump will not form and the super-
critical flow will continue down the channel.
The following table summarizes the necessary conditions for the formation
of a hydraulic jump.
Table 9.1: conditions for the formation of a jump at a considered section
Jump Condition Flow Fr1
Formation at considered section y1 < yc Super-critical >1
No formation at considered section y1 > yc Sub-critical <1

1 No hydraulic jump with if y1 = yc.

37
CdP User’s Manual February 2017 (second edition revised and expanded)

10.ANALYTICAL DESIGN OF THE HYDRAULIC JUMP


10.1. HYDRAULICS OF THE STILLING BASIN

The equations proposed below are applicable if:


1. The jump takes place abruptly.
2. The flow remains streamlined throughout (refer to article 1.3).
3. The friction loss is negligible (phenomenon takes place in a short
distance)1.
4. Rectangular cross-section.
Always take care to no apply a particular equation to the wrong channel
shape.
10.2. SPECIFIC ENERGY EQUATION

Energy is the most convenient parameter for determining the flow


characteristics. In uniform flow, the total mechanical energy falls gradually as
energy is lost through friction 2. But specific energy remains constant along the
channel because there are no changes in depth and velocity.
The value of the specific energy E2 of the sub-critical reach of the stilling
basin and of E1 of the super-critical reach is calculated as the sum of the
pressure head term3 and the velocity head term measured with respect to the
channel bottom as shown with the following equations (with Z = 0) (from
Bernoulli’s energy equation)4:

V22 V21 𝒒𝟐
E2 = y2 + ; E1 = y1 + ;𝐄 = 𝐲 + (10.1)
2g 2g 2g𝒚𝟐
Where:
y = static pressure head (flow depth) (m).
v2/2g = dynamic pressure head (kinetic energy) (m).
Refer to list of symbols.
With the required tailwater, the velocity leaving a properly designed stilling
basin (V2) should equal the velocity V3 of the receiving canal or channel.
10.3. COMPUTATION OF THE CRITICAL AND SEQUENT DEPTHS OF FLOW

10.3.1. Determination of critical water depth yc

1 For chute (glacis) less than 9 m long, the friction in the chute can be neglected.
2 In other words, the friction slope is equal to the bottom slope when the head loss is equal to the elevation drop.
3 = (ρ*g*h) where p is pressure (kN/m2); ρ is mass density of water (kN/m3); g is gravity constant (m/s2); h is depth

of water (m).
4 By reordering the equation, we can have the specific energy in terms of a flow rate per unit width and the

channel water depth (E = y + q2/(2 * g *y2). In other words, we removed the velocity term which, sometimes, is
not useful, and we replaced it with a flow rate term. We now have rewritten the equation in terms of only one
parameter: depth of flow (refer to graphs of figures 9.2).

38
CdP User’s Manual February 2017 (second edition revised and expanded)

The point of inflection on the specific energy curve marks the approximate
position of the critical depth yc where the specific energy is minimum for a given
discharge or where the unit discharge is maximum for a given specific energy
(refer to figure 9.1 and 9.2). The Froude number equals unity and the minimum
energy is given by equation 9.1.
In a rectangular cross-section, the critical depth yc is determined by applying
Bernoulli’s energy equation.

yc = (q2 / g)1/3 (10.2a) with q = Q / Bstructure = y * V (10.2b)


Where:
Refer to list of symbols.
From the above equation, the critical depth is influenced only by the discharge
per unit width (which means also the velocity) and the structure geometry: yc
increases with increasing discharge and, for a constant discharge, increases with
decreasing channel width. On the contrary of yn, it has nothing to do with the slope
or the roughness coefficient of the channel (refer to figures 14.1 and 14.4).
10.3.2. Determination of super-critical water depth y1
The determination of the super-critical depth y1 and the velocity V1 before
the jump (usually at the toe of glacis) is based on the assumption of no energy
loss between the upper pool and the toe of the glacis (theoretical base point
formation of the jump)1, as shown in the opposite figure. Therefore, the energy
equation (Bernoulli equation) applies. Knowing the upper pool elevation, the
u/s velocity head (if significant) and the
discharge, the super-critical depth y1 and
entering velocity V1 can be solved by trial
and error for an assumed stilling basin
floor elevation (in the User’s Manual, the
assumed floor elevation initially equals
Figure 10.1: energy grade line the d/s channel bed elevation)2. As an
in glacis and jump important hydraulic characteristic, the
flow depth at the bottom of the glacis is
also the flow depth preceding the hydraulic jump in a properly designed stilling
basin.

1
For short and smooth transitions, the energy losses are negligible and the Bernoulli equation may be applied
quite successfully. Short transitions may be gates (e.g. or radial gates), weirs and glacis.
2 Glacis are not functioning as a chute having sub-critical uniform flow because the energy grade line is not parallel

to the slopping glacis; the condition for functioning as a chute with sub-critical uniform flow is to have the energy
level parallel to the slopping chute dissipated by friction. The glacis is to short or too smooth so that the friction
loss is very small and ignored. Thus, the energy dissipation takes places not in the glacis reach but only in the
stilling basin thanks to the hydraulic jump and eddies.

39
CdP User’s Manual February 2017 (second edition revised and expanded)

However, the energy equation cannot be used between y 1 and y2 because


there is an unknown energy loss at this portion of the flow. The momentum
equation is used instead to describe the flow between those two positions.
10.3.3. Determination of the u/s Froude number
The Froude number is the ratio of the velocity of a stream divided by the
speed of a small wave on the water surface relative to the speed of the water,
called wave celerity 1 (Fr = water velocity/wave celerity).
The Froude number Fr (or Fr1 before the jump) is computed according to the
following equation (rectangular cross-section)2:

Fr1 = V1 / (g * y1)1/2 (10.3a) or Fr1 = q / (g * y13)1/2 (10.3b)


Where:
V1 = stream velocity before the jump.
Refer to list of symbols.
V1 is usually considered at the toe of the glacis if any. At critical flow, the
wave celerity equals the flow velocity and the Froude number = 1. Any
disturbance to the surface will remain stationary.
10.3.4. Relationship between the sequent depths
The ratio of the flow depths across the hydraulic jump is computed thanks
to the momentum conservation principle and the continuity equation 3 across
the jump. The u/s and d/s specific energies are no more assumed as equal (E1 ≠
E2) but the u/s and d/s specific forces are the same4. The forces of this
momentum equation ignore the forces of the baffle blocks if any in the stilling
basin.
Theoretically speaking, in a rectangular channel, a hydraulic jump takes
place at a point where the u/s and d/s sequent depths and the approaching

1 The celerity wave is typically equal to (y*g) 1/2 with y = the water depth in the channel at the considered point.
2 There are several ways to calculate the Froude number.
3 The continuity equation relates mass rate of flow along a streamline. It links u/s and d/s discharges (or areas

and velocities) and means that the amount of water flowing into a system must be equal to the amount of water
flowing out of it.
4 The momentum equation, stated as the sum of all external forces, is equal to the rate of change of momentum.

This equation links force (or mass) to velocity. The forces of this momentum equation are shown in the figure
below.
Assumptions: the friction can be
neglected (length very small) and the
slope close to 0 (gravitational
component of flow may be
neglected)

40
CdP User’s Manual February 2017 (second edition revised and expanded)

Froude number satisfy the below equation. This relationship between the initial
depth y1 and the sequent depth y2 of a hydraulic jump on a horizontal floor in
a rectangular channel is a quadratic equation in terms of the entering Froude
number and is solved by iteration. This equation is applicable even when the
flow enters the jump at an appreciable angle to the horizontal.

y2 / y1 = 1 / 2 * [(1 + 8 * Fr12)1/2 – 1)] (10.4)


Where:
Refer to list of symbols.
This equation can be swapped around and be in terms of not Fr1 as above
but critical depth yc1. There is no hydraulic jump formation when yc = y1.
10.4. CHARACTERISATION OF THE HYDRAULIC JUMP

10.4.1. Concentration and degradation


In the Manual, the jump characteristics are determined taking into account
the following hypothesis:
 No concentration of flow 2.
 No degradation (retrogression) of d/s level because of the
protective works (refer to article 24).
10.4.2. Base point and surface point elevation of jump formation
For a given energy head loss ΔE and a discharge intensity q, there is a
definite value of the quantity of d/s specific energy (E2 or Ef2) in sub-critical flow
range required for the jump formation. This quantity of d/s specific energy E2
or Ef2 indicates the theoretical position of the base point of the jump formation,
hence the elevation of the stilling basin floor.
The elevation of the position of the base point of jump formation is
computed as the difference between the total d/s mechanical energy elevation
HE2 and the calculated or estimated specific energy head E2 or Ef2 in sub-critical
flow range of the hydraulic jump as shown in the following equation:

Elevation of base point formation = HE2 – E2 (10.5)


Where:
Refer to list of symbols.
The elevation of the surface point of jump formation is obtained by adding
to the elevation of the base point of jump formation the super-critical depth y1.

1 The equation becomes, taking y1 as an example: y1 = (y2 / 2) * [(1 + 8 (yc / y2)3)0.5 – 1] or y1 = (y2 / 2) * [1 + 8 * q2
/( g * y23)0.5 – 1]. The 2 conjugate depths y1 and y2 may replace each other in the equations.
2 While calculating the flow, the possibility of non uniform flow is not taken into account by providing a suitable

concentration factor chosen arbitrarily (usually 20 % which means that the calculated maximum discharge rate is
increased by 20 % during the design phase).

41
CdP User’s Manual February 2017 (second edition revised and expanded)

10.4.3. Hydraulic jump efficiency


The dimensionless ratio of the specific energy after the jump (E2) to that
before the jump (E1) is defined as the efficiency of the jump and indicates how
much of the original energy stays after the jump. It is calculated with the
following equation:
𝟑
𝐄𝟐 (𝟖𝐅𝐫𝟏𝟐+𝟏)𝟐 −𝟒𝐅𝐫𝟏𝟐 +𝟏
= (10.6a)
𝐄𝟏 𝟖𝐅𝐫𝟏𝟐 ∗ (𝟐+𝐅𝐫𝟏𝟐 )
Where:
Refer to list of symbols.
The above equation shows that the efficiency depends only on the u/s
Froude number (Fr1). A greater Fr1 corresponds to a lower ratio E2/E1 and a
higher relative loss or energy dissipation in the jump as shown in figure 10.2
and as calculated with the following equation:

Relative loss in energy = 1 - E2 / E1 = ΔE / E1 (10.6b)


Where:
Refer to list of symbols.

Figure 10.2: relative loss in specific energy head of hydraulic jumps in


horizontal rectangular structure
10.4.4. Hydraulic jump energy head loss
The total mechanical head loss ΔE in the basin equals the difference of the
specific energies before and after the jump as shown in the following equation:

(𝐲𝟐 −𝐲𝟏 )𝟑
𝚫𝐄 = 𝐄𝟐 − 𝐄𝟏 = (10.7a)
𝟒∗ 𝐲𝟐 ∗ 𝐲𝟏
Where:

42
CdP User’s Manual February 2017 (second edition revised and expanded)

Refer to list of symbols.


For energy to be lost, it requires y2 > y1.
This quantity of mechanical energy dissipated is converted into thermal,
sound and potential energy as a result of frictional effects during the jump.
Although the energy dissipated in the highly turbulent motion in the jump can
be considerable, the difference in temperature is not great because the flow
rate is usually quite high. The energy per unit width dissipated by the hydraulic
jump at design flow conditions is calculated with the following formula:

Elost = γ * q * 𝚫𝐄 (10.7b)
Where:
Elost = energy dissipated per unit width in metre (kWs).
γ = specific or unit weight of water (9.81 kN/m 3)1.
Refer to list of symbols.
10.4.5. Hydraulic jump height
The height of the jump hj is the difference between the water depth after
(y2) and before the jump (y1). The jump elevation is the sum of the elevation of
the base point of the jump formation and the jump height 2.
The relative height of the jump can also be expressed as a ratio with respect
to the initial specific energy E1 as shown in the following equation:

𝐡𝐣 𝐲𝟐 𝐲𝟏
= − (10.8)
𝐄𝟏 𝐄𝟏 𝐄𝟏
Where:
Refer to list of symbols.
This relative height increases up to Froude number 2.77 and then decreases
non linearly as the value of Froude number increases to eight (refer to figure
10.3).

10.4.6. Hydraulic jump length


The length of a hydraulic jump may be defined as the distance measured
from the front face of the jump to a point on the surface d/s from the rollers
where it gets back to normal depth again. This length cannot be determined
easily by theory. Some empirical equations are proposed below (refer to article
12.3.2).

1Specific or unit weight of water γ =ρ*g with ρ in kg/m3(water density).


2The above calculation assumes that the complete jump is located upstream of the sill and that no additional
energy loss takes place at the sill.

43
CdP User’s Manual February 2017 (second edition revised and expanded)

In the User’s Manual, the length of the hydraulic jump equals the length of
the stilling basin, which may include an end sill (refer to article 12.5).
10.4.7. Characteristic curves of the hydraulic jump
Since the relative loss, the efficiency, the relative height and the relative
initial and sequent depth of a hydraulic jump in a horizontal (or nearly so)
rectangular cross-section are function of Fr1, they can be plotted against the u/s
Fr1, resulting in a set of characteristic curves shown in the following figure:

Figure 10.3: characteristic curves of hydraulic jumps in horizontal rectangular


structure

Theses characteristic curves provide the designer with a general idea about
the range and conditions under which the proposed structure is to be operated.
With reference to these curves, the following features may be noted:
1. The maximum relative jump height hj/E1 is 0.507, occurring at Fr1
= 2.77 as per USBR standard.
2. The maximum relative sequent depth y2/E1 is 0.8, which occurs at
y1/E2 = 0.4 and Fr1 = 1.73. For this value of Fr1 = 1.73, the transition
from undulating jump to a direct jump takes place (refer to figure
12.1).
3. When Fr1 = 1, the flow is critical and y1 = y2 = 2/3*E1 (refer to article
9.3). It means also that the velocity of the stream is the same as
the propagation of the wave (refer to article 10.3.3).
4. When Fr1 increases, the changes of all characteristic ratios become
gradual.

44
CdP User’s Manual February 2017 (second edition revised and expanded)

5. The efficiency E2/E1 decreases non linearly with Froude number


from 1 to 7.
10.5. STABILITY AND CONTROL OF THE HYDRAULIC JUMP

Theoretically speaking, in a rectangular channel, a hydraulic jump occurs at


a location where the sequent depths and the approaching Froude number
satisfy the equation 10.4. However, there can be conditions in a channel, such
as a d/s control, that can alter where the conjugate depths form. For instance,
tailwater depth can play a very influential role on where the jump will occur in
the channel and changes in this depth can shift the jump either u/s or d/s.
With the presence of a glacis, the jump may occur in either the steep glacis
(preferably at the lowest) or the cistern floor.
10.5.1. Pattern of jump positions
There are 3 alternatives patterns of jump positions or locations.
1. yn = y2: jump rating curve (y2 versus Q) coincides with the tailwater
rating curve in the canal (yn versus Q) at all discharges. This is the ideal
condition for scour protection because the jump will form and remains
at the toe of the glacis at all discharges. In such a case, a simple apron
is generally sufficient to provide protection in the region of the
hydraulic jump.
However, a little difference between the assumed and actual values of
the relevant hydraulic coefficients may cause the jump to move from
its estimated position. Consequently, some devices to control the
position of the jump are always necessary.
2. yn > y2: the tailwater rating curve in the canal (yn versus Q) is lying
above the jump rating curve in the basin (y2 versus Q) at all discharges.
The jump moves upwards and will drown out at the source by the
tailwater, becoming a submerged (under water) jump. It is probably
the safest design because the position of the submerged jump can be
most readily set. However, the design is not efficient and very little
energy will be dissipated1. The wave celerity may continue to move at
high velocity along the channel bottom for a considerable distance.
The remedy for this problem is to adjust the glacis slope to create the
proper conditions for a jump to occur somewhere on the glacis at all
discharges.
3. yn < y2: the tailwater rating curve in the canal (yn versus Q) is lying
below the jump rating curve (y2 versus Q) at all discharge. The jump
moves downwards to a point where the equation y 2/y1 for a given
specific energy is satisfied and a jump takes place. This case must be

1 In fact, the super-critical flow will rush underneath and still cause erosion for some distance downstream.

45
CdP User’s Manual February 2017 (second edition revised and expanded)

avoided in design because the jump, repelled from the scour resisting
apron of the stilling basin, will take place in the unprotected channel,
resulting in severe erosion.
The below table summarizes the pattern of the jump positions or locations,
taking into account that we must have the presence of a super-critical flow to
assure the formation of the jump (refer to table 9.1).
Table 10.1: summary of the pattern of the jump positions or locations with
super-critical flow
Jump location Conditions
Standing wave at the considered cross-section yn = y2
Moves upwards from the considered cross-section yn > y2
Moves downwards from the considered cross-section yn < y2

There is no simple remedy for a deficiency in tailwater depth, even with the
aid of appurtenances. Increasing the length of basin, which is the remedy often
attempted in the field, will not compensate for deficiency in tailwater depth.
Baffle piers and sills are only partly successful in substituting for tailwater
depth. Furthermore, if it is not possible to depress the basin floor because of
difficulties in excavation, then a lateral expansion of the basin remains the only
possibility for guaranteeing the required dissipation of energy. In such basins,
there are mainly two problems faced by the designer: the determination of
sequent depth and the estimation of energy loss.
10.5.2. Tailwater consideration
The tailwater depth1 fluctuates owing to changes of the discharges in the
canal or river. Usually, the tailwater levels need only be checked at Q min and
Qmax, since the flow should also be modular at intermediate flows, if the flow is
modular at Qmin and Qmax.
In most of the design, it is assumed that the tailwater depth remains fixed.
However, there are two situations of interest with regards to the d/s channel
tailwater:
1. The tailwater level d/s of the structure may be influenced only by d/s
channel friction. When this occurs, the flow is said to be uniform and
the water depth is the normal water depth yn (Manning’s equation)2.
However, owing to seasonal changes of discharge in the canal and of
the hydraulic resistance of the d/s channel, the flow velocity changes
the water depth (in doing so, the relation between the sequent depth

1 Although flow is still non uniform in the diverging transition, the energy losses due to friction are estimated by
applying the Manning equation.
2 When the flow in the d/s channel is at normal depth, the roughness, cross-section and bed slope of the channel

are equal for a sufficient distance such that the water level at the structure site is controlled solely by the frictional
resistance of the d/s channel (i.e., no backwater effects).

46
CdP User’s Manual February 2017 (second edition revised and expanded)

y2 and the tailwater depth yn). Therefore, the tailwater levels should
be determined for the seasonal maximum d/s channel roughness (n =
0.035 in the User’s Manual)1.
2. When there are obstacles in the channel d/s of the structure, the
tailwater level is not governed by normal depth y n, but by backwater
(or drawdown) from of the obstruction (overfall etc.). In such cases,
the tailwater level depends greatly on the properties and settings of
this d/s obstacle. From a practical standpoint, the easiest way to
determine the resulting tailwater level is to measure it during the
worst-case conditions.
Therefore, even with a broad-crested weir, tailwater needs to be checked
at both minimum and maximum flow, since backwater can cause high tailwater
depths even at low flows. Also, a small difference between the actual and
calculated values of hydraulic coefficients may move the jump formation point
from its estimated or calculated position.
The usual case represents the conditions in which the jump rating curve is
at a higher stage than the tailwater rating curve at low discharges but at a lower
stage at high discharge. As a result, the hydraulic jump position moves up and
down on the glacis and the channel, hunting for the right depth of flow. This
unstable behavior is often undesirable. The safest method (adopted here) to
ensure a more stable jump is:
 Check the stilling basin for a jump formation at low discharges.
 Combine with the basin a sloping apron for developing a satisfactory
jump at high discharges.
 Ensure a hydraulic jump to be formed within the stilling basin for the
maximum tailwater and discharge and ensure the energy dissipation
is sufficient while the stilling basin is not in full capacity.
The choice of the glacis and the basin involves consideration of economics
and hydraulic performance:
Tailwater rating curves are extremely important. The variation in a tailwater
rating curve may shift toward more flow capacity, less flow capacity, or oscillate
from one to the other and back again. The shift in rating may be abrupt, gradual,
or sporadic2. The selected tailwater curve is used in design of stilling basins, wall
heights, erosion protection and many other critical elements that make up a

1 For broad crested weir with a uniform cross-section and with tailwater at normal depth y n, the submergence
need only be checked at Q max because the tailwater level will generally decline faster than the u/s depth if the
flow rate is reduced.
2 The variation in a tailwater rating curve may be caused by sediment erosion or aggradation, deposition or

excavation of channel bed or bank material, variations in hydrologic events, loops in rating curves as flow
transitions from the raising to falling flood stages, inaccurate estimates of channel roughness (resistance to flow)
or by man-induced events.

47
CdP User’s Manual February 2017 (second edition revised and expanded)

total project design. It has been observed however that the majority of the
basins were designed for conjugate tailwater depth or less.
Therefore, it is imperative that the hydraulic engineer not only has an
accurate estimate of what the tailwater curve will be after the project
construction and throughout the life of the project, but designs a basin deep
enough to provide for full conjugate depth (or some greater depth to include a
factor of safety) at the maximum weir design discharge 1. For example, for
projects with loop rating curves, raising stages should be used for design of
stilling basins and erosion protection and falling stages used for setting wall
heights. The use of an average tailwater rating curve may yield inadequate
design for both wall height and the high-velocity flow areas. In many cases, a
cross-regulator structure is provided d/s to maintain the tailwater depth yn.
10.5.3. Control of the position of the hydraulic jump
Depending on the corresponding value of y1, the jump may occur either on,
after or before the break in slope as shown in the following figure (with y0 & y1
< yc):

y1 > y0 (y0 = current depth on glacis) : jump d/s of break point


y1 < yo (y0 = current depth on glacis): jump u/s of break point
y1 = y0) (y0 = current depth on glacis: jump on break point
Figure 10.4: theoretical position of the jump on the glacis

The cistern elevation usually matches the elevation of the jump base point
formation2. In the User’s Manual, the formation of all types of jump is usually
stabilized at the lowest end of the sloping glacis and not swept out of the stilling
basin causing excessive scouring. Therefore, the floor of the basin is set at such
a level that the tailwater surface elevation stays greater than the water surface
elevation in the basin under varying conditions of head (at the higher flow
range).
Taking into account that the tailwater rating curve at best is only
approximate and the exact position of jump and the type of flow cannot be

1
The designer is cautioned against spending too much effort in refining inconsequential parameters, such as
spillway pier shape coefficients, without paying sufficient attention to potential shifts in tailwater rating curves
that can, of course, have drastic influences on submerged structure capacity.
2 As a first approximation, the cistern elevation is provided at the level of the d/s channel bed elevation.

48
CdP User’s Manual February 2017 (second edition revised and expanded)

determined only on an analytical way, a set of four conditions are adopted in


the User’s Manual during the design phase to stabilize the position of the
hydraulic jump:
1. Condition Nº 1.A is met when the tailwater surface elevation yn is 10
% greater than the sequent depth y2 of the jump1.
Condition Nº 1.B is met when the depth of the cistern floor is increased
by 25 % of the value of E2.
2. Condition Nº 2 is met with the presence of a d/s end sill (refer to article
12.5) in cast with the cut-off wall in order to diffuse the residual
portion of high velocity jet reaching the end of the basin, as shown in
figure 12.1. The sizing of the end sill d/s of the dissipater is given by
the following equation (the fore slope of the end sill should be set at
1H on 0.5V):
hes = y1 (0.0536 * Fr1 + 1.04) (10.9)
Where:
hes: height of the end sill in relation to d/s channel or canal bed
elevation (m).
Refer to list of symbols.
3. Condition Nº 3 is met with the presence of
the proposed glacis slope (2H or 3H on 1V) or
other appropriate structure (refer to articles
10.5.4 and 12.2.2).
4. With a straight glacis creating a super-
critical flow with Froude number Fr1 between
Figure 10.5: chute
2.5 and 9, a single row of triangle blocks (glacis
blocks
blocks or chute blocks) as shown in the opposite
figure are provided at the foot of the flumed glacis to ensure a more
uniform flow. These chute blocks tend also to stabilize the jump and
shorten the length.
In addition, a balance has to be stricken between adequate dissipation of
energy and stability (fixed position on the glacis) of the hydraulic jump, taking
into account that the jump formation depends upon the hydraulic conditions
immediately after it.
At this point of the design, the assumed stilling basin elevation is checked
against the existing tailwater surface elevation and a new stilling basin floor
elevation is assumed until all the conditions for controlling the hydraulic jump
are satisfied.

110 % for basin type IV; 5 % for basin type II. In the Manual, the 10 % increase is generally introduced for all the
basin types as a FoS.

49
CdP User’s Manual February 2017 (second edition revised and expanded)

10.5.4. Glacis slope


The glacis type of fall is suitable up to a flow rate of 60 m 3/s and/or a
topographical drop up to 2 m. In case of topographical drops greater than 2 m,
drop structures are usually designed, such as vertical drop, chute drop, pipe
drop or cascade drop (refer to article 12.2.2)
The slope of the glacis u/s of the stilling basin has little effect on the jump
as long as the distribution of velocity and depth of flow are reasonably uniform
on entering the jump. Furthermore, if the slope of the glacis is not too
pronounced, the gravity component of the momentum equation can be
neglected on the grounds that the error so introduced is practically cancelled
by the errors involved in other assumptions. Therefore, in this User’s Manual,
a 2H on 1V or 3H on 1V slope is adopted for the sloping glacis and the equations
derived for a horizontal bed may be applied to a sloping glacis.
This proposed glacis slope is usually sufficient to create on the glacis a flow
velocity V1 greater than the critical velocity Vc, triggering the formation of the
jump, ideally at the toe of the glacis.

50
CdP User’s Manual February 2017 (second edition revised and expanded)

11.HYDRAULIC JUMP DESIGN WITH DESIGN CHART


11.1. DESIGN CHART

The hydraulic jump characteristics can be estimated with the following


abaci:
 Blench’s design chart for estimating the specific energy Ef2 of the flow
in sub-critical state d/s of the jump.
 Montague’s design chart for estimating the sequent depths.
 Khosla’s design chart for uplift pressure (refer to article 23).
11.2. ESTIMATION OF THE SPECIFIC ENERGY

Blench’s curves shown in the following figure give the estimated specific
energy Ef2 d/s of the jump by taking the discharge intensity per unit width q and
the head loss HL (ΔE) as known variables 1.
At this level, a difference might occur between the computed and estimated
energy ΔE due to the following:
 The specific energy E2 is calculated with the sequent water depth y2;
the velocity head is calculated with the sub-critical velocity V2.
 The specific energy Ef2 is estimated by means of the tailwater depth yn;
the velocity head is calculated with the d/s velocity V3.
This difference is summarized in the following table.
Table 11.1: variables used for the calculation of E2 and estimation of Ef2
Water depth
Energy Velocity (m/s)
(m)
E2 y2 V2 (sub-critical reach)
Ef2 yn V3 (d/s of the stilling basin)
11.3. SEQUENT DEPTHS ESTIMATION WITH DESIGN CHART

Knowing the specific energy Ef, the conjugate depths y1 and y2 can be read
directly from the energy of flow curves given in Montague’s design chart,
corresponding to certain discharge intensity per unit width of stream.

1
The head loss considered may be the difference between the u/s and d/s water surface elevation.

51
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 11.1: Blench’s design chart for the estimation of the specific
energy Ef2 with known discharge intensity and energy loss

52
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 11.2: Montague’s design chart for the determination of energy flow
curves for different values of discharge intensity per unit width q

53
CdP User’s Manual February 2017 (second edition revised and expanded)

11.4. POSITION OF THE POINTS OF JUMP FORMATION

11.4.1. Elevation of the base point of jump formation


The position of the base point of jump formation is determined by checking
the curves of the Blench’s design chart (refer to the above figure) relating the
energy head loss ΔE (HL) and the unit discharge q to the estimated specific
energy Ef2.
11.4.2. Elevation of the surface point of jump formation
The elevation of the surface point of jump formation is obtained by adding
to the elevation of the base point of jump formation the estimated super-
critical initial depth y1.
11.5. CHARACTERISATION OF THE HYDRAULIC JUMP

The approach is the same as in the case of the analytical design (refer
to article 10.4).

54
CdP User’s Manual February 2017 (second edition revised and expanded)

PART IV:

EXTERNAL FLOW ENERGY DISSIPATER

55
CdP User’s Manual February 2017 (second edition revised and expanded)

12.STANDARD ENERGY DISSIPATER STRUCTURES


12.1. INTRODUCTION

Energy dissipation can occur in different manners: (i) due to friction, (ii) by
impact of the flow against the floor, (iii) by turbulence in a stilling basin and (iv)
by heat and sound. Therefore, different types of energy dissipater structures
can be distinguished.
Stilling basins are external energy dissipaters placed at the outlet of the
structure. These basins are characterised by some combination of chute blocks,
baffle blocks and end sill designed to trigger and/or maintain a hydraulic jump
in combination with tailwater conditions. The stilling basin design guidance
presented in this chapter is for free flow. Stilling basins designed for submerged
flow normally require a model study.
USBR has developed various types of generalized water energy dissipater
structures with rectangular cross-sections that will be used in the User’s
Manual. Each type of energy dissipater has to be investigated. In many cases,
more than one solution to a particular problem will be possible. The choice will
depend upon considerations of economics, material and complexity for
construction.
Other forms of energy dissipater include the vertical drop structure with
hydraulic jump or impact blocks where the dissipation takes place by impact of
the inflow on a vertical baffle.
12.2. SELECTION OF THE ENERGY DISSIPATER

The stilling basin is the common type of dissipaters for weirs. The selection
of a stilling basin depends upon:
• Approach flow conditions (Froude number).
 Hydraulic limitations.
 Basin size, constructability and cost
• Tailwater characteristics.
• Scour potential.
• Personal preference and experience.
12.2.1. Froude number
The u/s Froude number is the main criterion to select among the different
types of standard stilling basins used for dissipating the energy.
In the following figure, the jumps and corresponding USBR stilling basin are
classified according to the Froude number Fr1 of the incoming flow. The higher
the Froude number is at the entrance of the basin, the more efficient are the
hydraulic jump and the shorter the resulting basin.

56
CdP User’s Manual February 2017 (second edition revised and expanded)

If the Froude number does not fall within an acceptable range, the following
methods can be employed:
 Reduce the unit discharge by expanding the crest width.
 Increase the drop height by raising the crest elevation of lowering the
stilling basin floor elevation.
 A combination of above.
To increase the Froude number, the potential1 energy has to be converting
into kinetic energy; opposite changes for decreasing the Froude number 2:
12.2.2. Topographical drop
When the topographical ground slope exceeds a 2 m drop (and less than 6.1
m), this excess will be accounted in the User’s Manual by providing an
alternative to the glacis, with a vertical drop or a fall type of structure (refer to
article 10.5.4 and 12.4).
12.3. CARACTERISTICS OF THE ENERGY DISSIPATER

12.3.1. Classification of energy dissipater


The classification of hydraulic jumps and dissipaters in figure 12.2 may be
considered as a guideline for hydraulic jumps in rectangular horizontal
channels.
12.3.2. Total length of dissipater
On horizontal floor, the total length of a stilling basin is considered in the
User’s Manual as related to the length of the hydraulic jump before the drop
off caused by the d/s channel conditions
(refer to article 10.4.6). This length may
include an end still.
This length cannot be determined
easily by theory and some empirical
equations and graphs are proposed
Figure 12.1: hydraulic jump below. In case of a jump on sloping floor,
length on glacis the depth ratio y2/y1 increases with the
slope of the floor as shown in the opposite
figure. It means that the jump on sloping floor requires more tailwater depth
than the corresponding horizontal floor jumps. However, in the User’s Manual,
the equations derived for a horizontal floor are applied to the sloping glacis
(refer to article 10.5.4), considering also that a proper tailwater depth is
provided for maximum flow (refer to article 10.5).

1 When there is potential energy only, the flow rate is zero (the water is stagnant). When there is kinetic energy
only, the flow rate is also zero because the water depth has to be nil The maximum flow rate lies between these
two extremes.
2 In practical design, the economies would be checked between the cost of the expanded crest or the cost of

raising the crest or the cost of lowering the stilling basin floor.

57
CdP User’s Manual February 2017 (second edition revised and expanded)

Hydraulic jump Froude Energy


Stilling basin type (USBR) Characteristics
classification number loss
No jump occurs (sub-
Fr1 < 1 ±0%
critical flow)
N jump occurs (critical y2 = ± 2*y1;
Fr1 =1 ±0%
flow) V2 = ± V1/2
-Two conjugate depths are close
- Flat floor with no appurtenances
1 < Fr1 <
- Extensive length to contain the jump ≤5%
1.7
- Costly due to length
Undular jump
Basin type I - No control on jump
- 2.0 < y2/y1 < 3.1
- Flat floor with no appurtenances
- Extensive length to contain the jump ≤ 15 %
- Costly due to length
Basin type I - No control on jump
1.7 < Fr1
- 2.0 < y2/y1 < 3.1
< 2.5
- Alternative for type IV or I
Weak jump - Chute blocks, baffle blocks and dentaded end sill
≤ 33 %
- Less simple for construction
- Decrease length of basin thanks to auxiliary
Basin type LFN devices

58
CdP User’s Manual February 2017 (second edition revised and expanded)

Hydraulic jump Froude Energy


Stilling basin type (USBR) Characteristics
classification number loss

- 2.0 < y2/y1 < 3.1


- Most common in our design
- Jump difficult to handle ≤ 45 %
2.5 < Fr1 - Try to change dimensions to be out of oscillating
Oscillating jump < 4.5 conditions and high d/s celerity waves
(entering jet Basin type IV
intermittently flowing
near bottom and -Recommended instead of type IV (refer to
Basin type LFN
along surface characteristics above)

- Incoming flow velocity V > 18 m/s


- No particular difficulties to handle
- No baffle blocks due to high velocity (cavitation)
- For large hydraulic structure

- 5,9 < y2/y1 < 12


Steady jump 4.5 < Fr1 Basin type II ± 70 %
(least sensitive in <9 - Incoming flow velocity V < 18 m/s (no cavitations)
terms of its position - qmax = 18 m3/s.m
to small fluctuations - Recommended because length basin type III <
in the tailwater length basin type II and type I (up to 60% of type I)
elevation) - Chute blocks, baffle blocks and end sill to
decrease length (baffle blocks can also be cube
shaped)
Basin type III

59
CdP User’s Manual February 2017 (second edition revised and expanded)

Hydraulic jump Froude Energy


Stilling basin type (USBR) Characteristics
classification number loss
- 12 < y2/y1 < 20
- High dam spillway with sloping glacis
- Bucket dissipates energy (and not the hydraulic
Fr1 > 9 ± 80 %
jump and basin because too large and expensive)
- When the tailwater depth is too great for the
Strong jump Basin type V – bucket formation of a hydraulic jump

1.7 < Fr1


Basin type SAF - Difficult to construct
< 17

- Drop < 6.1 m


- Basin dimensions determined with design chart
Straight drop with impact
- Impact block basin length < hydraulic jump drop
blocks
basin
2m≤ - Fr1 ≤ 4.5
drop ≤
6.1 m
- Drop < 6.1 m
Straight drop with hydraulic
- Basin dimensions determined with design chart
jump and stilling basin
- Stilling basin type II, III or IV

Figure 12.2: classification of hydraulic jumps and types of stilling basins according to the incoming Froude number Fr1 (USBR)

60
CdP User’s Manual February 2017 (second edition revised and expanded)

1. Total dissipater length according to the type of stilling basin and the
value of the Froude number.
Table 12.1: empiric relations for the calculation of the total basin length
according to the type of basin (USBR) and the value of the Froude number
(rectangular cross-section)
Fr1
Type Length of basin (m)
(dimensionless)
IV 2.5 < Fr1 < 4.5 5*(y2 – y1)
III 2.5 < Fr1 < 4.5 2.7*y2
II 4.5 3.6*y2
II 6 4.0*y2
II 8 4.2*y2
II ≥9 4.3*y2
LFN 1.7 < Fr1 < 4.5 3*y2
I 1.7 < Fr1 < 2.5 6*y2

2 Total dissipated length according to graphs plotted from experimental


data with the Froude number Fr1 against a dimensionless ratio L/y2).

Figure 12.3: dissipater length in terms of sequent depth y2 of jumps in


horizontal rectangular channels

61
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 12.4: dissipater length in terms of sequent depth y2 (type II, III & IV)

Figure 12.5: dissipater length in terms of sequent depth y2 (type LFN)

Figure 12.6: dissipater length in terms of sequent depth y2 (type II)

62
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 12.7: dissipater length in terms of sequent depth y2 (type III)

Figure 12.8: dissipater length in terms of sequent depth y2 (type IV)

In practice, the stilling basin is seldom designed to confine the entire length
of a free hydraulic jump on the paved apron because such a basin will be too
expensive. Consequently, accessories to control the jump are usually installed
in the basin (refer to figure 12.1).
12.4. STRAIGHT DROP

12.4.1. Introduction
The variables needed in the analysis for the inclined drop are much more
complex than that of the vertical drop. The drop structures discussed here
(refer to figures 12.1 and 12.9) require an aerated underside nappe to prevent
a pulsating and fluctuating jet and, in general, for sub-critical flow in the u/s as
well in the d/s channel.
A straight drop structure has a vertical wall between the control section
and the stilling basin. The small portion of energy loss occurs by impact of the

63
CdP User’s Manual February 2017 (second edition revised and expanded)

jet on the floor. The major portion of energy loss occurs by turbulence in the
stilling basin.
The hydraulic problem is concerned with the dissipation of flow energy in
the d/s basin. The aerated free falling nappe will strike the basin floor and turn
d/s. Beneath the nappe, a pool is formed that supplies the horizontal thrust
required to turn the nappe d/s (redirection of the flow). It is assumed first that
there is no loss in specific energy head between the reservoir and the point
where the jet strikes the basin floor (however, because of the impact of the
nappe on the stilling basin floor once the jet strikes the floor, some energy is
already lost). The main dissipation is obtained in the following hydraulic jump
(for straight drop with jump) or the turbulence induced in a basin with the
presence of impact blocks (straight drop with impact blocks).
12.4.2. Aeration of nappe
The negative pressure drags the lower side of the nappe towards the
surface of the cross-regulator wall
as shown in the opposite figure. If
the atmospheric pressure exists
beneath the nappe, it is known as
a free nappe. A free nappe is
obtained by ventilating a cross-
Figure 12.9: weir ventilation and regulator. With a partially
types of nappe: unaerated (clinging), ventilated nappe, the pressure
partially aerated (depressed) and below the nappe is negative
aerated ( free) nappe (depressed nappe). Sometimes, no
air is left below the water, and the
nappe adheres or clings to the d/s side of the cross-regulator (clinging nappe or
adhering nappe).
The discharge depends upon the amount of ventilation and the negative
pressure. Generally, the discharge of a depressed nappe is 6% to 7% more than that
of a free nappe. If the flow clings to the d/s side, reducing the head on the cross-
regulator (draw down), it gives a false value of discharge when it is put into the
formula. The discharge of a clinging nappe is 25% to 30% more than that of a free
nappe.
In a suppressed weir (weir width/Bstructure = 1), the sides of the structure may
prevent air from circulating under the nappe (aeration is automatic in a
contracted weir with weir width/Bstructure < 1). If used as a flow measurement,
the air beneath the nappe may be exhausted, causing a reduction of pressure

64
CdP User’s Manual February 2017 (second edition revised and expanded)

beneath the nappe with a corresponding increase in discharge rate for a given
head1.
12.4.3. Drop number
The flow geometry of the straight drop is described by functions of the
dimensionless drop number as shown in the following equation:

D = q2 / g * Y3 (12.1)
Where:
D = drop number (dimensionless).
Y = drop height (crest elevation minus cistern elevation) (m).
Refer to list of symbols.
12.4.4. Basin elevation
The first approximation of the basin elevation for various basin widths is
given in the design chart of the following figure. Because the basin is directly
d/s from the crest, there is no loss in specific energy head between the reservoir
and the point where the jet strikes the basin floor (scale A).
12.4.5. Straight drop with hydraulic jump
The aerated free-falling nappe of the jet strikes the floor of the stilling basin,
reverses its curvature and turns into super-critical flow in the apron of the
stilling basin. Consequently, a hydraulic jump can be formed d/s on the flat
apron as determined by the tailwater depth (refer to figure 12.11).
The jump characteristics of the straight drop are basically the same as those
for other jumps except that the position of the start of the jump cannot be
determined as readily as it can for other basins. The position of the depth y 1
can be approximately determined by a straight line from the point on the axis
of the nappe at the height of the pool depth joining the apron.
The value of y1 and of Fr1 at the start of the jump in relation to the drop
number D are shown in the below figures. Theses relations are used in the
User’s Manual to determine the basin dimensions and characteristics. If the
Froude number does not fall within an acceptable range, changes in either drop
height or in the unit discharge should be considered.
12.4.6. Straight drop with impact blocks
The impact block type is considered when Fr is < 4.5. The dissipation of
energy of the vertical drop structure with impact blocks is principally by
turbulence induced by the impingement of the incoming flow upon the impact

1 Suppressed weirs must have proper ventilation of the cavity underneath their nappes. This ventilation is
commonly done by installing properly sized pipes in the walls to vent the cavity under the nappe. Standard
equations and tables are valid only when sufficient ventilation is provided. The design of pipe size to introduce
sufficient air depends upon the discharge, drop, and the loss of accuracy that is tolerable. Sizing air piping and air
vents requires some knowledge of fluid mechanics and is difficult to do (outside the scope of the Manual).

65
CdP User’s Manual February 2017 (second edition revised and expanded)

blocks (refer to figure 12.11). The required tailwater depth is more or less
independent of the drop height.
The length is considerably smaller than the straight drop with hydraulic
jump. However, the foundation for an impact blocks basin must be of better
quality because of the concentrated forces involved.
In practice, the straight drop with stilling basin has a much wider application
than the straight drop with impact blocks.
12.5. END SILL

An end sill is effective in spreading the flow. The higher the end sill, the more
effective it will be in spreading the jet. A higher end sill results in shallower
depths in the d/s channel and possibly higher velocities over the d/s floor
protection or channel bed.
The end sill should not be appreciably above the exit channel. For a sub-
critical approach flow, the depth of flow is reduced. The flow drop may be
enough to generate a critical flow transition and to form a secondary jump d/s.

66
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 12.10: USBR stilling basin depths versus hydraulic heads

67
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 12.11: USBR hydraulic characteristics of straight drops with hydraulic


jump or with impact blocks

68
CdP User’s Manual February 2017 (second edition revised and expanded)
102.00 Section A (approach) Section C-C
100.00
EL 102.70 Section 1-1 Section 2-2 Section 3-3
Martin, the Supervisor EL 100.50

HE0 = 101.76 HE1 = 101.76


2
101.53 Va /2g = 0.23
101.32 HE1 = 100.09

RIVER or
PARENT CANAL h0 max = 1.23 yc = 1.02 ∆Edrop = 2.19
95 m3 dumped stones Va = 2.12 H0 = 1.46 Vc = 3.17 ∆Edrop = 2.31

Φ min. 0.05 El. of reference level 100.30 Top weir length = 4.00

EL 100.00 1V:1H E1 = 3.96 HE2 = 99.57


Thick. = 0.60 E1 = 2.91 HE2 = 99.45 h2 = -0.85
d1 = 2.50 Φ EAc =

0.60 75.04 Zc = 2.94 ∆Ejump = 1.77 99.42

EL 97.50 Weir width = 25.00 Zc = 3.12 ∆Ejump = 0.65 Design jump 99.45
y2 = 2.06
v2 = 1.57 EL 99.15 yn (rating curve) = 1.95
Depressed jump

292 CC blocks of 0.60 x 0.60 x 0.60 y2 = 2.15 Basin type IV

32 m3 packed stones 1 97.77 97.65 E2 = 2.19 v2 = 1.51 V3 = 1.23


1 E2 = 2.27
df = 1.56 y1 = 0.41
97.39 y1 = 0.47 Field trial of Arnaud the Designer 417 CC blocks of 0.60 x 0.60 x 0.60
97.68
1 H; 0.5 V hes = 0.18 97.50
EL 97.00 0.50

Thick. = 1.00 135 m3 dumped

Thick. = 1.20 Inverted filter stones


Thick. = 2.00 Thick. = 1.80 d2= 3.00 Φ min. of 0.06
95.00 95.20 Φ halfway = 36.85
95.00 Φ CAc = 70.92 Φ t = 47.86 0.60

9.30 94.50

3.50 3.50 3.00 4.30 Ld = 5.30 12.49 Vertical drop + basin 5.00 5.00

Lp = 5.60 LB - Lp = 2.61 Vertical drop + blocks

Hydraulic jump type of vertical drop b = 30.40 m including cut-offs)


Impact blocks type of vertical drop b = 21.11 m (including cut-offs)

Figure 12.12a: spreadsheet showing the general scheme for the hydraulic and structural design of a drop structure with
straight drop

69
CdP User’s Manual February 2017 (second edition revised and expanded)

103.0
Elevation (m)

101.88
101.88 101.88 101.88 101.88 101.88
102.0 101.58
101.69
101.69 101.69
101.69 101.29
101.31 101.00
101.0
100.30 100.30 100.41
100.00
100.00 100.12
100.70 99.82 d/s river bed elevation with structure
100.0
99.53
99.24
98.94
99.1599.15
99.0 98.97
98.78 98.81 98.81
98.40
98.22
98.0 97.94 97.68
97.84 97.50 97.50
97.18 97.18 97.18 97.18

97.0

96.0
0.0 5.0 10.0 15.0 20.0 25.0 30.0 35.0 40.0
Hydraulic grade line (water surface el. of jump) (m) Total energy line or energy grade line elevation (m) Elevation of structure floor (m) Critical depth Distance (m)

Figure 12.12b: spreadsheet showingfloor,


FIGURE Be: cross-regulator structure floor
hydraulic jump, critical(black), the
and total energy line water
(SI units) surface of jump (blue and green), the critical depth
Dessinque river (dotted
Proutheadred)
work (CR + drop) August 2015
and the energy grade line (Printed:
(red)16-Sep-16 )
of a drop structure

70
CdP User’s Manual February 2017 (second edition revised and expanded)

13.HYDRAULIC JUMP WATER SURFACE ELEVATION IN THE BASIN


13.1. INTRODUCCION

The knowledge of the free water surface profile of the jump is necessary to:
 Design the freeboard of the side walls of basin.
 Determine the water weight in the basin to counteract the uplift
pressure under the structure.
In the User’s Manual, the profile of the water surface of the hydraulic jump
is determined for the high flood flow Qmax without taking into account the
concentration and degradation factors.
13.2. JUMP WATER SURFACE ELEVATION IN THE SUB-CRITICAL REACH

The curves for plotting the post jump profile for different values of F r1 are
given in the below figure

Figure 13.1: design chart for plotting post jump water surface profile in the
stilling basin

By taking arbitrary values of x along the profile after the point of jump
formation, x/y1 can be tabulated. Corresponding to these values of x/y 1 for a
known Fr1, different values of y2/y1 can be read out1 of the design chart. Hence,
different values of x and y2 are known and the post jump profile can be slowly
plotted.

1 y from abacus corresponds to the sub-critical sequent depth of flow y2 in the User’s Manual.

71
It should be noted that the considered jump surface elevation occurs when
the u/s flow has a Froude number of 2.5 or more. At lower F r value, the jump
tends to be either undular in nature or weakly developed (refer to figure 12.1).
13.3. JUMP WATER SURFACE ELEVATION IN THE SUPER-CRITICAL REACH

The water surface profile in the super-critical conveyance reach before the
base point formation can easily be plotted with the help on the Montague’s
design chart (refer to figure 11.2).
At different points on the glacis corresponding to increasing values of x, the
specific energy E1 will go on increasing, being equal to the u/s specific energy
level minus glacis level at the considered point (refer to table 25.1 with ΔZ ≠ 0
and B2 = B1).
For each value of E1, a corresponding value of y1 can be estimated from the
curves. The values of y1 will go on reducing till reaching the surface point of the
jump formation.

72
PART V:

DESIGN OF CANAL HEAD REGULATOR AND


FLOW CONDITIONS

73
CdP User’s Manual February 2017 (second edition revised and expanded)

14.WATER REQUIREMENT AND CONTROL AT HEAD REGULATOR


14.1. WATER SUPPLY

Water level and discharge control and measurement are required at all off-
taking (branching) canals, as to control the discharge that leaves the parent
canal or the river.
The discharge control method generally used is the supply oriented
operation or d/s control. On the contrary, u/s control takes place when the
water is released from the head work and distributes over the canals 1.
14.2. IRRIGATION REQUIREMENT AND DESIGN DISCHARGE

The required irrigation flow rate in the canal is calculated according to the
cropped area, the water requirements of each crop and the effective rainfall.
The maximum irrigation water need adopted in the river basin is 2.51 l/s/ha
for the design calculation, based on local practices and field losses.
The design of the head regulator (also called discharge regulators) considers
the gate(s) fully open and takes into account the maximum flow rate passing
down the cross-regulator in the river or the parent canal (h0 max or H0 max).
14.3. GATE CHAMBER

In the User’s Manual, the gate chamber comprises the piers and gate(s)
system (open channel or breast wall type) and performs usually as an
obstruction2 in the open channel. The gate chamber is considered with a
nominal horizontal (or nearly so) slope with or without the presence of a sill
(weir)3 below the gate (refer to article 16.4).
14.3.1. Pier
Piers are provided between each bay (refer to figure 2.2). The effect of the
piers is to contract the flow and, hence, to alter the effective crest width of the
weir. The piers are usually of mass concrete and founded on the floor.
Nominal reinforcement shall be provided at the faces as protection against
surface cracking.
14.3.2. Under-flow gate
One spindle under-flow (under-shot) gate such as sluice gate is the most
common type of gate.

1 As opposed to a demand oriented operation or downstream control that maintains a constant water level at
the d/s side of the structure, without regarding discharges. The effect is that the discharge at each regulator is
automatically adjusted to the accumulated d/s demand for irrigation water.
2 An obstruction in open channel flow presents a phenomenon very similar to that of a constriction, since both

have the effect of reducing the cross-sectional area of the flow. However, the constriction reduces the cross-
section into a single opening, whereas the obstruction creates at least two openings.
3 The presence of a sill under the gate reduces the gate size and sediments flowing d/s. Water depth can be

affected (refer to table 25.1).

74
CdP User’s Manual February 2017 (second edition revised and expanded)

The illustration of terminology of a sluice gate is given by the following


figure:

Figure 14.1: illustration of the sluice gate terminology with mild slope or
horizontal floor

The advantages of the vertical gate with under-flow as a discharge or as a


water level regulator are: the structure is relative cheap, simple and sturdy; it
provides a reasonable constant discharge for varying u/s water levels in the
parent canal.
The disadvantages of regulators with vertical gates are: the gate does not
pass floating debris (refer to figure 16.4) and the discharge cannot be always
measured accurately (refer to article 17.5.4).
14.3.3. Approach flow velocity
The velocity to be considered is the approach (or accelerated) velocity Va at
the entry of the structure (Manning’s equation). For solving the problems in
hydraulic engineering, the velocity used is the average velocity of flow over a
section.
Permissible limits of velocity are given in the table 6.1. However, the
conveyance reach u/s of the chamber should have a sub-critical flow ("normal"
flowing water) as to limit the head loss. It means that Fr < 1 (refer to article 6.4.2
and equation 6.2a)).
14.4. EFFECTIVE OR CLEAR WATERWAY

14.4.1. Pier(s) contraction


For a gated structure, the effective waterway Bcl has to be taken into
consideration instead of the structure waterway Bstructure between abutments
in order to account for the obstruction (horizontal contraction) due to the
presence of the pier(s) of the gate chamber (refer to figure 2.2).

75
CdP User’s Manual February 2017 (second edition revised and expanded)

This horizontal contraction of the flow alters the effective crest width
(transverse to the water flow) of the gated
structure. A weir in which the crest width equals
the channel width is referred to as suppressed and
the number of side contractions N = 0. If both sides
of the weir are far enough removed from the sides
of the approach channel (end contraction of the
weir), the weir is considered to be contracted (or unsuppressed), and N = 2; if
one side is suppressed and one is contracted (or unsuppressed), N = 1.
The effective or clear width of the crest is given by the following equation:

Bcl = Bstructure -Wpier- K * Nsc * h0 (14.1)


Where:
K = pier contraction coefficient.
Wpier = total with of piers (transverse to the water flow) (m).
Nsc = number of side contractions (2 for each gated bay; 0 for
suppressed side contraction weir).
Refer to list of symbols.
The pier contraction coefficient K varies mainly with the shape and position
of the pier nose and the head conditions. In the User’s Manual, the u/s nosing
and d/s end of pier are curved to ease the flow. Values of the contraction
coefficient are given in the following table for different pier shapes:
Table 14.1: pier contraction coefficients and shapes
Pier shape K
Square nosed pier without any rounding 0.10
Square nosed pier with rounded corners 0.02
Rounded nose pier 0.01
90° cut nosed pier 0.01
Pointed nose pier 0.04

With one open gate adjacent to closed gates, these values of K become
roughly 2.5 times larger.
While designing the head regulator, the open water width of the structure
is chosen according to the obstruction as shown in the following table:
Table 14.2: obstruction and waterway
Flow Open water width Hydraulic
Obstruction
regulation Name Width coefficients
Same or less than
None Unregulated Bstructure Fairly constant
channel or river
Unregulated Bcl

76
CdP User’s Manual February 2017 (second edition revised and expanded)

Bstructure minus
Related to pier(s)
Pier(s) & obstruction
shape (horizontal
bay shape effects of pier(s)
contraction)
in gate chamber
Related to pier(s)
Bstructure minus
Pier(s) & shape and gate(s)
obstruction
bay shape Regulated Bcl lip and position
effects of pier(s)
& gate(s) (horiz. and vertical
in gate chamber
contractions)

However, a minimum width of the off-take has to be considered to avoid


super-critical flow at the entrance and the possibility of an u/s jump (refer to
figure 14.4b and table 14.3).
14.4.2. Pier length and critical depth
In the User’s Manual, the minimum length of the pier is equal to 2 m, taking
into account the gate chamber and the 1 m minimum width operating platform;
the d/s edge or brink of the pier corresponds to the beginning of the glacis (u/s
inflexion point).
With the presence of pier(s), the calculation of the critical depth yc takes
into account the effective or clear waterway B cl and is given in the following
equation.
yc = (q2 / g)1/3 (14.2) with q = Q / Bclear = y * V (14.3)
Where:
Refer to list of symbols.
Therefore, the d/s edge or brink of the pier must correspond to the
beginning of the glacis (u/s inflexion point) from which B structure is taken into
account in the calculations.
14.5. GATE OPENING, ENERGY AND MOMENTUM

Under-flow gates operate in a variety of flow modes including the weir flow
when the gate(s) are out of the water and the orifice flow for normal operation.
The following figure of the hydraulic jump d/s of a gate outlet gives a clear
idea about how conservation of energy and conservation of momentum apply
in a gate chamber with rectangular cross-section and a constant unit discharge
q (refer to article 9.3).

77
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 14.2: hydraulic jump, specific energy and specific force diagrams of an
under-flow gate with rectangular cross-section

As shown in the centre of the above figure, a deep u/s flow (position 1)
encounters a sluice gate. The gate imposes a decrease in flow depth at position
2 and a hydraulic jump occurs between the positions 2 and 3 (note that the gate
opening w is lower than the critical depth yc).
The left part of the figure shows the specific force diagram of these three
positions, while the right part shows the specific energy diagram for these same
three positions. The energy loss can be neglected between the positions 1 and
2 (E1 = E2)1, but the external thrust on the gate causes significant specific force
loss (Fs1 > Fs2). By contrast, between the positions 2 and 3, turbulence in the
hydraulic jump dissipates energy (E3 < E2), while the momentum is assumed to
be conserved (same specific forces Fs2 = Fs3), provided the channel bottom slope
or the datum are the same.
In other words, the energy-momentum applies with the energy equation
from position 1 to position 2 and the momentum equation from position 2 to
position 3.
If we know the unit discharge q and the flow depth at position 1, by applying
energy conservation between the positions 1 and 2 and momentum
conservation between 2 and 3, the flow depths at position 2 (y 1) and 3 (y2) can
be computed.
From the above figures, it can be inferred:
 Total energy line (total mechanical energy): neglecting friction
losses (short distance), there is no energy loss as water flows
under the gate (refer to opposite figure) and the Bernoulli’s
equation applies 2. This is reasonable because the flow is

1 The specific energy remains constant for idealized gate(s) with negligible frictional effects.
2 On the contrary, in the hydraulic jump, momentum principle applies (refer to article 10.3.4).

78
CdP User’s Manual February 2017 (second edition revised and expanded)

converging (contracting) under the gate and this tends to


suppress turbulence which means
little or no energy loss (turbulence is
the main cause of energy loss).
 There is a significant change in the
components of the total mechanical
energy across the gate even though
the total is the same. Upstream flow is
Figure 14.3: total energy line (or slow and deep whereas d/s flow is very
energy grade line) of a sluice gate shallow and fast but the discharge is
the same on both sides of the gate.
 Energy dissipation d/s of gate(s): it takes place in the hydraulic
jump1. The hydraulic jump has a tendency to move towards the
gate because the normal depth yn in the d/s channel is usually
greater than the sequent depth y2 (refer to article 10.5.3).
 Energy dissipation u/s of gate(s): with a super-critical flow u/s of
the gate(s), the transition from super-critical to sub-critical flow
takes place u/s of the gate(s). A hydraulic jump occurs before the
gate(s) (refer to figures 14.4b and 14.4c). No hydraulic jump
occurs d/s of the gate because the flow is going from one super-
critical flow to another super-critical flow (steep slope); there is
no d/s transition between super-critical and sub-critical flow.
14.6. WATER SURFACE PROFILES WITH GATE

14.6.1. Introduction
The following figures show the different water surface profiles encountered
with the presence of a sluice gate with horizontal, mild, critical, and steep slope:

(a) Mild slope (b) Steep slope

1 It is usually not a well-formed jump. The basin often has entering Froude number less than 4, which means that
the jump can be weak or oscillating between the bottom and the water surface, resulting in irregular wave
formation propagating downstream.

79
CdP User’s Manual February 2017 (second edition revised and expanded)

(c) Critical slope (d) Horizontal slope


Figure 14.4a, b, c, d: individual water surface profile with a gate

As shown in the above figures, the critical depth remains the same in the
channel because it depends only on discharge and not on slope.
Generally of most interest to the design engineer is the situation where GVF
occurs as a result of an obstruction (fallen tree, gates, weirs etc.) that raises the
water surface above the uniform flow normal depth line u/s of the obstruction 1,
as shown in the following figure (refer also to figures 25.1a & 25.1b):
This flow profile
represents a backwater curve
as the depth of flow increases
continuously in the direction of
flow. This happens only once
critical conditions are achieved
Figure 14.5: water surface profile due to on the sill in the gate chamber
obstruction (refer to article 25.2.1).
14.6.2. Surface flow profiles encountered with a gate
The following table and figures give the type of curves illustrated in the
above figures that can be generated by a sluice gate placed across the channel,
so that it produces a super-critical flow under the gate.
Table 14.3: description of flow profiles encountered with the presence of a sluice
gate with regulated flow
Profiles Description
Sub-critical flow. Represents the backwater curve on a mild slope M (for
M1
example, occurs u/s of a gate)
Sub-critical flow. Represents the drawdown curve on a mild slope M (for
M2
example, occurs u/s of a sudden enlargement or overfall)
Super-critical flow. Starts at a vertical angle slope and terminates with a
M3
hydraulic jump2

1 When analysing the GVF with the Manning equation, the longitudinal energy slope must be used instead of the
longitudinal bed slope.
2 In other words, M is the gradually varied flow (GVF) curve leading to the alternate depth y , followed by the
3 1
rapidly varied flow (RVF) leading to the sequent depth y 2 (for example, occurs when a super-critical flow enters a
mild slope or d/s of a sluice gate)

80
CdP User’s Manual February 2017 (second edition revised and expanded)

(reach with mild slope M)


The lowering of the gate on a sub-critical flow in a horizontal or mild reach frictionless
rectangular section produces 3 surfaces profiles: M3 immediately d/s of the gate
which then undergoes a hydraulic jump to sub-critical M2 GVF curve which depends
upon the d/s control (overfall here). It could be a M1 profile also (instead of the local
overfall control, a very long channel gives a friction control ensuring uniform flow
with the normal depth being controlled by the friction of the channel). Upstream of
the gate, the initial/s portion of a M2 curve evolves to an M1 curve as the lowering
the gate progresses. This M1 curve represents in fact the energy built up in front of
an obstruction and M3, the energy dissipation.

Sub-critical flow. Begins with a hydraulic jump at the u/s and becomes
S1
tangent to the horizontal pool level at d/s end.
S2 Super-critical flow. Drawdown curve.
Super-critical flow. Formed between an issuing super-critical flow and the
S3 normal depth line to which the profile is tangent (for example, d/s of a
sluice gate in a steep slope when the gate opening w < yc).
J Hydraulic jump.

(reach with steep slope S)


The lowering of the gate (artificial control) on a super-critical flow in a steep reach
frictionless rectangular section backs up water to form a S1 profile on the u/s side
with the formation of a jump.

81
CdP User’s Manual February 2017 (second edition revised and expanded)

14.7. FLOW BEHAVIOURS THROUGH A GATE

The flow behaviour associated with gated structures (open channel or


breast wall type) is often complex under real-time conditions and the rating
curve is not straight forward.
14.7.1. Regulation and flow types
Hydraulically speaking, the under-flow gate(s) of the gate chamber acts like
an orifice when regulating the flow1 (refer to articles 16) and like a non orifice
when not regulating the flow 2 (refer to articles 15). With unregulated flow, a
critical water level control section may or may not exist in the chamber,
depending of the magnitude of the specific energy at the contraction.
Basically, there are three types of flow behaviours through a gated
structure, as shown in the following table. The identification of the flow type is
crucial for the selection of the correct equation to be applied.
Table 14.4: flow behaviours through a structure with a gated discharge
regulator
Regulation Types Remarks
Rectangular broad or short crested weir
state if presence of gate sill; horizontal
Unregulated (non Free
constricted or obstructed open channel
orifice flow)
flow if no gate sill (Fr > 1)
Submerged Tailwater quite high (Fr < 1)
Free Outflowing jet open to atmosphere (Fr > 1)
Partially Outflowing jet still super-critical (Fr > 1);
Regulated (orifice
submerged some influence of the tailwater
flow)
Fully Outflowing jet sub-critical (Fr < 1); influence
submerged of the tailwater

14.7.2. Pivot table


The different criteria used in the Manual for flow regulation and for distinguishing
the conditions, types and state of flows below a gate or over a weir are summarized in
the following pivot table:

1 The gate bottom interferes with the water flow.


2 The gate bottom is above the water surface.

82
CdP User’s Manual February 2017 (second edition revised and expanded)

Table 14.5: flow regulations, types, conditions and statess in relation to the adopted pivot
Pivot
Flow Non orifice or Reference
Orifice flow
weir flow
Regulation
Regulated h0 > 1.1*w Not applicable
Articles 15
Unregulated Not applicable h0 ≤ 0.9*w
& 16
Semi-regulated 0.9*w ≤ h0 < 1.1*w
Types
h2/w < modular h2/h0 < 75%
Free
limit*0.9
h2/w > modular h2/h0 > 75% Articles
Submerged
limit*1.1 14.7 &
0.9*modular limit 16.2.2
Transitional < h2/w <
1.1*modular limit
Conditions
Orifice h0 > w Not applicable
Articles 15
Non orifice or weir h0 ≤ w
Not applicable & 16
flow
State
Cannot be used as
H0/L < 0.08
broad crested Not applicable Article 6.4
H0/L > 0.33
measuring device
Can be used as 0.08 ≤ H0/L ≤ 0.33
broad crested Not applicable (0.08 < H0/L < Article 6.4
measuring device 0.15)

83
CdP User’s Manual February 2017 (second edition revised and expanded)

102.00 Albert, Supervisor 4 gate(s) 1.20 m width & 1.36 m height 102.00
Dessinque river

3 pier(s) of 0.80 m width X 1.75 m length and 103.00 m elevation

Prout head work (off-take) Bottom el. of operating platform 103.00 Top el. of operating platform 103.20
EL 103.00 Section 1-1 Section 2-2 Section 3-3
EL 101.90
HE0 = 101.83 ∆E = 0.46
2
Pond level 101.76 Va /2g = 0.07 101.76 E1 = 2.44 HE2 = 101.36 100.95
or HFL E1 = 1.90 ∆E = 0.87 HEchannel = 100.96
For Qmax =

80.04

Va = Vc = 3.11 V1 = 5.50 101.30


Design jump
1.14 h0 max = 1.36 yc = 0.99 V1 = 6.47 y2 = 1.33 EL 100.89
13 m3 dumped stones H0 max = 1.43 V2 = 1.51 EL 100.74
El. of reference level 100.40 Depressed jump
d50 min. of 0.07 100.29 h2 = 0.49

EL.100.00 E2 = 1.44 y2 = 1.48 yn (from rating curve) = 0.92


Thick. = 0.60 1 y1 = 0.36 E2 = 1.57 V2 = 1.35 V3 = 1.17
2 99.92 99.70 Basin type IV Field trial of Louis the Designer
100.17 133 CC blocks of 0.60 x 0.60 x 0.40

0.60 d1 = 1.20 u/s horizontal weir slope = 1.00 hes = 0.20 EL 99.97
99.39 1 H; 0.5 V
EL 98.80 Bstructure = 7.20 y1 = EL 99.26 0.71
50 CC blocks of 0.60 x 0.60 x 0.40 Bcl = 4.69 0.31 Thick. = 0.80
34 m3 dumped
5 m3 packed stones flow is NOT reasonably uniform; single row of triangle Inverted
Thick. = 1.20 stones
chute blocks (glacis blocks) NECESSARY at the glacis toe filter
(User's Manual, figure 10.5) 97.47 0.60 d2 = 2.50 d50 min. = 0.09

1.50 1.50 7.86 2.15 horizontal length) 2.28 (horizontal length) 8.41 (horizontal length) 4.00 4.00
Min. distance for measur. station = 2.86 Crest length = 1.75 2.55 (slope length)

Pier(s) length = 2.00

b= 21.30 (horizontal length including cut-offs)

Figure 14.6a: spreadsheet showing


FIGURE C2: head
Dessinque river
the general
reguator longitudinal section (intake scheme ofAugust
the2016
or off-take structure)(sluice
Prout head work (off-take)
hydraulic
gate(s) not shown) and
(printed:
(SI units)structural
(not to scale)
16-Sep-16 )
design of a gated head regulator

84
gate(s) 1.20 m width & 1.36 m height 102.00
CdP User’s Manual February 2017 (second edition revised and expanded)
pier(s) of 0.80 m width X 1.75 m length and 103.00 m elevation

op el. of operating platform 103.20


Section 1-1 Section 2-2 Section 3-3
EL 101.90
∆E = 0.46
E1 = 2.44 HE2 = 101.36 100.95

E1 = 1.90 ∆E = 0.87 HEchannel = 100.96

V1 = 5.50 101.30
Design jump
V1 = 6.47 y2 = 1.33 EL 100.89
V2 = 1.51 EL 100.74
0 Depressed jump
100.29 h2 = 0.49

E2 = 1.44 y2 = 1.48 yn (from rating curve) = 0.92


y1 = 0.36 E2 = 1.57 V2 = 1.35 V3 = 1.17
2 99.92 99.70 Basin type IV Field trial of Louis the Designer
100.17 133 CC blocks of 0.60 x 0.60 x 0.40

.00 hes = 0.20 EL 99.97


99.39 1 H; 0.5 V
y1 = EL 99.26 0.71
0.31 Thick. = 0.80
34 m3 dumped
triangle Inverted
Thick. = 1.20 stones
he glacis toe filter
97.47 0.60 d2 = 2.50 d50 min. = 0.09

th) 2.28 (horizontal length) 8.41 (horizontal length) 4.00 4.00


2.55 (slope length)

Figure 14.6b: spreadsheet showing a detail of the general scheme of the hydraulic and structural design of a gated head
regulator
b= 21.30 (horizontal length including cut-offs)

ake structure)(sluice gate(s) not shown) (SI units) (not to scale) 85


ke) August 2016 (printed: 16-Sep-16 )
CdP User’s Manual February 2017 (second edition revised and expanded)

102.0101.83 101.83
101.83 101.83 101.83101.83
101.75
Elevation (m)

101.67
101.57 101.59
101.48 101.51
101.39 101.43
101.5 101.35
101.28
101.20
101.12 Tail water elevation
101.04
100.96
100.89 100.89
101.0
100.74
100.59 100.59

100.40
100.5

100.09 100.17
100.20 100.04
100.00 99.97 99.97
100.0
99.70 d/s canal bed elevation with structure
99.59 99.57

99.5
99.39 99.26 99.26 99.26

99.26 99.26 99.26


99.0 99.26
0.0

5.0

25.0
10.0

15.0

20.0

30.0
Water surface elevation of depressed jump (m) Total energy line or energy grade line elevation (m) Elevation of structure floor (m) Critical depth (m) Distance from u/s structure (m)

Figure 14.6c: spreadsheet showing floor,


FIGURE C2: cross-regulator structure floor
jump profile, (black),
critical the
and energy water
grade line withsurface
flow of jump (blue and green), the critical depth
(dotted red)
Dessinque river Proutand the (off-take)
head work energy Augustgrade
2016 line (red) of16-Sep-16
(printed: a gated
) head regulator

86
CdP User’s Manual February 2017 (second edition revised and expanded)

15.GATE(S) OPENING WITH NON ORIFICE FLOW CONDITIONS


Since the gates are used for flow distribution, the gate opening is, at some
height, equivalent to a non orifice flow.
The hydraulic coefficients used in the discharge equations account for:
 Approach velocity head.
 Approach velocity distribution.
 Decrease in jet velocity caused by friction.
 Amount of jet contraction caused by the flow curving around the
corner of the orifice perimeter.
 Flow states.
15.1. CASE Nº1: GATE LOWERED INTO A RECTANGULAR HORIZONTAL CANAL
ABOVE FLOW

15.1.1. Outflow conditions with h0 < w


When the gate is kept above the u/s water head ho (in relation to weir sill if
existing), the gate(s) is not able to regulate the flow (refer to figure 16.8).
The flow through a one-gate chamber performs hydraulically as an open
channel flow with constriction (refer to article 25.2); in case of more than one
gate, the chamber performs as an open channel flow with obstruction.
15.1.2. Outflow discharge equation for non orifice conditions
The flow beneath the gate is described by the horizontal broad crested or
short crested weir equations (refer to equation 6.1a and figure 6.1).
Free flow type
With fully open gate(s) (refer to figure 16.8), the free water discharge
available through the gated bay(s) (with pier if more than 1 gate) is given by the
discharge formula of a rectangular weir as shown below:

Q = Cd * Cv * Bcl * h03/2 (15.1)


Where:
Q = discharge through gate(s) (m3/s).
Refer to list of symbols.
The product of Cd * Cv is called the effective discharge coefficient C e (refer
to articles 6.1). For practical reasons, the energy head H0 is replaced by the
water head h0 (refer to equation 6.1b and article 6.1.1).
Submerged flow type
In case of submerged flow, the head discharge relationship can no more be
applied and the d/s water surface elevation must also be measured.

87
CdP User’s Manual February 2017 (second edition revised and expanded)

With fully open gate(s) (refer to figure 16.8), the water discharge available
through the gated bay(s) section is given by the discharge formula shown
below:
Q = Cd * Cv * Bcl [2 * g (ho – h2)]0.5 (15.2)
Where:
Q = discharge through gate(s) (m3/s).
Refer to list of symbols.
The product of Cd*Cv is called the effective discharge coefficient Ce (refer to
article 15.1.3 and 15.1.5).
Submerged flow in the structure will always decrease accuracy of flow
measurement (refer to article 17.5.4).
In the User’s Manual, discharges under drowned conditions are also
obtained by applying a flow reduction factor f to the free flow discharges
(equation 15.1) as shown below:

Qsubmerged = f * Qfree (15.3)


Where:
Qsubmerged = submerged discharge through gate(s) (m3/s).
Qfree = free discharge through gate(s) (m3/s).
f = flow reduction factor.
This correction factor f depends upon the u/s and d/s slope of the weir faces
as pictured below and the submergence ration as shown in the following table:

Figure 15.1: u/s or d/s vertical or sloping weir faces of off-taking canal

Table 15.1: flow reduction factor for submerged (non modular) flow with
broad crested weir
Submergence
Weir shape (H:V) f
ratio
u/s or d/s vertical or sloping weir faces ≤ 0.75 1.0
0.80 0.95
0.85 0.88
u/s & d/s vertical faces
0.90 0.75
0.95 0.57
0.80 ≈1

88
CdP User’s Manual February 2017 (second edition revised and expanded)

0.85 0.98
u/s face 1:1 & d/s face 2:1 0.90 0.90
0.95 0.73
0.80 ≈1
0.85 0.95
u/s face 5:1 & d/s face 2:1
0.90 0.82
0.95 0.62

Transitional flow type


In case of transitional type of flow, the flow is considered as semi-regulated
and can be free or submerged according to the d/s conditions. The transitional
flow type is calculated as orifice flow (ho > w) or non orifice flow (ho ≤ w) in the
User’s manual (refer to table 14.5, conditions).
15.1.3. Effective discharge coefficient
The effective discharge coefficient depends upon different parameters such as
the u/s water depth, the contraction coefficient and the approach velocity.
In practice, for free flow type, Ce is taken as 1.7 for rectangular broad
crested weir state or 1.84 for short crested weir state (refer to article 6.1.2).
15.1.4. Contraction coefficient
In the case of non orifice flow, the contraction corresponds to the amount
of curvature of the jet in the direction of flow due to the geometry of the piers
(horizontal contraction) and is reflected in the effective waterway Bcl
(horizontal contraction) (refer to tables 14.1 and 14.2).
15.1.5. Velocity coefficient
In the User’s Manual, the practice to calculate the discharge rate above the
weir is by measuring h0 instead of H0, taking into account that the u/s water
surface elevation in the canal equals the u/s water surface elevation in the river
or parent canal. By using h0 instead of Ho, the correction coefficient Cv for
neglecting the velocity of approach head has to be introduced with Cv is ≥ 1
(refer to article 6.1.3).
However, in irrigation systems, Cv can be assumed to be unity since most of the
canals have a very flat hydraulic gradient and the flow velocities very low.
If significant u/s approach velocity occurs, Cv may be approximated from the
following equation for a rectangular cross-section:

Cv = (H0 / h0)1.5 (15.4) (6.1d)


Where:
Refer to list of symbols.

89
CdP User’s Manual February 2017 (second edition revised and expanded)

16. GATE(S) OPENING WITH ORIFICE FLOW CONDITIONS


The gate(s) installed in the chamber creates orifice flow conditions when
the edge of the gate(s) goes under the water surface. Between regulated and
unregulated flow, the flow type is semi-regulated: the edge of the gate(s)
touches or slightly goes under the water surface (refer to table 14.5).
The hydraulic coefficients used in the discharge equations account for:
 Approach velocity head.
 Approach velocity distribution.
 Decrease in jet velocity caused by friction.
 Amount of jet contraction caused by the flow curving around the
corner of the orifice perimeter.
 Flow state.
16.1. CASE Nº2: GATE(S) LOWERED INTO A RECTANGULAR HORIZONTAL
CANAL TO A HEIGHT BELOW CRITICAL DEPTH WITH FREE FLOW TYPE
D/S OF THE GATE(S)

16.1.1. Flow description downstream of the gate(s)


When the gate(s) is progressively lowered (h0 ≥ 1.1* w), the velocity of flow
d/s of the gate(s) becomes larger and larger and, eventually, super-critical flow
will occur immediately d/s of the gate(s) with the formation of a hydraulic
jump1 when the initial and sequent depths y1 and y2 and the approaching
Froude number satisfy the equation 10.4 (refer to figures 14.1 and 14.3).
Provided the jump has not reached the gate (submerged flow), the water depth
u/s of the gate remains constant due to the super-critical flow d/s of the gate.
In other words, the control is provided by the structure itself and the u/s
conditions.
The tailwater depth yn however plays a very influential role on where the
jump will occur in the channel, and changes in its depth can shift the jump
either u/s or d/s (refer to table 10.1).
If the gate opening w is reduced too much, it can represent a temporary
choke condition since the energy may not be sufficient to pass the required
amount of discharge per unit width. An increase in u/s water depth occurs (with
a surface profile M1; refer to table 14.3 and figure 14.4) with an instantaneous
reduction of the unit discharge 2 due to the change in geometry.

1
Super-critical outflow below the gate occurs as long as the roller of the hydraulic jump does not submerged the
section of minimum depth of the jet that is located at a distance from 1.54 to 1.60 the height of the gate opening.
2 After an instantaneous reduction in unit flow rate q due to the closing of the gate, q steadily increases, eventually

reaching the same steady state discharge as before the movement (closing) of the gate (the time that it takes to
get back to steady state discharge can be calculated). With the discharge Q remaining constant, the discharge per
unit width q within the contraction must increase again to reach the initial q, so that the critical depth yc must
also increase. Therefore, the specific energy ( E = 3/2 * yc ) increases (within the contracted section and u/s of the
gate), so that the u/s water depth increases and a M1 curve appears u/s of the gate. This behaviour follows well

90
CdP User’s Manual February 2017 (second edition revised and expanded)

16.1.2. Free flow discharge equation


The flow under the gate is very similar to orifice flow but not quite. First, the
flow contracts on its upper surface as it goes under the gate (and lower surface
with the presence of a gate sill) and, second, there is additional friction from
the bed of the channel. So to find a formula for discharge for this structure, the
orifice formula is a good starting point, but it needs modifying thanks to the
introduction of coefficients.
The formula for discharge from an orifice modified for a gate with free
outflow is given by the following equation:

Q = Cd * Cv * w * Bcl [(2 * g (h0 – Cc * w)]0.5 (16.1)


or
q = Cd * Cv * w [(2 * g (h0 – Cc * w)]0.5 (16.2)
Where:
Q = discharge through gate(s) (m3/s).
Refer to list of symbols.
This discharge through a gate in orifice conditions with free outflow is
governed by the u/s flow depth h0 and the gate opening w.
Due to the square root of the u/s water head h0 in this and following
equations, under-shot gates are well suited for controlling the d/s flow by
changing the gate(s) height. Contrary to the weir, variation in the u/s water
head h0 has little effect on the flow rate (refer to figure 17.4).
16.1.3. Contraction coefficient
Orifices may be partially contracted in two senses: one is the amount of
curvature of the jet in the direction of flow due to the piers (horizontal
contraction; refer to article 15.1.4) and the other in the amount of orifice
opening perimeter which produces less or no curvature of the outflowing jet
passing through the opening for greater gate openings1 (vertical contraction).
The contraction coefficient Cc varies with the relative gate opening and the
relative submergence2. The passage from orifice (regulated) to non orifice

with the intuitive notion that a severe constriction in the channel will cause the water to “back up”. The extra
specific energy, which was acquired u/s must be lost, even if there is no energy loss in the contraction itself. The
required drop of energy E can occur only through the d/s development of super-critical flow (refer to figure 14.4a
and article 25.2).
1 Because effective discharge coefficients are not well defined where suppression exists, the use of a standard

fully contracted (or fully suppressed) orifice is desirable wherever conditions permit. For a rectangular cross-
section with fully contracted submerged orifice, the discharge coefficient C d equals 0.61.
2 At large gate opening, the value of C increases, meaning that its influence on Q decreases. Very little vertical
c
contraction occurs when the gate hardly penetrates into the water stream and, usually, the contraction comes
only from the bottom weir below the gate.

91
CdP User’s Manual February 2017 (second edition revised and expanded)

(unregulated) conditions increases the flow discharge. This increase in flow is


due to the decreasing influence of the vertical contraction in the equations.
At small gate(s) openings, the
streamlines of the flow through the
gate(s) will initially not be parallel, as
shown in the opposite figure. The section
where parallel streamlines occur is called
the vena contracta. The vena contracta
effect is a result of the inability of the fluid
Figure 16.1: sluice gate with to turn a sharp 90° corner as shown in the
sharp 90° corner, full same of figure1.
contraction at orifice and vena In the User’s Manual, the
contracta effect determination of the contraction
coefficient with super-critical outflow d/s of the gate(s) is related to the depth
ratio h0/w (refer to table 16.1).
16.1.4. Discharge coefficient
In the range 1.5 < h0/w < 5, orifice flow conditions prevail with free outflow
for which the discharge coefficient Cd is applicable in the equation 16.1. The
discharge coefficient Cd and the contraction coefficient Cc can be taken from
the following table:
Table 16.1: values of discharge coefficient Cd and contraction coefficient Cc
for the depth ratio h0/w and free outflow below a gate
h0/w Cc Cd
1.5 to 2.5 0.633 0.597
2.5 to 3.5 0.625 0.599
3.5 to 5 0.624 0.605

When the ratio h0/w is less than 1.5, the table does not give any value of Cc
or Cd and the equation 16.1 can no longer be applied. At larger gate(s) openings
where the head differences between u/s water level and the tailwater level
become small (or h0/w less than 1.5), the determination of the value of Cd may
result in considerable errors and has to be derived experimentally. The flow is
considered in the User’s Manual as unidentifiable.

1The depth of flow yvena is in reality smaller than the gate opening and, therefore, partially suppressed due to the
vertical (gate) and horizontal (pier) contractions in the gate chamber. The ratio of the area of the jet, at vena
contracta, to the area of the orifice is known as coefficient of contraction C c. In other words, the coefficient shows
how the water depth contracts after the gate opening w. When the contraction is partially suppressed such as a
bottom suppressed sluice gate (no bottom weir) allowing sediments and trashes to pass the structure, the
coefficient Cc is not so well defined (only the submerged fully contracted sharp edged rectangular orifice has a
contraction coefficient well defined in laboratory tests and equals 0.61).

92
CdP User’s Manual February 2017 (second edition revised and expanded)

When the ratio h0/w is above 5, the value of the discharge coefficient Cd is
considered in the User’s Manual to be equal to 0.61; below 1.5, Cd = 0.60.
16.1.5. Velocity coefficient
For orifice conditions, it is assumed that, on the u/s side of the gate, the
depth of flow is much greater than the velocity head and so, the velocity is
neglected. Furthermore, in practice, the User’s Manual considers that the
orifice is designed and maintained so that the approach velocity to the orifice
is negligible1, thus assuring that Cv approaches unity (Cv ≈ 1)2.
If significant u/s approach velocity occurs, Cv may be approximated from the
following equation for a rectangular cross-section:

Cv = (H0 / h0)1.5 (16.3) (6.1c)


Where:
Refer to list of symbols.
16.2. CASE Nº3: GATE(S) LOWERED INTO A RECTANGULAR HORIZONTAL
CANAL WITH SUBMERGED FLOW D/S OF THE GATE(S)

16.2.1. Flow description downstream of the gate(s)


For usually large opening of the gate(s), the sub-critical flow on both sides
of the gate(s) can be established. There is some energy lost d/s the gate(s)3 and
the gate(s) is said to drown.
Independently of the gate(s) position, a d/s obstruction in the channel
forces the tailwater to a depth above the conjugate depth y2, pushing the
hydraulic jump u/s and submerged flow can be established. The gate(s) inhibits
the movement of the jump further u/s so that super-critical conditions d/s of
the gate(s) cannot be attained. Thus, the d/s conditions control (that is the
obstruction) influences the water depth u/s of the gate(s) by raising it.
16.2.2. Modular limit
To ensure free flow, the following ratio should not be exceeded by h2/w:

h2 / w < Cc / 2 [1 + 16 (h0 / (Cc * w - 1) - 1]0.5 (16.4)


Where:
Cc = theoretical minimum contraction coefficient for modular flow (C c
= 0.611 for sharp edged gate to Cc = 0.99 for rounded edge).

1 The value of the velocity coefficient varies slightly with the different forms of the gate lip.
2 For Cv = 1, the velocity of approach is zero, as would be the case if the weir were the outlet of a deep reservoir
or lake. Furthermore, in irrigation systems, Cv may be assumed to be unity since most of the canals have very flat
hydraulic gradients and the flow velocities very low.
3 A jump will emerge (but invisible) in the submerged flow.

93
CdP User’s Manual February 2017 (second edition revised and expanded)

Refer to list of symbols.


This equation is very useful since only depths are involved, which allows for
any cross-section.
The maximum gate opening for free outflow is obtained when the d/s
undisturbed sub-critical flow depth yn (considered equal to y2 on horizontal or
nearly so floor) is the conjugate flow depth of y.
The following figure shows the graphical representation of the relations
between h2/w and h0/w and the types of flow.
h2/w

h0/w

Figure 16.2: graphical representation of the modular limit and types of flows

16.2.3. Submerged flow discharge equation


The upper pool elevation is controlled by both the submergence effect of
the tailwater and the gate(s) opening: the u/s water head h0 is replaced in the
equation by the effective head (difference between the initial u/s water head
h0 and the d/s water head h2).
The applicable equation is the basic head-discharge equation for a
submerged fully contracted rectangular orifice flow given below1.

Q = Cd * Cv * w * Bcl * [(2 * g (h0 – h2)]0.5 (16.5)


Where:
Q = discharge of fully suppressed or contracted orifice (m 3/s).
Refer to list of symbols.

1 In submerged conditions, the submerged flow may also be estimated by using the modular flow with a flow
reduction factor f (refer to article 15.1.2 and equation 15.3).

94
CdP User’s Manual February 2017 (second edition revised and expanded)

The discharge rate decreases with higher tailwater depth; either the u/s
water depth must increase or the gate(s) opening must be adjusted to keep the
same discharge.
16.2.4. Contraction coefficient
The gate is considered as a submerged orifice with regulated flow and a
perimeter contraction partially suppressed with the presence of a gate sill.
The passage from free to submerged flow usually corresponds to a small
gate opening towards a large gate opening. For greater gate(s) opening, the
amount of orifice opening perimeter produces less or no curvature of the
outflowing jet passing through the opening (vertical contraction).
The contraction coefficient may be similar in submerged flow and free flow
at small openings but not at large openings. Therefore, its determination may
result in considerable errors at large gate(s) openings where the head
differences between u/s water level and the tailwater level become small.
16.2.5. Discharge coefficient
The discharge coefficient Cd depends upon h2/w (refer to article 16.1.4).
Where the head differences between u/s water level and the tailwater level
become small (or h0/w < 1.5 and close to 1), the flow is considered as
unidentifiable, because the determination of the value of C d may result in
considerable errors and should be derived experimentally.
16.2.6. Velocity coefficient
Practically, its value is also taken as 1 (refer to article 16.1.5).
the orifice opening (refer to article 16.1.5).
16.2.7. Froude number
The Froude numbers depends upon the ratio h0/w and the value of Cc. The
Froude number of the jet under submerged and under free flow conditions
equals is calculated with equation 16.5.
Submerged conditions occur for both Froude numbers Fr < 1 (fully
submerged flow d/s of the gate is sub-critical), as well as for Froude numbers Fr
> 1 (partially submerged transitional flow and super-critical d/s of the gate).
16.3. CASE Nº4: GATE(S)
LOWERED INTO A
RECTANGULAR HORIZONTAL
CANAL TO A HEIGHT CREATING
A TRANSITIONAL TYPE OF FLOW

The developed equations


Figure 16.3: flow types and conditions apply for defined flow
at gated head regulator conditions but are of typically
gate with orifice is partially under water low accuracy in the transitional
combined
95
CdP User’s Manual February 2017 (second edition revised and expanded)

zone where the flow can move from any behaviour to any other, from orifice
(regulated) to non orifice (unregulated) conditions and from free to submerged
type, as pictured in the opposite figure (refer to table 14.5).
The instantaneous unit discharge of transient flow passing through the
gate(s) is difficult to calculate.
16.4. GATE SILL

The use of an under-gate sill affects the flow behavior below and d/s of the
gate. The crest level or gate sill of the branching canal is generally kept 0.1 m to
0.5 m higher that the crest level in the river or parent canal.
In non orifice flow conditions, the flow of a gate without a sill behaves as
broad crested weir.
The energy head H and the opening w are both related to the height of the
sill. It is usually admitted that the value of Cd increases with the increase of the
Froude’s number and the sill height1. A high sill level means that the energy
head H and the opening height w are smaller, so the width of the orifice Bcl
should be increased at higher costs. The minimum costs of the structure are
often obtained with a sill at canal bed level.
Some uncontrolled variations of contraction under the gate are shown in
the following figures:

Figure 16.4: factors affecting the flow under a gate


16.5. FORCES AND MOMENT ON GATE(S)

16.5.1. Forces on closed gate(s)


Gates must be made strong enough to withstand the force created by the
hydrostatic pressure (horizontal thrust). The force acting on the whole
immersed vertical plane of the closed gate is the resultant hydrostatic force

1The increases of Cd are due to increase in the velocity through the gate opening and the gradually decreasing
value of the velocity after the gate due the d/s slope of the sill. The rate of increase depends also upon the
configuration of both the sill and the gate.

96
CdP User’s Manual February 2017 (second edition revised and expanded)

(from a distributed load to a point load) and is calculated using the following
formula1 (neglecting friction):

FR gate = A * γ water * ӯ (16.6)


Where:
FR gate = resultant hydrostatic force acting
on the gate (N).
A = area of the gate (m2).
γ water = water specific weight (9,810 N/m3).
ӯ = water depth from water surface to Figure 16.5: terminology of
centroid/centre of the closed gate (m). horizontal pressure forces
16.5.2. Forces on closed gate(s) on open gate
The point of application of the resultant force Fr gate acting perpendicular to
the immersed gate is the centre of pressure located at some depth below the
free surface. The depth to the application point of the resultant force is given
by the following formula (rectangular gate):

DRF gate = ӯ + Hgate2 / (12 * ӯ) (16.7)


Where:
DRF gate = depth to the application point of the
resultant hydrostatic force acting on the gate
(m).
Hgate = gate height (m).
ӯ = water depth from water surface to
centroid/centre of the closed gate (m).
This water depth DRF gate (depth to force D in
figure) is always greater than the water depth ӯ
from the water surface to the centroid / centre Figure 16.6: FR gate and
of the gate. In other words, the centre of depth of application
pressure is always below the centroid.

1 Pascal’s principle states that the pressure exerted by a fluid at a depth is transmitted equally in all directions.
The total pressure force on a plane area is equal to the area multiplied by the intensity of pressure at its centroid.
The computation of hydrodynamic forces acting on partially opened gates is far more complicated as it is closely
related to flow conditions.

97
CdP User’s Manual February 2017 (second edition revised and expanded)

1.60 800
Water level h0 (m)

FR gate (kg/m2)
667
1.40 700
589
1.20 600
491
1.00 441 500
392
0.80 343 400
294
0.60 245 300
196
0.40 200

0.20 100
0
0.00 0
1 2 3 4 5 6 7 8 9 10
Upstream water depth ho (m) FR gate (kg/m2)

Figure 16.7: water depth ho and FR gate acting on a vertical gate


16.6. FLOW RECAPITULATION

The following table summarizes the different flows, including flow types,
flow conditions and flow regulation.
Table 16.2: flow types, conditions and regulations with corresponding
equations
Flow types Flow conditions Regulation Equations
Orifice Regulated 16.1 or 16.2
Free
Non orifice Unregulated 15.1
Between orifice 15.1 or 15.2
Transitional Semi-regulated
& non orifice 16.1 or 16.5
Orifice Regulated 16.5
Submerged
Non orifice Unregulated 15.2 or 15.3

For conditions outside the range covered above, a comprehensive


treatment of the effects of gate(s) location and geometry on discharge for free
controlled outflow has to be carried out.
The visual difference between free and submerged flow is that the jet does
not have a free surface open to the atmosphere.

98
CdP User’s Manual February 2017 (second edition revised and expanded)

Regulation & type


Description Equation Side view
of flow
Weir or non orifice conditions
Fully open gate(s) (gate
bottom above water Q = Ce*Bcl*h03/2(eq. 15.1)
surface) with horizontal  Ce = Cd*Cv *2/3*(2/3*g)1/2 (eq. 6.1c)
Unregulated with
constriction or  Cd = 1 (with Ce =1.70) or > 1 (with
free flow under the
obstruction due to Ce =1.84) (flow state)
gate(s)
gate(s) chamber  Cv > 1
Flow non affected by  h2/h0 < 75%
d/s water conditions

Fully open gate(s) (gate Q = Ce*Bcl[2*g(h0-- h2)]0.5(eq. 15.2)


bottom above water
 Ce = Cd*Cv *2/3*(2/3*g)1/2 (eq. 6.1c)
surface) with horizontal
 Cd = 1 (with Ce 1.70) or > 1 (with Ce
Unregulated with constriction or
1.84) (flow state)
submerged flow obstruction due to
 Cv > 1
below the gate(s) gate(s) chamber
 h2/h0 > 75%
Flow affected
Q = f*Qfree flow (eq. 15.3)
(reduced) by d/s
tailwater conditions  f from table 15.1

99
CdP User’s Manual February 2017 (second edition revised and expanded)

Regulation &
Description Equation Side view
type of flow
Orifice conditions

Q = Ce*w*Bcl[(2*g (h0 - Cc*w)]0.5 (eq. 16.1)


Gate(s) partially open  Ce = Cd*Cv
Regulated with
with super-critical  0.597 < Cd < 0.605 (table 16.1)
free flow d/s of
outflow open to the  Cv ≈ 1 or (H0/h0)0.5 (eq. 16.3)
sluice gate(s)
atmosphere  h2/w < modular limit (eq. 16.4)
 Fr1 > 1 (eq. 16.5)

(with or without sill)

Q = Ce*w*Bcl[(2*g (h0-- h2)]0.5 (eq. 16.5)


Gate(s) partially open  Ce = Cd*Cv
Regulated with with flow partly  Cd = 0.61 for fully contracted
submerged submerged super- submerged orifice
flow d/s of critical or fully  Cd = 0.61 [(1 + 0.15 (Ps/P0)] (eq.
gate(s) submerged sub- 16.8) or experimental
critical flow  Cv ≈ 1
 h2/w > modular limit (eq. 16.4)
(with or without sill)
Figure 16.8: governing flow types and regulations and corresponding equations for sluice gate with orifice and non orifice
conditions

100
CdP User’s Manual February 2017 (second edition revised and expanded)

17. RATING CURVE


17.1. DISCHARGE MEASUREMENT AND CONTROL

17.1.1. Discharge measurement


Weirs, flumes and orifices can all be used for discharge measurement. But
weirs and flumes are better suited to measuring discharges in rivers when there
can be large variations in flow. Weirs and flumes not only require a simple head
reading to measure discharge but they can also pass large flows without causing
the upstream level to rise significantly and cause flooding.
Orifice structures too can be used for flow measurement but large
variations in flow also mean that the gates will need constant attention for
opening and closing (refer to article 17.7).
17.1.2. Discharge control
Orifices are rather cumbersome for discharge measurement but very useful
for discharge control because the discharge through an orifice is not very
sensitive to changes in u/s water level. If the orifice opening is assumed to be
fixed, so the discharge Q changes only when the upstream depth d changes.
In contrast, if a weir (or flume) is installed at the head of a canal, it would
be very easy to use for discharge measurement but it would not be so good for
controlling the flow because it is very sensitive to water level changes (refer to
article 17.7).
17.2. DISCHARGE RATE AND STRUCTURE DESIGN

The following table shows the type of structure to be designed according to


its location and the discharge rate to be considered.
Table 17.1: structure design according to location and discharge rate
considered
Discharge
Location Type of structure
rate
Cross-regulator width & stilling basin
Qmax Cross-regulator cut-off walls
River or
Cross-regulator protective works
parent
Number & gate(s) sizing of head regulator to
canal
Qmin reach Qcrop
Head regulator weir width to reach Qcrop
Head regulator stilling basin
Qmax Head regulator cut-off walls
Branching
Head regulator protective works
canal
Number & gate(s) sizing of head regulator
Qcrop
corresponding to Qmin in river or parent canal

101
CdP User’s Manual February 2017 (second edition revised and expanded)

The required irrigation water Qcrop through the head regulator (also called
discharge regulator) has to take into account the flow depth in the parent canal
and the effective waterway Bcl corresponding to a certain opening (bay) fitted
with sliding gates.
17.3. GATE SIZING AND NUMBER OF GATES

The hydraulic gate(s) width (Bcl) is large enough to pass the irrigation
discharge Qcrop under low flow conditions in the river or parent canal (Qmin), but
small enough to restrict the under flow to the maximum capacity of the canal,
assuming a functioning spillway.
To check the flood water entering the canal, a breast wall is provided
between designed upper pool (pond) level and high flood level. Unless the
difference between the high flood elevation and the upper pool elevation is
nominal, the construction of a breast wall is recommended (usually more
economical than higher gates).
17.4. FLOW CONDITIONS

17.4.1. Flow conditions


The flow conditions and types have to be checked (orifice or non orifice) and
the relevant equation selected for calculating the discharge rate through the
head regulator (refer to figure 16.8).
The User’s Manual recommends that, in order to avoid any significant
backwater effect in the u/s channel, the Qmax or high flood flow should be totally
unobstructed. Therefore, the equation to be selected for calculating Qmax
through the head regulator is the non orifice equation with the gate(s) opening
at least equal or greater than h0 max. With fully open gates, the Froude number
should not exceed 0.5 at the flow measuring gauge (refer to article 6.4.1).
17.4.2. Gate opening and position of the hydraulic jump
The regulated flow is computed assuming that all gates are set to the same
position (all gates are considered as one).
Free outflow remains as long as the ratio h2/w with the value of the limiting
contraction coefficient for modular flow C c = 0.611 does not exceeds the
modular limit (refer to article 16.2.2). With free flow, the adjustment of the
gate opening is based upon information from upstream (usually water head h0).
In this case, the gate opening is given by the following equation:

w = y1 / Cc (17.1)
Where:
Refer to list of symbols.

102
CdP User’s Manual February 2017 (second edition revised and expanded)

The maximum gate opening wmax with d/s free flow is obtained when the
sequent depth of y1 is exactly matching the normal depth yn in the d/s channel
obtained by the Manning’s formula (refer to table 10.1, figure 14.4a)1.
For gate openings w < wmax, super-critical flow occurs d/s of the gate(s) with
y2 > yn. The jump runs away from the gate(s) (refer to table 10.1, figure 16.8 and
article 10.5.1), forcing the water depth before the jump y1 to gradually move
downwards and increase in depth2 up to a point where the depth y1, the
sequent depth y2 and the approaching Froude number satisfy the equation 10.4
and, therefore, triggering the jump3.
For gate openings w > wmax, super-critical flow initially occur d/s of the
gate(s) with y2 < yn. The jump runs towards the gate, eventually flooding the
reach between the gate(s) and y1 (refer to table 10.1, figure 16.8 and article
10.5.1).
17.5. RATING CURVE

17.5.1. Introduction
The approach of the User’s Manual is to combine water distribution to
water level control and water measurement is utmost important for the
implementation of the Integrated Water Resources Management (IWRM) at
basin level as shown in the below schematic.
17.5.2. Rating curve for natural channel
The rating curve presents the relationships between the discharge (m3/s)
and the river stage above the arbitrary datum. In the User’s Manual, rating
curves are based on the slope-area method.
The procedure for determining a rating curve in a particular section for a
natural channel is the slope-area method based on the empirical Manning's
equation:
1. Obtain a cross-sectional profile of the channel at the point of interest.
2. Set datum at lowest point in channel.
3. Set maximum in the channel.
4. Find difference between max. and min. and subdivide into equal
increments of depth.
5. For each increment, solve for parameters of Manning’s equation.
6. Plot the resulting flows versus stage to find the rating curve.

1 With a gradually varied flow (GVF) and in certain conditions, an analog to (Fr)2 is developed for applying equation
10.4 and (Fr)2 may be substitute by (yc/y)3, y being the water depth at the point of interest (refer to article 10.3.4).
2 If depth of y is increasing while moving downwards, depth of y is decreasing.
1 2
3 In other word, the gate does not release enough water to permit the formation of a jump close to the gate.

103
CdP User’s Manual February 2017 (second edition revised and expanded)
Integrated Water Resources Management (IWRM) schematic

Weir rating curve with broad and / or short crested regime

Depth of water y n downstream of structure (m j


1.80

Depth of water ho upstream of weir crest (m)


1.60

1.40

WUA role & responasbility 1.20

1.00

1. Water request from river to SBA 0.80

0.60

2. Water allocation and distibution from main canal 0.40

0.20

3. O&M 0.00
0.00 10.00 20.00 30.00 40.00 50.00 60.00 70.00 80.00 90.00
Weir Discharge (m3/s)

RIVER GAUGE STATION


RBA/SBA role & responsibility
1. River basin planning
2. River flow forecast
3. Water allocation from river to main canal

OFF-TAKE (HEADWORK)
Off-take 1.5
Panj-Amu river basin
85

G ate ope ning he ight w and de pth of flow ho (m) 1.4 80

Flow in structure (m3 /s)


75
50.00 1.3
70
other Water User's

1.2
th ro u g h g a te ( s ) ( m 3/s )

65
45.00
1.1
Groups (WUG)
Associations &

60
1.0
40.00 55

0.9
50
Irrigation

35.00 0.8 45

0.7 40

30.00 0.6
35
D is c h a rg e Q

30
0.5
25.00 25
0.4
20
0.3
20.00 15
0.2
10

15.00 0.1 5

0.0 0
March 2 March 3 April 1 April 2 April 3 May 1 May 2 May 3 Jun. 1 Jun. 2 Jun. 3 Jul. 1 Jul. 2 Jul. 3 Aug. 1 Aug. 2 Aug. 3 Sep. 1 Sep. 2 Sep. 3
10.00 (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade)
Decadal average river or parent canal depth of flow hₒ (m) with crest = 0.40 m Decade (day)
Decadal gate opening height w (m) with max. crop water requirements (obtained from orifice and non orifice rating curves)
5.00 Decadal gate opening height w with decadal calculated crop water requirements (m)

0.00
0.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40 1.60
ho= 1.52m ho= 1.30m ho= 1.10m ho= 0.90m Gate(s) opening(m)
ho= 0.70m ho= 0.40m Qcrop =3.14 m3/s during 24hours Qcrop =9.42 m3/s during 8 hours

PANJ-AMU RIVER

Figure 17.1: Integrated Water Resources Management (IWRM) schematic

17.5.3. Rating curve of non orifice flow


The approach of the User’s Manual is to combine water level control and
flow measurement functions and to establish the best estimate rating curves of
the cross-regulator and head regulator based on the theoretical equation.

104
CdP User’s Manual February 2017 (second edition revised and expanded)

The min. u/s distance between the fixed


Weir rating curve with broad and /or short crested regime(s) staff gauge and the u/s weir toe =
1.50 3.05 m
Depth of water ho upstream of weir crest (m)

1.40 1.36
1.29
1.30 1.22
1.20 1.16
1.09
1.10 1.02 Ref. level of 0 corresponds to 100.40 m
1.00 0.95
0.88
0.90 0.81 u/s water level h0 on staff Type of flow
0.74 Q (m3/s)
0.80 gauge (m) From To
0.67
0.70 0.60 1.36 Free Free 80.04
0.60 0.53 Ref. level = 100.40 m
0.46
0.50
0.39
0.40 0.31 u/s water level h0 on staff Regime of flow
0.30 0.24 Q (m3/s)
gauge (m) From To
0.16
0.20
0.08 1.36 OK OK 80.04
0.10 0.00 Ref. level = 100.40 m
0.00 OK corresponds to broad crested weir regime
0.00 5.00 10.00 15.00 20.00 25.00 30.00 35.00 40.00 45.00 50.00 55.00 60.00 65.00 70.00 75.00 80.00 85.00 (see User's Manual. article 6.4.2)
Weir Discharge (m3/s) WARNING corresponds to short crested weir regime
(see User's Manual. article 6.4.2)
FIGURE E3: rating curve of Figure
Prout head work (CR) 17.2: rating curve
(calculated: for non
August 2015 orifice16-Sep-16
) (Printed: flow )

105
CdP User’s Manual February 2017 (second edition revised and expanded)

The weir (non orifice) flow has standard rating curves based on the general
weir equation but with particular coefficients for each type such as the broad
crested weir operating under two conditions: modular and non modular flow.
The u/s water head h0 is measured at a point where the h0 across the width of
the channel is uniform (refer to article 6.4.1).
For a weir that always operates in free flow, the calculations can offer free-
flow rating curves that condense the entire range of operations into a single
table. On the other hand, for a weir that experiences submerged flow, multiple
tables are required to provide information covering the range of u/s and d/s
water levels.
A weir rating curve is shown in the above graph taken from the attached
spreadsheet.
17.5.4. Rating curve of orifice flow
In the case of the gated head regulator (orifice flow), for a given u/s water
head h0, the opening of the gate(s) needed to produce a given discharge rate is
calculated and presented in the following graph form
For a gate that always operates in free flow, the calculations offer
continuous free-flow rating curves that condense the entire range of
operations (varying u/s water level and varying gate setting or discharge) into
a single table. On the other hand, for a gate that experiences submerged flow,
multiple tables are required to provide information covering the range of u/s
and d/s water levels, as well as a varying gate settings or discharges.
The increase of discharge rate from orifice flow to weir flow is due to the
decreasing influence of the vertical contraction coefficient C c and/or the change
of flow state (refer to 6.4.2).

106
CdP User’s Manual February 2017 (second edition revised and expanded)

10.50

10.00
Discharge Q through gate(s) (m3/s)

9.50
Qcrop = 1.89 m3/s during 24 hours
9.00
Qcrop = 5.66 m3/s during 8 hours
8.50

8.00

7.50

7.00

6.50

6.00

5.50

5.00

4.50

4.00

3.50

3.00

2.50

2.00

1.50
Ref. level = 100.00 m
1.00

0.50 The min. u/s distance between the


0.00 fixed staff gauge and the u/s gate
0.00
0.01
0.02
0.03
0.04
0.05
0.06
0.07
0.08
0.09
0.10
0.11
0.12
0.13
0.14
0.15
0.16
0.17
0.18
0.19
0.20
0.21
0.22
0.23
0.24
0.25
0.26
0.27
0.28
0.29
0.30
0.31
0.32
0.33
0.34
0.35
0.36
0.37
0.38
0.39
0.40
0.41
0.42
0.43
0.44
0.45
0.46
0.47
0.48
0.49
0.50
0.51
0.52
0.53
0.54
0.55
0.56
0.57
0.58
0.59
0.60
0.61
0.62
weir sill toe = 2.88 m

Qcrop = 1.89 m3/s during 24 hours Qcrop = 5.66 m3/s during 8 hours ho = 2.96 m ho = 2.60 m Gate(s) opening (m)
ho = 1.20 m ho = 1.00 m ho = 0.80 m ho = 0.40 m
ho = 2.40 m ho = 2.00 m ho = 1.60 m ho = 1.40 m

Figure
FIGURE D3B: 17.3a:
orifice and gate(s)
non orifice flow ratingrating
curves of curves for different
Abdullah head regulatorvalues of upstream
for a range ofwater depths
u/s water heads hoa in
(h ) and the supply
maximum
0
canal
gate opening of 0.59 m
(designed: February 2017 ) (printed: 19-Mar-17 )

107
CdP User’s Manual February 2017 (second edition revised and expanded)

17.5.5. Hydrograph and gate(s) openings


A hydrograph combines the decadal river flow (expressed in m 3/s and water
depth ho) with the corresponding decadal gate(s) opening height (w) of the
canal head regulator to reach the required irrigation flow in the main canal as
shown in the figure 17.3b.
The left ordinate represents the gate(s) opening height (w in m) and the
depth of flow (ho in m) in the river. The right ordinate represents the river flow
(m3/s) and the abscissa indicates the time in decades. The solid blue curved line
stands for the discharge in the river (m3/s) with respect to decades. The blue
vertical bars show the corresponding decadal river water depth (ho in m). The
brown columns state the decadal gate(s) opening height (w in m) calculated
with the maximum irrigation water need (water duty) while the grey bars
represent the gate opening height (w in m) calculated with the monthly
irrigation water need. Finally, the red solid line represents the minimum depth
of flow hₒ (m) in the river or parent canal corresponding to the required
discharge based on the maximum irrigation water need (July).

108
CdP User’s Manual February 2017 (second edition revised and expanded)

Decadal gate opening height w with decadal maximum crop water requirement is obtained from figure D3A or D3B above

1.5 85
Gate opening height w (m) and depth of flow ho (m)

1.4 80

Flow in structure (m3/s)


75
1.3
70
1.2
65
1.1
60
1.0
55

0.9
50

0.8 45

0.7 40

35
0.6

30
0.5
25
0.4
20
0.3
15
0.2
10

0.1 5

0.0 0
March 2 March 3 April 1 April 2 April 3 May 1 May 2 May 3 Jun. 1 (decade) Jun. 2 (decade) Jun. 3 (decade) Jul. 1 (decade) Jul. 2 (decade) Jul. 3 (decade) Aug. 1 Aug. 2 Aug. 3 Sep. 1 Sep. 2 Sep. 3
(decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade) (decade)
Decadal average river or parent canal depth of flow hₒ (m) with crest = 0.40 m Decadal gate opening height w (m) with max. crop water requirements (m3/s) Decade (day)
Decadal gate opening height w (m) with decadal calculated (real)crop water requirements (l/s/ha) Minimum depth of flow hₒ (m) in river or parent canal with crest = 0.40 m for user's defined Qcrop = 3.14 m3/s
Decadal average river or parent canal flow (m³/s) with crest = 0.40 m

Figure 17.3b: river


FIGURE D4: hydrograph
gate(s) opening andmax.&
heights w (m) with decadal corresponding gate(s)
calculated crop water requirements (l/s/ha)openings
versus river or canalof
waterthe
depthscanal head
h (m) of Prout regulator calculated
head regulator with flow during 24 with
hours the maximum 0

Based on 2014 Gerdab gauge station discharge rates (printed: 13-Mar-17 )


and decadal irrigation water need

109
CdP User’s Manual February 2017 (second edition revised and expanded)

17.6. ACCURACY OF THE RATING CURVE

17.6.1. Introduction
Stage-discharge relation is non linear and it is difficult to obtain high
accurate rating curves. One of the greatest problems with standard structures
is that they do not always conform to the chosen coefficients1.
Significant errors δ are possible because of the unique approach conditions
at proposed projects and of tailwater inaccuracies. An error can be random
(reading errors), systematic or spurious (gate malfunction or human mistakes).
17.6.2. Source of errors
For discharge measurement structures, the main sources of error to be
considered are:
1. δcoefficient: error in Ce2 (Cd*Cv) (refer to article 6.1.2).
2. δf: error in the drowned flow reduction factor f (refer to table 15.1).
3. δmn: error in dimensional measurement of the weir, such as the width
or the height of the weir.
4. δwater depth: error in gauge position and or the measurement of u/s
water depth in relation to weir sill elevation and/or Δh (refer to figure
6.1 and article 6.4).
The error δcoefficient is considered systematic3. When the flow is modular, the
drowned flow reduction factor f is constant (f = 1) and is not subject to error (δf
= 0). With submerged flow, the error consists of a systematic error 4. The error
δmn depends upon the accuracy with which the constructed structure can be
measured; it is also a systematic error. Finally, the error δwater depth is a
systematic and random error. Possible contributions to errors are an improper
maintenance of the gate(s), construction faults not included in δmn etc.
Therefore, whenever a flow rate or discharge is measured in a structure, the
value obtained should be considered as the best estimate of the true flow rate
that can be slightly greater or less than the true discharge.
17.6.3. Submerged flow type
The submerged flow type is frequently found on many installations where
the tailwater covers the jump, but is seldom recognized and corrections are
rarely made. Submerged flow in the structures will always decrease accuracy of
flow measurement. Failure to make corrections is surprising since a
considerably greater discharge may be indicated than is actually flowing.
Submerged flow type weirs can occur unintentionally by poor design,

1 For example, Cd has a standard value for standard structures.


2 Ce can vary with the presence of silt and debris accumulated in the region of dead water.
3 This classification is not entirely correct because C d and Cv are function of h1.However, the variations of the
errors in Cd and Cv as a function of h1 usually are sufficiently small to be neglected.
4 Δf is a systematic and constant error being in the numerical value of f (refer to table 15.1)and a systematic and

random error caused by the fact that f is a function of the submergence ratio H 2 / H0 or h2 / ho.

110
CdP User’s Manual February 2017 (second edition revised and expanded)

construction errors, structural settling, attempts to supply increased delivery


needs by increasing d/s heads, accumulated sediment deposits, or weed
growths.
The value of the discharge coefficient C d for submerged flow is not very
accurate and may result in considerable errors if the head differences between
u/s water level and d/s tailwater level become small (refer to article 16.2.5).
Hence, the calculation of the discharge of a vertical gate under submerged
conditions will be of lower accuracy.
17.7. CONCLUSIONS

Gate rating curves, as computed by the above equations, should be


considered as a preliminary rating curve. In finalizing the rating curves for major
head regulators, special measurements on similar existing projects should be
considered (current-meter method is more commonly used).
Although gates are rather cumbersome for discharge measurement, they
are very useful for water level control. This is because the discharge through a
free flow orifice is not very sensitive to changes in u/s water head h0, as shown
in the following graphs (refer to equations 16.1 and 15.1):

Figure 17.4: water level control using a gate (orifice flow) and weir

Suppose the parent canal water head h0 or energy head H0 (d1 and H0 in the
graphs) rises by say 20%. The effect on the discharge into the branch through
the gate is to change by only 5%. So even though there is a significant change
of h0 in the parent canal, this is hardly noticed in the branch canal. This can be
very useful for ensuring a reliable, constant flow to a farmer even though the
parent canal water head h0 may be varying considerably.

111
CdP User’s Manual February 2017 (second edition revised and expanded)

112
PART VI:

SCOUR DEPTH IN CHANNEL BED

113
CdP User’s Manual February 2017 (second edition revised and expanded)

18.WATERWAY AND REGIME SCOUR DEPTH


18.1. WATERWAY DETERMINATION

18.1.1. Wetted perimeter


For shallow and meandering rivers, the minimum stable width can be
calculated from Lacey’s modified formula:

Pw = a * (Qmax)0.5 (18.1)
Where:
Pw = Lacey’s wetted perimeter or minimum waterway width1 (or stable
width of waterway) (m).
Qmax = maximum flow discharge in the river (HFL) (m3/s).
a = coefficient depending on stability of river channel.
B = overall waterway width between river banks or abutments (m).
For large river, the wetted perimeter Pw is practically equal to the river
width. If the width of the river is considerably larger in comparison to the water
depth, the computed perimeter is provided as width only.
With these tentative values, the adequacy of the waterway and the crest
level for passing the design flood within the permissible afflux (refer to figure
5.2) need be checked and readjusted in such a way that the permissible values
of afflux are not exceeded.
18.1.2. Looseness factor
The proposed above Lacey’s equation applies when the looseness factor is 
1. The ratio of actual width to the regime width of the river is the looseness factor,
as shown in the following equation:

Looseness factor = B / Pw (18.2)


Where:
Pw = Lacey’s wetted perimeter or minimum waterway width2 (or stable
width of waterway) (m).
Qmax = maximum discharge in the river (HFL) (m3/s).
a = coefficient depending on stability of river channel.
B = overall waterway width between river banks or abutments (m).

1 For large river, the wetted perimeter Pw is practically equal to the river width. If the width of the river is
considerably larger in comparison to the depth of water, the computed perimeter is provided as width only.
2 For large river, the wetted perimeter P is practically equal to the river width. If the width of the river is
w
considerably larger in comparison to the depth of water, the computed perimeter is provided as width only.

114
CdP User’s Manual February 2017 (second edition revised and expanded)

A looseness factor  1 is obtained when Lacey’s waterway is restricted due


to fluming or contraction of normal/regime waterway i.e. when the actual
waterway provided for the structure is less than Lacey’s waterway.
Table 18.1: coefficient a in Lacey’s minimum stable width of waterway
Channel type a Remark
Stable channel in scour resistant material 3.3 In most literature
Shifting channel in sandy material 4.9 4.75 is adopted

Generally, with a high looseness factor, there is a tendency for shoal formation
d/s of the structure, which may cause maintenance problems.
18.2. SCOUR DEPTH

Depending on the bed material, the river bed, during peak flood flows, may
HFL
become mobile to several
meters below the normally
observed river bed. Unless
bed level during low flow R founded on rock, the structure
needs to be designed to
bed leve during high flow withstand such deep scour
(= scour line)
depths (refer to opposite
Figure 18.1: mean scour depth R figure).
measured from the high flood level The Lacey’s method of
estimating the river scour depth
R is used in this Manual. This method is basically empirical and essentially gives
total scour below high flood level (HFL) in the case of meandering rivers in flood
plain and is meant for non cohesive sandy material with mean sediment size of
about 0.15 mm to 0.43 mm. The method is not valid outside this range1.
18.3. SOIL CONDUCTIVITY

The grain size distribution of a soil is one of the geotechnical aspects of soil
mechanic properties that affect the hydrogeological conductivity. A sorted soil
with larger grains has a high hydraulic conductivity. If a sediment contains a
mixture of grain sizes (multi-graded soil), the porosity will be lowered, and thus
the hydraulic conductivity. This is because the void between the larger grains is
filled up with smaller grains.

1 Lacey’s theory is applicable only to stable alluvial rivers and no to rocky or boulder river stages and unstable
(aggrading or degrading) alluvial rivers. In the case of coarser material with larger standard deviation, as scour
progresses, scouring occurs by selective removal of finer material from scour hole and hence smaller scour depth
will occur. For very fine material, having cohesion, it is generally considered that there will be greater resistance
to scour and hence reduced scour depth will result. Further, due to effects being site specific, larger variations in
scour depth are likely to occur which cannot be related to Q and f alone. However, the method is commonly used
in design of structures.

115
CdP User’s Manual February 2017 (second edition revised and expanded)

Hydraulic conductivity (or transmissivity) is correlated to the particle size


and can be estimated by using methods based on grain size analysis. This grain-
size analysis estimates representative value of the grain size (d501, d40 etc.).
The following figure gives a simple classification of the major soil groups
according to the particle size (mm).

Figure 18.2: classification of major soil groups according to particle size (mm)

The following table gives the average particle size d50 for different types of
material.
Table 18.2: indicative values of average sediment grade scale (size particle
d50 in mm)
Average grain size d50
Type of material (soil)
(mm)
Very fine 0.0005 to 0.00024
Fine 0.001 to 0.0005
Clay
Medium 0.002 to 0.001
Coarse 0.004 to 0.002
Very fine 0.008 to 0.04
Fine 0.0016 to 0.008
Silt
Medium 0,031 to 0.016
Coarse 0.062 to 0.031
Very fine 0.125 to 0.062
Fine 0.250 to 0.125
Sand Medium 0.500 to 0.250
Coarse 1.000 to 0.500
Very coarse 2.000 to 1.000
Very fine 4 to 2
Fine 8 to 4
Gravel Medium 16 to 8
Coarse 32 to 16
Very coarse 64 to 32

1
d50: average particle size of the soil which 50% of the material is finer.

116
CdP User’s Manual February 2017 (second edition revised and expanded)

Small cobble 128 to 64


Large cobble 256 to 128
Small boulder 512 to 256
Boulder/cobble
Medium boulder 1,024 to 512
Large boulder 2,048 to 1,024
Very large boulder 4,096 to 2,048
18.4. SILT FACTOR

With known average particle size d50 (mm) of the bed material (where the
structure will be imbedded), the silt factor f may be calculated from the
following relationship:
f = 1.76 (d50)1/2 (18.3)
Where:
d50 = average particle size or diameter of the channel bed material
(mm)1 (refer to table below).
The following table gives some indicative values of the Lacey’s silt factor:
Table 18.3: indicative values of the Lacey’s silt factor f
Soil type Lacey’s silt factor f
Boulders and shingle 20.0 to 15.0
Boulders and gravel 12.5
Medium boulders, shingle and sand 10.0
Gravel 4.75
Coarse sand 1.5
Medium sand 1.25
Standard silt 1.0
Medium silt 0.85
Fine silt 0.6
Clay 0.05
18.5. CALCULATION OF REGIME SCOUR DEPTH IN CHANNEL

Scour depth depends largely on the following factors:


 Particle size of the channel bed material.
 Intensity of flow per unit width.
 Velocity of flow.
During major floods, the main incised channel tends to increase in size
towards the geometry corresponding to the peak flow rate. It is unlikely that a
full adjustment to this geometry will be achieved during an individual flood
because the appropriate changes in channel width and longitudinal hydraulic
gradient take a considerable time. Nevertheless, because of the uncertainty of

1 The sieve size through which 50% of the material passes by weight in mm)

117
CdP User’s Manual February 2017 (second edition revised and expanded)

how far any short-term changes will progress, it is assumed in the User’s
Manual that the geometry corresponding to the design flood would be reached.
18.5.1. Mean scour depth
When the river width does not equal the wetted perimeter1 of 4.75 √Q, the
mean scour depth for natural channels flowing in non cohesive soils2 (where
the structure will be built) measured from the high flood level (Qmax) is
calculated from the Lacey’s mean scour depth equation3 as follows:

R = 1.35 (q2 / f)1/3 (18.4)


Where:
R = mean scour depth measured from the high flood level (m).
q = discharge per unit width of stream allowing for concentration of
flow where the width is the actual river width taken as the flood water
width at the given site (m3/s.m).
f = Lacey’s silt factor related to grain size d50.
The above equation applies when the looseness factor < 1.
18.5.2. Normal scour depth
When the river width > or = the regime width of 4.75 √Q, the normal scour
depth for natural channels flowing in non cohesive soils (where the structure
will be built) with Qmax (HFL) is calculated from the Lacey’s normal scour depth
equation4 as follows:

R = 0.473 (Qmax / f)1/3 (18.5)


Where:
Qmax = design flood discharge at the given site (m3/s).
Refer to list of symbols.
The above equation applies when the looseness factor > 1.
18.5.3. Maximum discharge intensity per unit width of weir
The unit discharge intensity q during high flows may be kept low to avoid
costly energy dissipater. The following table provides values of discharge
intensities as a criterion to design the minimum required width of weirs.
Table 18.4: foundation material and corresponding maximum unit discharge
per unit width of stream

1 It means, the river width is still active.


2
Alluvium.
3 The equation is based on alluvial regime and may not be quite correct for large river or and for boulder or clayey

reaches.
4 The equation is based on alluvial regime and may not be quite correct for large river or and for boulder or clayey

reaches.

118
CdP User’s Manual February 2017 (second edition revised and expanded)

Foundation Unit discharge q per m


material width
(m3/s.m)
Fine sand 5.0
Coarse sand 10.0
Sand & gravel 15.0
Sandy clay 20.0
Clay 25.0
Rock 50.0

In addition, the discharge intensity is related to the depth of flow and the
velocity of flow which should be kept as much as possible below the value of
the erosion threshold.

119
CdP User’s Manual February 2017 (second edition revised and expanded)

19.CUT-OFF WALL
19.1. CUT-OFF WALL ROLE

The purpose of providing cut-off walls is two-folded: (i) increases the flow
path and (ii) reduces the uplift pressure, ensuring stability to the structure as
shown in the following figure:

Figure 19.1: effect of cut-off walls on seepage lines

The u/s cut-off wall is more efficient in reducing the uplift pressure while
the d/s cut-off pile is more effective in reducing piping.
While designing a cross-regulator, d/s cut-off from the maximum scoured
depth considerations is, first of all, provided and then checked for Ge. If a safe
value of Ge is not obtained, then the depth of cut-off or the length of the
impervious floor is increased (refer to article 22.2).
19.2. CUT-OFF WALL DEPTH

The depth of the u/s pile line will be governed by the scour depth R alone,
while, on the d/s end, both the scour depth R and exit gradient Ge have to be
considered.
The intermediate sheet pile lines are not required from consideration of
scour or exit gradient but they act as important secondary lines of defense.
They are also helpful in the matter of distribution of pressure due to uplift
pressure.
The scour depth D measured from the channel bed level is given by the
following equation:
D = FoS * R – yn (19.1)

120
CdP User’s Manual February 2017 (second edition revised and expanded)

Where:
D = depth of scour measured from channel bed level (m).
yn = depth of flow from rating curve (m).
Refer to list of symbols.
To the calculated natural scour depth R of the river from the water surface
level, a safety factor (FoS) is applied as given by the following table:
Table 19.1: values of safety factor (FoS) for scour depth
Location) FoS
U/s cut-off depth 1.25
D/s cut-off depth 1.50 or 1.75
Pier 2.00
Flow concentration 2.00 to 3.00

In the d/s expansion of section, the velocity flow distribution is uneven in


the cross-section, which could lead to asymmetry of flow and the development
of erosion in places of highly concentrated velocity. Therefore, a higher value
of the factor of safety (FoS) is always used in the design of the d/s protected
transition.
The cut-off wall depth d1 (u/s cut-off) or d2 (d/s cut-off) is equal to the scour
depth D measured from the channel bed level multiplied by a safety factor.
It has to be noted that the depth of the d/s cut-off wall will be governed not
only by the local scour depth but also by the exit gradient value (refer to article
22.5.3). This cut-off depth is also limited by practical considerations, as very
deep walls can be difficult or impracticable to build at the site.
The cut-off walls must extend from one bank of the cross-regulator to the
other bank, and underneath the wing walls and return walls.

121
CdP User’s Manual February 2017 (second edition revised and expanded)

122
CdP User’s Manual February 2017 (second edition revised and expanded)

PART VII:

STABILITY OF THE STRUCURE ON PERMEABLE


FONDATIONS

123
CdP User’s Manual February 2017 (second edition revised and expanded)

20.FLOW PATH IN PERMEABLE SOILS


20.1. SUB-SOIL DATA

The design should match the permeability and the bearing capacity of the
sub-soil. Permeability is the measure
of the soil ability to permit water to
flow through its pores or voids.
Explorations in the river bed shall
be confined to periods of low river
flows. Test pits through manual
labour/back hoe will be excavated to
expose the top stratum for physical
Figure 20.1: soil permeability examination, in-situ testing and
sampling.
20.2. STREAM LINES

The stream lines represent the paths along which the water flows through
the sub-soil. Every particle entering the sub-soil at a specific point upstream of
the structure will trace out its own path and will represent a stream line. The
stream line flow is a flow in which each liquid particle has a definite path and
the paths of adjacent particles do not cross each other.
Every stream line possesses a difference of head. Further, at every
intermediate point along the stream line, there is a residual head still to be
dissipated in the remaining length to be traveled to the d/s end1.
20.3. SUB-SOIL PRESSURE

The seepage pressure is due to the head difference between two points in
a given mass of soil and it acts on the soil particles. The unit pressure is
measured in kN/m2 or in m (head).
The seepage flow, moving through the pores of the sub-soil, causes two
types of sub-soil pressure affecting the stability of the structure:
1. Direct underneath uplift pressure throughout the sub-soil of the
structure river or parent canal bed that tends to lift up the hydraulic
structure floor.
2. Upward raising pressure d/s of the solid apron which causes sand
particles to erupt upwards, creating the piping phenomenon
(retrograde erosion).

1
Khosla’s theory shows that the loss of head along the flow net does not take place uniformly but
depends upon the whole geometry of the structure, including the shape of the foundations, the u/s
and d/s bed elevation etc.
124
CdP User’s Manual February 2017 (second edition revised and expanded)

21.PRINCIPAL CAUSES OF INESTABILITY


21.1. PRINCIPAL FAILURE OF A STRUCTURE

The design of a safe structure has to meet the surface flow requirements
and guard against uplift pressure and seepage due to the residual force (head
potential) of sub-surface water flowing from u/s end to the d/s end of a
structure1.
The main causes of failure of a structure constructed on a permeable
foundation can be classified broadly into the two following categories.
21.1.1. Failure due to sub-surface flow
Failure by piping or undermining: the water from the u/s side continuously
percolates through the bottom of the foundation and emerges at the d/s end
of the cross-regulator
floor. When the seepage
water retains sufficient
residual pressure at the
emerging d/s end of the
structure, it may lift up
and remove the soil
particles by scouring at
Figure 21.1: seepage problem under levees the point of emergence,
(U: d/s uplift pressure force; W: submerged leading to increased
weight of soil) porosity of the soil and
formation of small
cavities. A depression occurs under the structure which extends backwards
towards the u/s through the bottom of the foundation (retrograde erosion).
The structure may ultimately subsides in the hollow so formed, resulting in the
failure of the structure 2.
Failure by direct uplift: the percolating water exerts an upward pressure along
the foundation of the cross-regulator. If this uplift pressure is not
counterbalanced by the self weight of the structure, it may fail by rupture.
21.1.2. Failure due to surface flow
Unbalanced head due to standing wave: with super-critical flows, a hydraulic
jump may develop. This jump causes a suction pressure or negative pressure
which acts in the direction of the uplift pressure. If the thickness of the
impervious floor is insufficient, then the structure fails by rupture (refer to
articled 23.4.4).

1 Calculated with the Khosla’s method


2 This failure is initiated by the sand boiling phenomenon, as shown in the above figure.

125
CdP User’s Manual February 2017 (second edition revised and expanded)

By scouring: a high discharge rate results in scouring effects on the d/s and u/s
side of the structure. Due to scouring of the soil on both sides of the structure,
its stability gets endangered by shearing.
21.2. ADOPTED SOLUTIONS AGAINST FAILURE

Solutions to the problems of instability due to the sub-soil pressures d/s of


the structure may be obtained by lengthening the journey of water from
beneath the structure (longer impervious floor, deeper cut-offs)1. In addition,
the provision of filters (granular filter or fabric filter) underneath the u/s and
d/s protective works of the structure is to check flowing out of fine subgrade
material (refer to figure 24.1 and 24.3).
The scouring of the d/s channel bed is prevented by the construction of an
adequate stilling basin and cut-off walls.
A failure by direct uplift due to the upward pressure of the percolating water
along the foundation of the cross-regulator can be checked by increasing the
self weight of the structure.
Those works can be expensive. For this reason, the trend is trying to reduce
its dimensions in thickness as its length. Particularly in the case of structures of
which large numbers need be constructed, the designer will be constrained by
economic pressure to keep the structure as short as possible. In this case,
hydraulic requirements will probably have the determining effect on the
horizontal length. The length of the vertical cut-offs, especially the d/s one, will
then be determined by the requirements for seepage resistance.
In practical terms, the designer must decide on an appropriate balance
between the length of the horizontal elements (H) and the vertical ones (V) in
order to arrive at an economic design for a particular type of soil; the V/H ratio
of length 1/1.5 to 1/1.0 is widely adopted. Practices of design in a given country
or region can effectively dictate this balance.

1More energy head will be dissipated by friction and other losses through the path of flow line ensuring the
decrease of the exit gradient.

126
CdP User’s Manual February 2017 (second edition revised and expanded)

22. EXIT GRADIENT AND STRUCTURE LENGTH


22.1. HYDRAULIC GRADIENT

The hydraulic gradient is the rate of loss in unit pressure in kN/m2 (or head
in m) due to friction per unit of distance of channel at a given point and in a
given direction of the flow path.
At the exit end of the structure, this gradient of pressure of water is called
the exit gradient Ge.
The minimum length of the floor of the hydraulic structure is determined
primarily from exit gradient considerations.
22.2. EQUATION OF THE EXIT GRADIENT

As per Khosla’s method, for a standard form of a hydraulic structure as used


in this User’s Manual consisting of a floor length b with a vertical cut-off of
depth d, the exit gradient at its d/s end is given by:

Hs 1
Ge  *
d π *  (22.1)
Where:
Ge = calculated exit gradient.
Hs max = maximum hydrostatic head where Ge is acting (without flow)
(m).
d = d/s vertical cut-off wall depth (m).
1/π*√λ = calculated or estimated factor depending on the length of the
floor of the structure and the depth of the d/s cut-off wall.
For the calculation of Ge, the maximum hydrostatic head Hs max is the
difference between the weir crest elevation (or the water surface elevation
with closed gate and with or without breast wall) and the d/s practically dry
channel bed elevation, where the exit gradient pressure is acting.
The above equation or its equivalent graphical form Khosla’s pressure
curves (not given in the Manual) gives a value of Ge equal to infinity if there is
no d/s cut-off (d = 0 in the equation). It is therefore essential that a d/s sheet
cut-off invariably be provided for any structure considered in the User’s
Manual.
22.3. SAFE EXIT GRADIENT

The exit gradient Ge is:


 Critical when the uplift (disturbing) pressure on the grain
comprising the river bed material is just equal to the submerged
weight of the grain at the exit point.
127
CdP User’s Manual February 2017 (second edition revised and expanded)

 Safe when a factor of safety (FoS) is used.


Values of the safe exit gradient Gse of the soil comprising the river bed
material are given in the table below.
Table 22.1: values of Khosla’s safe exit gradient Gse for different types of
soil
Type of soil Safe exit gradient Gse
Fine sand 1/6 a 1/7 0.17 to 0.14
Coarse sand 1/5 a 1/6 0.20 to 0.17
Shingle 1/4 a 1/5 0.25 to 0.20
Clay 1/3 a 1/4 0.33 to 0.25

For example, a safe exit gradient Gse equal to 1/5 of the critical exit gradient
Ge is necessary to maintain the structure safe on coarse sand.
22.4. DETERMINATION OF THE EXIT GRADIENT

To keep a structure safe against erosion (piping or other types), the value of
the calculated exit gradient from equation 22.1 must be less than the value of
the safe (permissible) exit gradient of the soil comprising the river bed material
encountered from table 22.1.
22.4.1. Estimated value of λ
. Knowing the value of the maximum hydrostatic head Hs max and the d/s cut-
off depth, the value of 1/π*λ1/2 (and, therefore, Ge) can be obtained from the
design chart shown in the following figure.
The α factor relates the floor length b of the structure with respect to the
depth d of the d/s cut-off. For any value of α, the corresponding value of the
factor 1/ (π*√λ) factor may be deducted from curve Nº1 in the design chart.
22.4.2. Calculated value of λ
. Knowing the value of the maximum hydrostatic head Hs max and the d/s cut-
off depth, the value of Ge can be obtained from analytical form with the
equation λ = [1 + (1 + α2)1/2]/2, with known values for the structure solid floor
length b.

128
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 22.1: design chart of λ in relation to α for estimating the exit gradient
Ge
22.5. EXIT GRADIENT CONTROL

22.5.1. Parameters of control


The extent of scour of the bed material varies at different places along a
hydraulic structure. The control of the exit gradient depends upon two
parameters:
 The length of the solid floor.
 The depth of end cut-off wall.
22.5.2. Length of the solid floor
Naturally longer the seepage lines of flow through the soil, more is the head
loss and thus lesser the uplift pressure and the value of exit gradient.
In case of Ge more pronounced, the length b of the impervious floor is
extended or the depth d of the cut-off made deeper in order to reduce the value
of Ge until a secure value be reached. Therefore, this approach allows the user
to analyze alternatives to reach a more economic design (refer to article 21.2).
22.5.3. Depth of cut-off wall
The depth of end cut-off wall is also helpful in reducing the exit gradient
value and thus the risk of piping (refer to article 19.2).

129
CdP User’s Manual February 2017 (second edition revised and expanded)

23.UPLIFT SUB-SOIL PRESSURE AND FLOOR THICKNESS


23.1. FLOOR THICKNESS

The structure floor should be thick enough to resist the uplift pressure and
withstand wear by moving bed load and suspended material and the impact of
stones and flowing water.
The determination of the hydrostatic pressure and the weight of the floor
itself counteract the destabilizing uplift pressure directly under the floor. The
thickness of the floor is calculated from the balance of both pressures, taking
into account the corrections to be applied as shown in the following equation:

tfloor = (Hmax * φc)/ u (23.1)


Where:
tfloor = thickness of floor at specified cross-section (m).
φc = corrected uplift pressure at specified cross-section (in % of the
total uplift).
Hmax = maximum static or dynamic head at specified cross-section (m).
u = submerged specific gravity or density of floor material.
The maximum difference of head and hence the maximum uplift pressure
imposed on the structure is when the water reaches its highest level with closed
gate(s) without any discharge passing down the structure. When a discharge is
passing through the structure, the maximum (seepage) head is the difference
between the u/s and d/s water level.
Floors can be designed with stone masonry covered by a 0.25 m protection
layer of reinforced concrete or with thinner PCC/RCC.
23.2. STANDARD PROFILES

In the User’s Manual, for the calculation of sub-soil uplift unit pressure
(kN/m2) for designing hydraulic structures on pervious foundations, the
Khosla’s method of independent variables is used, breaking a complex profile
like that of a cross-regulator structure into a number of simple standard
profiles, each of which can be solved mathematically, as shown in the following
figure:

130
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 23.1: standard profiles with (a) u/s sheet pile and (b) d/s sheet pile and
key points E, C and D
23.3. UPLIFT PRESSURE

23.3.1. Key points


The key points of the standard profiles, where the uplift pressure is
computed, are:
 U/s and d/s junction points of the sheet pile with the floor.
 Bottom point of each sheet pile.
 Bottom corners in case of depressed floor and/or at the foot of the
glacis.
The maximum hydrostatic head Hs max considered in the case of floor
thickness calculation is the head difference between the u/s water elevation up
to the highest level (weir crest elevation or pond water surface elevation1 with
closed gate) and the d/s practically dry cistern bed elevation. In other words,
seepage pressure is the most dominating for gates closed conditions (Hs max),
average during normal flow and almost none during high flow as shown in the
following figure:

Figure 23.2: profiles of the sub-soil seepage grade line

1
With cross-regulator without shutters, crest elevation equals pond elevation.

131
CdP User’s Manual February 2017 (second edition revised and expanded)

Mathematical solutions1 for the determination of the residual hydrostatic


head percentage at various key points of the flow net in a vertical plan for the
adopted simple standard profiles are presented in the form of equations given
below.
23.3.2. Upstream pile
For the u/s pile, the main equations are:
φE1 = 100 % of height of the maximum hydrostatic head Hs max.
φC1 = 100% of height of the maximum hydrostatic head Hs max – φE.
φD1 = 100 % of height of the maximum hydrostatic head Hs max – φD.
23.3.3. Downstream pile
For the d/s pile, the main equations are:
φE2 = (1/π) COS-1((λ-2)/λ
φC2 = 0% of height of the maximum hydrostatic head Hs max
φD2 = (1/π) COS-1((λ-1)/λ
23.4. CORRECTIONS OF UPLIFT PRESSURE

However, the reality does not comply with the hypothesis of a standard
simplified independent profile for the individual key points of the hydraulic
structure.
The real profile is a complex profile specific to the whole structure.
Therefore, the application of simple standard profiles is valid for any complex
profile only if the percentage pressure at the key points mentioned above is
corrected for:
1. Correction for mutual interference of piles or cut-off walls.
2. Correction for floor thickness.
3. Correction for slope of floor only if a cut-off is positioned at the
start or end of the slope 2.
23.4.1. Correction for mutual interference of cut-off walls
The correction to be applied at C1 or E2 as percentage of head Hs max due to
the effect of the presence of the other cut-off wall is given by:

𝐃 𝐝+𝐃
𝐂𝐢𝐧𝐭𝐞𝐫𝐟𝐞𝐫𝐞𝐧𝐜𝐞 = 𝟏𝟗 √𝐛′ ( ) (23.2)
𝐛
Where:
Cinterference = correction to be applied at C1 or E2 due to mutual
interference of piles (%).
b = total floor length (m).

1 In the Manual, to enhance the precision, we only use the analytical approach and not the Khosla’s abacus of
pressure curves.
2 In this User’s Manual, this correction is neglected owing that, usually, there is no cut-off at the toe of glacis.

132
CdP User’s Manual February 2017 (second edition revised and expanded)

d = depth of the cut-off on which the effect is considered (m).


D = depth of the cut-off, the influence of which has to be determined
on the neighboring cut-off (measured below the level at which
interference is desired) (m).
b’ = c/c distance between the 2 cut-off walls considered (d & D) (m).
The correction due to the interference is positive in the rear of the pile
(points C) and subtractive for the points in the forward direction of the flow
(points E).
23.4.2. Correction for the floor thickness
The thickness of the (concrete) floor is designed according to the (sub) up-
thrust pressure, using the weight of gravity in each reach. Therefore, the
pressure percentages φE1 and φC1 calculated by the Khosla’s equation1 pertain
to the top level of the floor.
In the standard form profiles (refer to figure 23.1), the u/s floor is assumed
to have negligible thickness, since the uplift pressure is more than
counterbalanced by the specific weight of the water γ above the floor. The u/s
floor however needs a minimum thickness in order to withstand wear by
moving bed load and suspended material and the impact of flowing water.
The corrected pressure at C1 with an u/s cut-off wall is calculated with the
following equation (the correction is positive at C1):

Correctionthick = [(φD1 – φC1) * tC1] / d1 (23.3)


Where:
Correctionthick = correction to be applied at C1 due to u/s floor
thickness as % of head Hs max (%).
φD1 = % of Hs max at D1 (%).
φC1 = % of Hs max at C1 (%).
tC1 = assumed u/s floor nominal thickness (m).
d1 = u/s cut-off depth including floor nominal thickness (m).
The corrected pressure at E1 with a d/s cut-off wall is calculated with the
following equation (the correction is negative at E2):

Correctionthickness = [(φE2 – φD2) * tC2] / d2 (23.4)


Where:
Correctionthickness = correction to be applied at E2 due to d/s floor
thickness as % of head Hs max (%).
φE2 = % of Hs max at E2 (%).

1The percentage pressures φE1 y φC1 can also be estimated thanks to available Khosla’s abacus (not given in the
User’s Manual).

133
CdP User’s Manual February 2017 (second edition revised and expanded)

φD2 = % of Hs max at D2 (%).


tC2 = assumed d/s floor thickness (m).
d2 = d/s cut-off depth including floor thickness (m).
23.4.3. Correction for floor slope
A correction is applied for the slope of the floor, only to the key points of
the pile line fixed at the start or the end of the slope. This correction is
neglected owing that, usually, there is no cut-off at the toe of glacis.
In most of the structures considered in the User’s Manual, the presence of
a cut-off wall at the toe of the glacis is not necessary. Therefore, there is no
correction for mutual interference and slope at this point.
23.4.4. Correction in the jump trough
In the User’s Manual, the percentage of pressure in the jump trough is
computed at the toe of the glacis considered as the theoretical base point of
the jump formation and is obtained with the following equation:

φt = φE2c + [(φC1c – φ E2c) / b] * bbasin (23.5)


Where:
φt = % of Hs max or Hd max at toe of glacis (%).
b = total floor length (m).
bbasin = basin floor length (m).
φE2c = % of Hs max at E2 corrected for mutual interference and floor
thickness (%).
φC1c = % of Hs max at C1 corrected for mutual interference and floor
thickness (%).
In the jump trough, for the determination of the maximum head, we must
examine two alternatives: the presence or absence of water flow. The thickness
of the floor is calculated from the balance of both pressures: hydrostatic uplift
pressure and dynamic pressure.
No water flow conditions (static conditions)
The maximum hydrostatic head Hs max and, hence, the maximum uplift
pressure imposed on the structure is the difference between the u/s water
surface elevation due to the presence of the weir (or with the gate(s) closed in
case of an off-take structure or discharge regulator) and the dry cistern
elevation1 (refer to figure 23.2).
The unbalanced hydrostatic head at the toe of glacis Hs glacis due to the
maximum uplift pressure Hs max and therefore the level of the sub-soil
hydrostatic grade line at the same point is given by the following equation:

1 The water surface elevation may be taken because v2/2g = 0.

134
CdP User’s Manual February 2017 (second edition revised and expanded)

Hs glacis = Φt * Hs max (23.6)


Where:
Hs glacis = unbalanced hydrostatic head at toe of glacis (m).
Hs max = maximum hydrostatic head (m).
φt = uplift pressure at the jump location (in % of the total uplift pressure
in key point E1).
The unbalanced hydrostatic head at toe of glacis Hs glacis is also the difference
between the sub-soil hydrostatic pressure elevation at this point and the cistern
elevation at the same point.
Water flow conditions (dynamic conditions)
When a certain discharge rate is passing over the weir with the formation
of a hydraulic jump, the maximum dynamic head Hd max, neglecting the velocity
head, is the difference between the u/s water surface elevation and the d/s
water surface elevation1 (refer to figure 23.2).
To avoid plotting the jump water surface elevation for knowing the dynamic
head Hd at the formation of the jump2, the unbalanced dynamic uplift pressure
(head) at the toe of glacis Hd glacis may approximately be calculated, using the
following equation:

Hd glacis = 0.5 (y2 – y1) + φt * Hd max (23.7)


Where:
Hd glacis = unbalanced dynamic head at toe of glacis (m).
φt = uplift pressure at the jump location (in % of the total uplift pressure
in key point E1).
Hd max = maximum dynamic head (m).
Refer to list of symbols.
The dynamic uplift pressure (in m head) Hd glacis due to the dynamic action
of the water flow where the jump is forming is compared to the hydrostatic
head Hs glacis at the same point (toe of glacis). The thickness of the glacis floor is
designed for the dynamic head Hd glacis or the hydrostatic head Hs glacis, whichever
is greater and divided by the submerged density of the floor material (concrete
or masonry).
The inflow conditions such as the velocity may modify substantially the flow
properties and affect the jump characteristics and its classification. The bottom
pressure fluctuation below the hydraulic jump may create negative pressures
that could lead to uplift pressure on the cistern floor. Therefore, since the base

1 We take the tailwater level in order to avoid plotting the water surface profile after the jump. Theoretically, it
is necessary to plot the post jump profile to determine Hs max.
2 Here considered at the toe of glacis (refer to article 10.5.3).

135
CdP User’s Manual February 2017 (second edition revised and expanded)

point of the jump formation is likely to shift with the variation in discharge rate
passing over the weir, the entire glacis has to be designed with the same floor
thickness from u/s to d/s.
23.5. HEADS AND GRADIENTS

In the User’s Manual spreadsheet, the green curve represents the elevation
of the water surface profile and the black curve represents the elevation of the
structure floor.
The sub-soil hydraulic grade line is drawn in the spreadsheet for the
maximum static head for high flood conditions (dotted red curve) and the no
flow conditions (solid red curve).
The (unbalanced) dynamic head is represented by the difference of
elevation between the static head for high flood conditions and the water
surface profile at the same vertical section (green curve). The (unbalanced)
static head is represented by the difference of elevation between the sub-soil
hydraulic grade line and the structure floor elevation (black curve) at the same
section. This unbalanced head is compensated by the floor thickness, taking
into account the buoyancy.

136
CdP User’s Manual February 2017 (second edition revised and expanded)

PART VIII:

UPSTREAM AND DOWNSTREAM PROTECTIVE


AND TRANSITION WORKS

137
CdP User’s Manual February 2017 (second edition revised and expanded)

24.UPSTREAM AND DOWNSTREAM PROTECTIVE WORKS


24.1. LOCATION

Upstream and downstream of the solid floor of the hydraulic structure, the
channel bed is protected by certain methods like concrete blocks protection,
loose stones protection, gabion boxes protection etc.
U/s of the structure, the flow velocity is lower than the approach (or
accelerated) velocity in the structure where the water depth decreases (refer
to table 25.1). D/s of the structure, in the stretch of expansion, the speed is
reduced and the water surface elevation rises leading to asymmetry of flow and
concentrated velocity (refer to table 25.1).
24.2. TYPE OF PROTECTION WORKS

Protection should, ideally, be flexible so that it can respond to changes in


the channel reach. The u/s and d/s protective works are of two types:
 Filter (granular, inverted or graded filter or fabric filter).
 Launching apron.
Many variations can be made on the basic filter construction. One or more
layer can be replaced or only the revetment layer is maintained, while the
underlyning layers are replaced by one single layer such as:
 Gabion boxes (matresses) on fine gravel.
 Concrete blocks on nylon filter (geotextile).
No concentration of flow is taken while designing the protective works.
24.3. INVERTED GRANULAR FILTER

24.3.1. Description
If the protective lining were to be installed on top of the fine material in
which the canal is excavated, grain of this subgrade would be washed through
the openings of the revetment1 (riprap or concrete blocks). To avoid damage to
the protective lining due to the washing of the subgrade, a filter must be placed
between the revetment (riprap or concrete blocks) and the subgrade.
The inverted graded filter consists of 2 layers of graded materials of
increasing permeability from bottom to top, as shown in the following figure.
The protective construction as a whole and each separate layer of the filter
must be sufficiently permeable to water entering the canal through its bed or
bank. Further, fine material from an underlying filter layer or from the original
material (subgrade) must not be washed into the void of the covering layer.

1This process is partly due to the turbulent flow of canal water in and out of the voids between the stones and
partly due to the inflow of water that leaks around the structure or flows into the canal.

138
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 24.1: inverted filter of graded gravel laying between the riparap
protection layer and original material (subgrade) in which canal is excavated

The protective construction as a whole and each separate layer must meet
the following 3 conditions:
1. Geometrical conditions: the protective construction as a whole and
each separate layer must be sufficiently permeable for water entering
the canal through its bed or bank.
2. Hydraulic conditons: the fine material from the underlying filter layer
must be sufficiently permeable to avoid pressure build up (such as the
exit gradient).
3. Stability conditions: fine material from the filiter layers or from the
subgrade must not be washed into the void of the covering layer.
To prevent the filter from dislocation under the surface flow, a revetment
(concrete blocks or riprap) is laid over the filter material.
To obtain a fair grain size distribution throughout the filter layers, each layer
should be sufficently thick as indicated in the following table:
Table 24.1: minimum layer thickness for filter construction made in the dry
Layer Minimum thickness (m)
Fine gravel 0.05 to 0.10
Gravel 0.10 to 0.20
Revetment Varies

To prevent from fine material of the filter being dragged into the holes in
the top layer of the filter, the gradation should be such that, while it allows free
flow of seepage water, the subgrade or foundation material does not penetrate
to clog the filter. For the stability of each layers (refer to figure 24.1), the
following requirements must be met:
1. d15 layer 2/d85 layer 1 ≤ 5.
2. d15 layer 1/d85 subgrade ≤ 5.

139
CdP User’s Manual February 2017 (second edition revised and expanded)

24.3.2. Concrete blocks revetment


Just beyond the u/s or d/s impervious floor, pervious revetment comprising
of cement concrete blocks of adequate size laid over the filter (packed stones1,
inverted graded filter) or fabric filter (geotextile) is to be provided, as shown in
the following figure:

Figure 24.2: u/s and d/s block protection works on filter

The concrete block size and thickness according to the canal discharge are
given by the following table:
Table 24.2: concrete block size according to canal discharge
Canal discharge Block size
(m3/s) (m3)
≤1 0.6 * 0.6 * 0.20
1 to 5 0.6 * 0.6 * 0.25
5 to 30 0.6 * 0.6 * 0.40
30 to 100 0.6 * 0.6 * 0.60

The gap between the blocks shall not be greater than 0.05 m and shall be
packed with pebbles.
The following figure presents the action of an inverted graded filter lying
under blocks in order to prevent the loss of soil through the joints.

1
Gravel or crushed rock.

140
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 24.3: flow lines with concrete blocks revetment on inveted graded filter

24.3.3. Riprap revetment


A riprap is a revetment of broken
rock or cobbles placed on a random
fashion. It is made from a variety of rocks
or stones types, as shown in the opposite
figure.
Figure 24.4: riprap revetment
Riprap is used where a rough
surface is needed to resist wave erosion, to protect against rainfall runoff
(gullying), and to dissipate high flow velocity energy as revetment in
embankment protection.
The stones should be predominantly angular in shape and nearly alike in all
directions so that they can easily move to lower depths without being hindered
by too large stone. River bed stone tend to be rounded and are more suitable
for masonry works and for filling gabion boxes.
24.3.4. Stone piching revetment
A stone pitching revetment is a regularly sized and shape stones or concrete
blocks placed in an ordered fashion as protection against erosion.
Pitching will be used
where a finished horizontal or
sloping surface is required as a
hydraulically smoother
protection than riprap. Slopes of
embankments should not be
steeper than 1:1.
Figure 24.5: pitching revetment Like for riprap, pitching is
placed on a sand/gravel bedding.

141
CdP User’s Manual February 2017 (second edition revised and expanded)

The large pitching stone are put first, with its longest axis perpendicular to the
surface. The smaller packing stone is driven in by hammer to support the
pitching stone and fill up some voids. The much smaller spalls are used for
wedging and to fill gaps to produce an even surface, without projection above
the neat lines shown on the figure. Surface grounting may be considered.
The diameter and weight of the pitching stones to be selected depend upon
the maximum flow velocity expected and the side slope of the embankment as
shown in the graph of the following figure:

Figure 24.6: size of pitching stone versus flow discharge rate

24.3.5. Boulders revetment


Boulders are basically rocks that are too big to pick up without mechanical
help. They are usualy used in embankments to break up the flow of storm-
water and to stop erosion.

142
CdP User’s Manual February 2017 (second edition revised and expanded)

24.4. LAUNCHING APRON

Beyond the filter protective works, a flexible launching apron is provided to


settle into the local scour hole formed around the structure in order to prevent
the erosion.
For practical reasons, the thickness of the flexible apron is the same as the
thickness of the filter system.
24.5. LENGTH OF PROTECTIVE WORKS

The length of protective works is determined by the scour depth D under


the original river bed (Lacey equation) with a safety factor (FoS) depending on
the type of structure (piers, river bank protection etc.).
Guidelines for the selection of the length of the protective works are given
below.
L = D * FoS (24.1)
Where:
L = length of protective layer (m).
FoS = safety factor (= 1,5 to 2 for d/s works; = 1.25 to 1.5 for u/s works).
D = refer to equation 19.1 (m).
For practical reasons, the thickness of the flexible apron is the same as the
thickness of the filter system.
24.6. ROCK SIZING EQUATIONS FOR REVETMENT

Several factors affect the stones or rocks size of a launching apron or a riprap to resist
the forces of flow which tend to move them. These factors are bottom flow velocity,
flow direction, turbulence and waves. Due to the possible combination of these factors,
the velocity at which the water strikes the riprap is rather unpredictable unless the
structure is tested in a laboratory.
The stability of a particular riprap particle is a function of its size, expressed either in
terms of its weight (specific gravity) or equivalent diameter. The rock sizing suitable for
rocks placed within a zone of highy turbulent water (immediately d/s of the end sil of
the energy dissipater) is given by the following formula:

d50 = 0.081 * V2.26 / (sr -1) (24.2)


Where:
d50 = nominal rock size (diameter) of which not more than 50% of the
rocks are smaller (m).
V = flow velocity (m/s).
sr = rock specific gravity (dimensionless).
The rock specific gravity is given in the following table:

143
CdP User’s Manual February 2017 (second edition revised and expanded)

Table 24.3: rock type and specific gravity


Type of rocks Specific gravity sr
Sand stone 2.2 to 2.4
Granite & limestone 2.6
Basalt 2.9

The following curve permits to determine the stone diameter of the


protective layer to be used d/sor u/s of the structure, in terms of flow velocity
leaving the structure1 based on discharge rate divided by cross-sectional area.
The average velocity above a end sill of the basin is calculated by dividing
the discharge by the cross-sectional area of flow above the end sill (if no stilling
basin is needed, the channel average velocity is used with Qmax). It is assumed
that d40/d50 = 0.75.

Figure 24.8: curve to determine stone size (d40 in m and weight in kg) as a
function of the average velocity of flow

With d40 (the particule average size wich 40% of the material is finer), more than
60% on the stone mixture of the protective layer should consist of stones of sizes
greater than d40, should be as homogeneous as possible in length, width and
thickness and should be of curve size (or curve weight or heavier).
The grading of the stone revetment should be as follows:
 Maximum stone size = 1.5*d50.
 Minimum stone size = 0.5*d50.

1To find the stone diameter, we should use the bottom velocity striking the protective layers. However, this
bottom velocity is rather unpredictable. For practical purposes, the average velocity is used.

144
CdP User’s Manual February 2017 (second edition revised and expanded)

25.TRANSITION STRUCTURES AND CHANNEL SIDE SLOPE


25.1. INTRODUCCION

Transition structures are required to guide the flow with a change of cross-
section (e. g. width, bottom etc.) designed to be accomplished in a short
distance with a minimum amount of flow disturbance and losses.
A transition in its general form may have a change of channel shape (from
the usual trapezoidal earth canal into the structure with usually vertical walls),
a provision of a hump or a depression and contraction or expansion of channel
width, in any combination. In addition, there may be various degrees of loss of
energy at various components.
A channel transition is usually designed so that the losses in the transition
are small. The form, friction and energy losses in the transition may be
neglected and, consequently, the energy equation is appropriate for the
analysis1 and implies conservation of total head HL.
The principal types of transition are as follows:
 Change in bed level with constant width.
 Change in channel width with constant bed level.
25.2. CHANGE OF SECTION AND SPECIFIC ENERGY

25.2.1. Change in bed level


The variation of water level is produced by a raised or a drop of the canal
bed level (ΔZ), with the channel width B remaining unchanged.
Upstream sub-critical flow
A sub-critical flow (with specific energy E1 and discharge per unit width q1)
is approaching a region where the bed is raised. No mechanical (friction) losses
are considered between sections 1 (before the hump) and 2 (on the hump)2.
The following situations occur for the water surface on the hump:
 ΔZ < yc: with a raised bed level (considered as a broad crested weir)
too low for critical flow to occur, the water level y2 at the constriction
drops with unaffected u/s level y1.or yinit) At this hump elevation, E2
decreases by ΔZ3, but still exceeds the minimum specific energy Emin
for this discharge. The flow remains sub-critical over the hump and

1 Energy losses if considered are consisting of friction loss and conversion loss. The friction loss is estimated by n
in the Manning’s formula (negligible). The conversion loss is expressed in terms of the change in the velocity head
(acceleration of the flow at inlet and drop in water surface; reduced flow velocity at outlet and rise in water
surface).
2
This region is sufficiently long for parallel flow to be established (hence “broad-crested”), but insufficiently long
for significant frictional losses.
3 Specific energies at sections 1 and 2 are given by: E = y + V2/2g and E = E – ΔZ.
1 2 1

145
CdP User’s Manual February 2017 (second edition revised and expanded)

resumes its original depth d/s of the hump (refer to figures 25.1a &
25.1b).
 ΔZ = yc: at this point, the critical height for the given discharge is
reached at the constriction. The flow reaches the critical energy E2 =
Emin = Ec (refer to figures 25.1a & 25.1b & 5.1; no backwater curve).
 ΔZ > yc: once critical flow is achieved at the constriction, if the hump
height is increased, the given flow will still go critical on the hump and
remain at the critical depth yc (it will not and cannot fall below this
value) as the specific energy E cannot be less than Emin as shown in the
below figure. To take into account ΔZ, only the u/s conditions can
change with an increase of specific energy and the u/s flow must “back
up” (refer also to figures 1.5 & 14.4).
What happens further d/s depends on other controls; the flow can be super-
critical or sub-critical.

Figure 25.1a: hump and water depth with u/s sub-critical

Upstream super-critical flow


A super-critical flow (with specific energy E1 and discharge per unit width
q1) is approaching a region where the bed is raised1. No mechanical (friction)

1
This region is sufficiently long for parallel flow to be established (hence “broad-crested”), but insufficiently long
for significant frictional losses.

146
CdP User’s Manual February 2017 (second edition revised and expanded)

losses are considered between sections 1 (before the hump) and 2 (on the
hump). The following situations occur for the water surface on the hump:
 ΔZ < yc: the water surface on the hump will rise due to a decrease in
the specific energy as shown in the following figure.
 ΔZ = yc: by raising more the hump, the critical height for the given
discharge will be reached. The flow will reach the critical energy E2 =
Emin = Ec (refer also to figure 5.1; no backwater curve).
 ΔZ > yc: once critical flow is achieved at the constriction, if the hump
height is increased even more, sub-critical flow will occur and a
hydraulic jump will take place.

Figure 25.1b: hump and water depth with u/s sub-critical or super-critical flow

Excessively tall broad crested weir is not a problem in terms of water


measurement or calibration; it is only troublesome with respect to
unnecessarily raising the u/s water surface elevation and increasing costs.
25.2.2. Change in channel width
The variation of water level is produced by a reduction/increase in the
channel width with the channel bottom remaining horizontal. It is convenient
to analysis the flow in terms of the discharge intensity q (refer to figure 9.3 with
constant specific energy).
Channel constriction with u/s sub-critical flow
With a horizontal contraction (B2 < B1), the specific energy at u/s section is
equal to the specific energy at d/s (constriction) since the bed elevation at both
sections is the same. Based on the constant specific energy graphs (refer to
figure 9.3), the water depth y2 decreases with an increase of q (q2 > q1) and Fr.
The limit of the contracted width is obviously reached when corresponding
to the critical depth yc where q = qmax for a given specific energy and critical-
147
CdP User’s Manual February 2017 (second edition revised and expanded)

flow condition at the constriction (Fr = 1). With a further reduction in the
channel width, the flow will not be possible with the given u/s conditions; the
u/s E will increase and the water level will have to raise provoking a decrease
in velocity (the water depth may be different but remains critical at the
constriction).
What happens further u/s or d/s depends on other controls and the flow
can be super-critical or sub-critical as shown in the following figure:

Figure 25.2: constriction and unit discharge with critical depth (u/s sub-critical
flow)

Channel constriction with u/s super-critical flow


If the u/s flow has super-critical flow conditions, a reduction of the channel
width (and hence an increase in the discharge intensity) causes a rise in water
depth.
Channel expansion
When the channel width increases, lower flow velocities take place in the
downstream section. The sediment transport capacity is smaller in the
downstream reach and accretion may take place 1.

1This geometry is well suited to the design of settling basins: e.g. at the intake of an irrigation channel, a settling
basin will trap sediment materials to prevent siltation of the irrigation system.

148
CdP User’s Manual February 2017 (second edition revised and expanded)

Table 25.1: summary table of changes in canal section


Type of Water depth y2 at
Section Characteristics
change new section
B2 = B1
ΔZ ≠ 0 y2 < y1 with u/s sub-
q2 = q1 (eq. 10.2b)
Raised canal critical flow
E2 < E1 (eq. 10.1)
bed level y2 > y1 with u/s super-
yc2 = yc1 (eq. 10.2a)
(E1 = E2 + ΔZ) critical flow
Emin2 = Emin1 (eq. 9.1a)
Constriction
B2 < B1
ΔZ = 0 y2 < y1 with u/s sub-
q2 > q1 (eq. 10.2b)
Decrease of critical flow
E2 = E1 (eq. 10.1)
canal width y2 > y1 with u/s super-
yc2 > yc1 (eq. 10.2a)
(E1 = E2) critical flow
Emin2 > Emin1 (eq. 9.1a) (1)
B2 = B1
ΔZ ≠ 0 y2 > y1 with u/s sub-
q2 = q1 (eq. 10.2b)
Drop in canal critical flow
E2 > E1 (eq. 10.1a)
bed level y2 < y1 with u/s super-
yc2 = yc1 (eq. 10.2a)
(E1 = E2 – ΔZ) critical flow
Emin2 = Emin1 (eq. 9.1a)
Expansion
B2 > B1
ΔZ = 0 y2 > y1 with u/s sub-
q2 < q1 (eq. 10.2b)
Increase of critical flow
E2 = E1 (eq. 10.1)
canal width y2 < y1 with u/s super-
yc2 < yc1 (eq. 10.2a)
(E1 = E2) critical flow
Emin2 < Emin1( eq. 9.1a) (2)
(1) Specific energy graph curve shifts to the right (q2 > q1) and up (Emin2 > Emin1)
(2) Specific energy graph curve shifts to the left (q 2 < q1) and down (Emin2 < Emin1)
25.3. TYPES OF TRANSITIONS

25.3.1. Transitions and wing walls


The variation of water level is produced by a reduction/increase in the
channel width with the channel bottom remaining horizontal. It is convenient
to use specific energy graph (refer to figure 9.2a).
The main designs of transitions are shown in the following figure:

149
CdP User’s Manual February 2017 (second edition revised and expanded)

Figure 25.3: isometric view of transitions

The most common transition structures used in the User’s Manual are the
abrupt transition wall (especially in the d/s diverging transition where energy
recovery is usually not necessary). This transition should be vertical walled with
walls at 45º to the canal center line. It may be used directly without elaborated
design. An apron, which is an integral part of the transition, protects from
erosion the channel bottom at the outlet1.
Special transition structures such as wrapped wall are useful when the
conservation of energy is essential because of allowable head water
considerations such as an irrigation structure in sub-critical flow.
It is left to preference of the engineer which type he/she prefers but
complex transition structures such as the warped type are costly and only
warranted if the least amount of energy loss must be afforded.
25.3.2. Head loss
If the losses with transition are accounting (E1 = E2+ HL), the head loss is
given by the following formula:

HL = K * V2 / 2g (25.1)
Where:
HL = energy head loss in transition (m).

1 Sudden expansion (or contraction) ratios like 1:1 or 2:1 are not very effective for energy conversion because the
high velocity jet leaving the throat cannot suddenly change direction to follow the boundaries of the transition.
In the flow separation zones that result, eddies are formed that convert kinetic energy into heat and noise.

150
CdP User’s Manual February 2017 (second edition revised and expanded)

K = coefficient depending of the type of transition1.


The transition head losses of inlet and outlets are summarized in the
following figure:

Figure 25.4: transition head losses of inlets and outlets


25.4. CHANNEL SIDE SLOPE

Erosion most commonly occurs when the velocity of flow exceeds the
velocity at which the soil of the channel will erode (refer to figure 8.1). Erosion
can be prevented by lowering the velocity below the soil-erosion velocity
(change of channel geometry), by lining the natural channel material with a
more erosion-resistant material or by changing the canal side slopes. The
following table shows the recommended canal side slopes.
Table 25.2: recommended canal side slopes
Type of channel Side slope (H:V)
Firm rock Vertical to ¼:1
Fissured rock ½:1
Stiff clay ½:1
Firm earth with stone lining 1:1
Firm earth and large channel 1:1
Firm earth and small channel 1½:1
Loose sandy earth 2:1
Sandy porous loam 3:1
25.5. STRUCTURE FREEBOARD

The freeboard of a channel is the vertical distance from the top of the
channel to the water surface at the design condition. The distance should be
sufficient to absorb sudden changes in water surface elevation due to errors in
water management or rainfall runoff entering the canal or prevent waves or
fluctuations in the water surface from overflowing the sides

1 For contraction: 0.1 < K < 0.6; for expansion: 0.3 < K < 0.8.

151
CdP User’s Manual February 2017 (second edition revised and expanded)

There are no universally accepted rules for the determination of a


freeboard, since wave action or water surface fluctuations in a channel may be
created by many uncontrollable causes.
The calculation of the freeboard is based on Qmax and the interpretation of
the following table containing freeboard in relation to design discharge rate.
Table 25.3: minimum freeboard for canal and structure
Discharge rate Total freeboard Lining freeboard
(m3/s) (m) (m)
0.01 0.30 0.15
< 0.5 0.40 0.20
0.5 – 1.5 0.50 0.20
1.5 – 5.0 0.60 0.25
5.0 – 10.0 0.75 0.30
10.0 – 15.0 0.85 0.40
> 15.0 1.00 0.50
> 30.0 1.00 0.60

Particularly for higher weirs in rivers with low embankments, it is important


to determine how high the embankment must be raised d/s to prevent flooding
that would normally not have occurred (backwater curve).

152
CdP User’s Manual February 2017 (second edition revised and expanded)

26.DESIGN PRINCIPLES OF GRAVITY WALLS


26.1. DESIGN PRINCIPLES

The four primary concerns for the design of nearly any retaining wall are:
1. That it has an acceptable FoS with respect to overturning.
2. That it has an acceptable FoS with respect to sliding.
3. That the allowable foundation soil bearing capacity is not
exceeded (bearing failure).
4. That the stresses within the components (stem and footing) are
within code allowable limits to adequately resist imposed vertical
and lateral loads.
The retaining wall design calculates the location of the resultant force, from
the soil backfill and the weight of the
wall1. This resultant force needs to
act within the middle third of the
retaining wall for stability, avoiding
tension in the masonry. The active
load against the retaining wall
combined with the dead weight of
the retaining wall and the resultant
Figure 26.1: retaining wall cross- force is then checked against the FoS.
section and acting forces Generally, free-standing gravity
retaining walls are economical for
retaining low walls, possibly up to 3 m.
26.2. PRESSURE

Retaining wall design considers all the forces acting on the wall, including
the effect of the water table. Pressures are shown in the following figure:

Figure 26.2: pressures on gravity wall

1 In the User’s Manual, the backfill material is typically coarse grained and non cohesive material.

153
CdP User’s Manual February 2017 (second edition revised and expanded)

26.3. VERTICAL PRESSURE FORCES

The acting vertical pressure forces are the soil and the wall components
weight. The resisting vertical pressure forces are developed in the underlying
soil. It varies uniformly between the toe and the heel of the footing and should
not exceed the allowable bearing capacity of the soil on which the footing is
built.
26.4. LATERAL PRESSURE FORCES

Knowing the properties of the soil behind the wall enables the engineer to
determine the lateral pressure distribution that has to be designed for. Lateral
earth pressure varies linearly with depth.
The relationship between the vertical earth pressure force (weight of soil)
and the lateral earth pressure force is through the appropriate earth pressure
coefficient Ka or Kp. The coefficients depend upon the shearing resistance of the
soil itself Ø as shown in the following simplified equations with a level back
slope of the embankment behind the wall.

Ka = (1-sinø)/(1+sinø) = tang2 (45° - ø/2) (26.1a)


and
Kp = 1 / Ka (26.1b)
Where:
Ka = coefficient of active pressure force.
Kp = coefficient of passive pressure force.
Ø = failure plane angle of internal soil friction or (critical) angle of
repose (degree).
The resultant pressure is located one-third of the height above the base of
the wall for triangular pressure diagram.
Because no adhesion or frictional forces (angle δ) are assumed to exist
between the soil and the wall, the lateral pressure is taken to act horizontally
and parallel to the surface of the backfill1 (Rankine’s theory with wall friction δ
= 0 and slope of back fill β = 0).
26.4.1. Active pressure forces
The total lateral active earth pressure (thrust) forces Pa on a vertical plane
(wall) above the point under consideration (here Ht at bottom from top)
develop when the wall is free to move outward and the soil mass stretches
sufficiently to mobilize its shear strength. The pressure force acts horizontally

1In practise, considerable friction may develop and, as a consequence, the earth pressure is inclined at a certain
angle to the normal to the wall. Rankine's assumption results is an overestimation of active earth pressure and
underestimation of passive earth pressure. This error is anyway on the safe side.

154
CdP User’s Manual February 2017 (second edition revised and expanded)

at a distant Ht/3 from the base of the wall 1 and is given by the following
equation (friction angle δ does not appear in equations):

Pa = ∑ C * Ka *γ * Ht2 (26.2)
Where:
C = coefficient (=1 for rectangular pressure diagram; = 0.5 for
triangular pressure diagram)
Pa = total lateral active (earth and water) pressure force on the wall
above the point under consideration (kN/m run of wall).
Ht = height (here total height) of backfill (m).
γ = unit or specific weight of soil or water (kN/m3).
26.4.2. Passive pressure forces
The total lateral passive earth pressure forces on the wall Pp occurs if the
wall moves towards the soil, then the soil mass is compressed which mobilizes
its shear strength and the passive pressure develops 2. The pressure force acts
horizontally at a distant usually Ht/3 from the base of the wall under
consideration3 and is given by the following equation (friction angle δ does not
appear in equations):

Pp = ∑ C * Kp *γ * Ht2 (26.3)
Where:
C = coefficient (=1 for rectangular pressure diagram; = 0.5 for
triangular pressure diagram)
Pp = total lateral active earth pressure above the point under
consideration (here at bottom from top) (kN/m run of wall).
γ = unit or specific weight of soil (kN/m2).
Ht = total height of passive earth layer under consideration (m).
26.5. BASIC INSTABILITY MODES

There are three basic instability modes to be checked: sliding, overturning


and soil bearing, as shown schematically in the pictures below.
26.6. SLIDING

The backfill and the other applied loads exert a lateral pressure against the
wall. This load actually pushes the wall out, so it tends to slide. This sliding force
is resisted by the friction between the soil and the footing and by the passive

1 When the lateral stress on the wall is triangular.


2 This situation occurs along the wall section below grade on the opposite side of the retained section of fill. In
the User’s Manual, it is ignored as additional restraint to lateral movement.
3 When the lateral stress on the wall is triangular.

155
CdP User’s Manual February 2017 (second edition revised and expanded)

pressure developed against the soil at the front of the wall. A portion of this
passive force is usually ignored to account for the
fact that the front soil may have been disturbed
during or after the construction. When more
sliding resistance is required, a shear key may be
provided. This component is very efficient since it
works in bearing against the foundation soil. The
Figure 26.3a: sliding factor of safety with respect to sliding equals the
resisting force divided by the driving force and the
minimum value should be 1.25 (refer to table 26.2).
To carry out this test, the horizontal and the vertical forces must be
calculated and the FoS verified (refer to table 26.2). In general, the structure
of the User’s Manual is very stable against sliding because of its wide base
combined with cut-off walls (or key walls), in one monolithic structure.
26.7. OVERTURNING

As a result of the lateral pressure forces on the back of the wall, a retaining
structure has the tendency to rotate outward about
the toe. The overturning moment from the applied
forces must be resisted by an opposite moment
produced by the vertical loads, including the wall
self weight and the weight of the backfill over the
heel, plus any surcharge. The factor of safety with
Figure 26.3b: respect to overturning is then defined as the
overturning resisting moment divided by the overturning
moment, and the minimum value should be 1.50
(refer to table 26.2).
To prevent the structure from overturning, the sum of the moments of
forces tending to resist overturning about 0 (Pp neglected in the User’s Manual)
must exceed the sum of the moment forces tending to overturn the structure,
with a safety factor overturning (refer to table 26.2).
26.7.1. Overturning moments
Overturning pressure forces include:
 Earth pressure (horizontal thrusts by soil).
 Hydraulic pressure (horizontal thrusts by water).
26.7.2. Stabilizing or resistance moments
Stabilizing or restoring pressure forces include:
 Structure weight.
 Soil weight above the heel.

156
CdP User’s Manual February 2017 (second edition revised and expanded)

The water weight is neglected in the resistance moment. Multiplying the


forces for their respective arms of leverage, the values of the overturning and
stabilizing moments are derived and the FoS checked (refer to table 26.2).
Providing a toe increases both the resistance to overturning and sliding. The
design is mainly concerned with keeping the resultant force within the middle
third of the base to prevent the base from losing contact with the soil at the
heel and tension in the foundation slab.
In general, the structure of the User’s Manual is very stable against
overturning and sliding because of its wide base (combined with cut-off walls)
in one monolithic structure.
26.8. SOIL BEARING
The bearing capacity of the soil must be adequate to support the structure.
Usually the bearing pressure under the footing is
trapezoidal with the maximum bearing pressure at
the end of the toe. It is important to keep the
resultant of the bearing pressure within the kern, so
that the minimum pressure at the heel side is not
negative, in which case the footing should be re-
proportioned.
Figure 26.3c: soil
The load per unit area of the foundation at
bearing
which shear failure in soil occurs (or the ultimate soil
bearing capacity) is given by the following table:
Table 26.1: bearing capacity for different types of rocks and soils
Bearing value
Category Types of rocks and soils
(kN/m2)
Dense gravel or dense sand and gravel > 600
Medium dense gravel, or medium dense sand &
350 to 600
gravel
Non Coarse river boulders and gravel, loose gravel or loose
350
cohesive sand and gravel
soils River gravel, boulder, sand 320
Alternate gravel and sand layers with silt content 210
Fine to medium sand, saturated: loose to medium
160
dense compact
Very stiff bolder clays & hard clays 300 to 600
Stiff clays 150 to 300
Cohesive
Firm clay 75 to 150
soils
Soft clays and silts < 75
Very soft clay Not applicable

157
CdP User’s Manual February 2017 (second edition revised and expanded)

26.9. DRAINAGE

Facilities for drainage from the retained soils are always provided. If water
pressure is allowed to accumulate behind a retaining wall, then the resulting
force along the back of the wall is increased considerably, as shown in the
following table:

Figure 26.4: hydrostatic pressure and drainage

Poor or inadequate drainage is the cause of nearly all retaining wall failures
when the soil is fully saturated and the resulting hydrostatic pressure pushes
the wall over. The drainage system is critical to the long term structural stability
of any retaining wall.
26.10. SAFETY FACTORS

The calculated and recommended safety factors used in the design of a


retaining wall in the User’s Manual are given by the following table:
Table 26.2: safety factors FoS
FoS
Movement FoS calculated
recommended
Overturning about > 1.50
Mresistance / Moverturning
the toe (1.50 to 2.50)
> 1.25 Friction coefficient*vertical
Sliding along base
(1.50) load/active earth pressure force
Bearing capacity 2.00 to 3.00 Bearing capacity / max. soil
failure foundation pressure

158
CdP User’s Manual February 2017 (second edition revised and expanded)

PART IX:

DID YOU OBSERVE THAT…..?

159
CdP User’s Manual February 2017 (second edition revised and expanded)

HYDRAULIC & ENERGY GRADE LINE

1. In uniform flow, the hydraulic grade line (or surface of water) is parallel
to the (total) energy grade line and to the channel bed.
2. In uniform flow, total mechanical energy falls gradually as energy is
lost through friction. But specific energy remains constant along the
channel because there are no changes in depth and velocity.
3. The total energy grade line lies over the hydraulic grade line (water
surface) by an amount equal to the velocity head. A turbine in the flow
decreases the energy grade line (total energy line) and a pump in the
line increases the energy grade line (total energy line).
4. The total mechanical gradient is the graphical representation of the
total head at any section of a conduct.
5. In an open channel, the specific energy is the total energy measured
with respect to the datum passing through the bottom of the channel
or the weir crest of a structure and not with any horizontal datum.
6. The water depth in a channel corresponding to the minimum specific
energy is known as critical depth.
7. The discharge in an open channel corresponding to critical depth is the
highest for a given specific energy.
8. It is important to draw a clear distinction between total mechanical
energy and specific energy. They are linked but they are quite different.
The total energy is measured from some fixed datum and its value can
only reduce as energy is lost through friction. Specific energy, in
contrast, is measured from the bed of a channel and so when the bed
level changes the specific energy also changes.
9. When there is a change in the bed level of a channel (e.g. when water
flows over a weir), there are also changes in the energy components
but the total mechanical energy remains the same. It means that
specific energy can rise as well as fall depending on what is happening
to the channel bed.
10. In a fluid flow, a particle may possess elevation energy, kinetic energy,
pressure energy and initial energy.
11. For conservation of energy, the losses are due to conversion of
turbulence to heat, sound and potential energy; it must account for
losses if applied over long distances. For conservation of momentum,
the losses are due to shear at the boundaries.
12. In problems involving the use of conservation of energy, the path taken
by the object can be ignored. The only important quantities are the
160
CdP User’s Manual February 2017 (second edition revised and expanded)

object's velocity (which gives its kinetic energy) and height above the
reference point (which gives its gravitational potential energy).

BERNOULLI’S EQUATION

13. For a perfect incompressible liquid flowing in a continuous stream, the


total energy of a particle remains the same, while the particle moves
from one point to another. This statement is called the Bernoulli's
equation. In other words, the main assumptions of Bernoulli's equation
are:
 Velocity of energy of liquid particle, across any cross-section
is uniform.
 No external force except the gravity acts on the liquid.
 There is no loss of energy of the liquid while flowing.
14. In the Bernoulli’s energy equation, two types of energy are considered:
gravitational potential energy and kinetic energy. The sum of the
energies is called the mechanical energy. If energy actually leaks from
the system via frictional head loss, the Bernoulli equation will overstate
the energy available to the flow and the related predictions of velocity
and depth will proportionately be in error.

FLOW

15. The flow in a channel is said to be non uniform when the liquid particles
at different cross-sections have different velocities.
16. The force present in a moving liquid is the inertia force, the viscous
force and the gravity force.
17. Reducing the hydraulic radius will decrease the flow velocity. This
decrease in hydraulic radius can be accomplished by increasing the
wetted perimeter in relation to the area. This can be done by widening
the channel, flattening the side slopes or widening the bottom. The
increases the wetted perimeter without materially increasing the area.
18. Our intuition often claims that the larger the flow depth, the larger the
specific energy. It is incorrect because the specific energy decreases as
the flow depth increases with super-critical channel flow.
19. Is the flow super or sub-critical? The answer comes from calculating
the normal depth of flow using a formula such as Manning’s and then
comparing with the critical depth.

161
CdP User’s Manual February 2017 (second edition revised and expanded)

20. The distinguishing difference between free and submerged flow in a


channel constriction is the presence of critical velocity in the vicinity of
the constriction (usually a little u/s of the narrowest section of the
constriction).

SLOPE

21. Two channels of the same slope can be classified differently (one mild
and the other steep) if they have different roughness and thus different
values of n.
22. In general, the slope of the free surface is not equal to the slope of the
bottom surface of the flow. However, in some situations, the conditions
are met to have both surfaces parallel and the flow is called uniform
flow.
23. The bottom slope alone is not sufficient to classify a downhill channel
as being mild, critical or steep.

DISCHARGE MEASURING

24. For steady, fully developed channel flow, the pressure distribution
within the fluid is merely hydrostatic, which means that the
streamlines are parallel. A hydrostatic pressure means that each water
particle pushes on the underlying particle with the same force,
resulting in a linear pressure distribution because the lowest particle
carries the accumulated weight of the water particles above.
25. For curved streamlines, the water pressure is no more hydrostatic.
26. The flow measuring devices used in the User’s Manual (broad crested
weir and gate) are based on the principles associated with rapidly
varied flow.

FROUDE NUMBER

27. The ratio of the inertia force to the gravity force is called the Froude
number.
28. If the forces are due to inertia and gravity and frictional resistance
plays only a minor role, the design of the channels is made by
comparing the Froude number.
29. The effect of gravity on the flow is introduced thanks to the Froude
number.

162
CdP User’s Manual February 2017 (second edition revised and expanded)

30. The Froude number is not a function of any temperature dependent


properties.
31. The higher the Froude number at the entrance of a stilling basin (Fr1):
a. The better the hydraulic jump is in energy dissipation.
b. The shorter the stilling basin.
32. To increase Fr1 by converting potential energy into kinetic energy:
a. Decreasing y1 with an expansion of the cross-section of the glacis.
b. Increasing the incoming velocity V1 with the appropriate slope of
the glacis.
33. Isabelle drops a large boulder in a channel, causing gravity water
waves to spread out from the source of the disturbance. With sub-
critical flow in the channel, the velocity of the stream Vstream is less than
the celerity of the wave or wave propagation cwave (Vstream < cwave) so
the gravity water waves can move u/s with a velocity of cwave – Vstream
and affect the water level there as well as the d/s conditions.
In super-critical flow, Vstream is greater than the wave speed cwave (Vstream
> cwave) so the wave or the ripples cannot travel u/s but will be swept
downwards with a velocity of Vstream – cwave.
With critical flow, Vstream = Cwave. In other words, the wave celerity sets
the boundary between sub-critical and super-critical flow for a given
flow.

HYDRAULIC JUMP AND SEQUENT DEPTH

34. The type of hydraulic jump that develops usually in the design of a
structure thanks to the User’s Manual is an oscillating jump.
35. The formation of a hydraulic jump on a sloping glacis as compared to
that on a horizontal floor is more definite and less efficient.
36. The sequent depth y2 depends upon Fr1 and y1
37. The ideal condition for energy dissipation in the design of a stilling
basin is the one when the tailwater rating curve coincides with the
jump rating curve at all discharges.
38. With increasing jump submergence, the efficiency of energy
dissipation is reduced.
39. The exit flow d/s of the stilling basin is sub-critical and, hence, it is
controlled by the d/s flow conditions i.e. by the tailwater flow
conditions. The location of the jump is determined by the u/s and d/s
flow conditions and can be at slope break point (toe of glacis), u/s of
break point or d/s of break point.

163
CdP User’s Manual February 2017 (second edition revised and expanded)

40. Deepening of the cistern level excessively is not hydraulically efficient,


though it might be required if there is considerable uncertainty as to
the stability of the tailwater channel.
41. As an alternative to the deepening of the cistern level, a wider basin
might be considered, providing a shallower basin.
42. Hydraulic jump action in a trapezoidal basin is less complete and less
stable than it is in a rectangular basin.
43. By changing the slope of the d/s channel and, therefore, changing yn,
the location of the hydraulic jump formation (and it submergence) will
change.
44. When a hydraulic jump occurs in a channel, conjugate depths have the
same pressure-momentum force (the momentum of the flow u/s and
that d/s of the jump are equal) but the energy is not equal. Alternate
depths however have the same specific energy. Two conjugate depths
can never be alternate depths or vice versa.

PERMEABLE FOUNDATION

45. The Khosla’s theory is applicable to structures founded on permeable


material only.
46. The safety of a hydraulic structure founded on pervious foundation can
be ensured by:
a. Providing sufficient length of its concrete floor.
b. Providing a d/s cut-off of reasonable depth.
47. The Khosla’s critical Ge is less for more porous soils.
48. The exit gradient Ge is independent of the depth of the u/s cut-off wall.
49. Seepage endangers the stability of a structure built on permeable
foundation because of piping which depends upon the value of the exit
gradient Ge.
50. The design of the d/s cut-off wall of a structure on permeable
foundation depends upon:
a. Scour depth.
b. Exit gradient Ge.
51. While designing a structure on permeable foundation, the correction
for mutual interference of cut-off walls is usually not applicable on an
intermediate cut-off wall if the outer pile goes only just as deep as the
intermediate pile and is within a distance of one and half times its own
length.
52. U/s and d/s cut-off walls are meant to control the uplift pressure and
the exit gradient (piping) respectively.
164
CdP User’s Manual February 2017 (second edition revised and expanded)

53. By increasing the impervious floor of the u/s reach of the structure, the
uplift pressure force will decrease below the d/s impervious floor.
54. The thickness of the u/s impervious floor of the structure is nominal
because the net uplift pressure governs it.
55. The term of piping used in connection with the type of hydraulic
structures presented in the User’s Manual is associated with failure
initiated by the sand boiling phenomenon.

PROTECTIVE WORKS & SCOUR DEPTH

56. The inverted filter should be such as:


a. To let out the residual seepage.
b. Not to allow the soil particles to escape.
c. Not to get clogged in itself.
57. An aggrading river is a silting river.
58. The river reach u/s of a newly build cross-structure usually behaves as
an aggrading river.
59. The scour depth as measured below the highest flood level in a river or
channel will be less in a boulder river than in an alluvial river and will
be more in fine silt than in core silt.
60. Ruptures of the d/s impervious floor of the structure can be due to:
a. Insufficient length of the u/s impervious floor.
b. Insufficient length of the d/s impervious floor.
c. Insufficient depth of the d/s pile.
d. Choking of the d/s inverted filter.

ORIFICE FLOW AND NUMBER OF GATES OF THE HEAD REGULATOR

61. For a same water requirement, the number of gates of the head
regulator can be increased if necessary by raising the gate sill elevation
and vice versa.
62. The value of the coefficient of discharge is less than the value of the
coefficient of velocity.
63. The coefficient of discharge is not a simple constant number like the
coefficient of contraction.
64. The discharge through a totally submerged orifice is directly
proportional to the square root of the difference in elevation of water
surface.

165
CdP User’s Manual February 2017 (second edition revised and expanded)

RETAINING WALL

65. As the wall moves away from the soil backfill, the active condition
develops and the lateral pressure against the wall decreases with wall
movement until the minimum active earth pressure force is reached.
66. As the wall moves towards (into) the soil backfill, the passive condition
develops and the lateral pressure against the wall increases with wall
movement until the maximum passive earth pressure is reached.
67. If y is the depth of water retained by a vertical wall, the height of centre
of pressure above the bottom is y/2.

166
CdP User’s Manual February 2017 (second edition revised and expanded)

PART X:

PERSONAL NOTES

167
CdP User’s Manual February 2017 (second edition revised and expanded)

168
CdP User’s Manual February 2017 (second edition revised and expanded)

169
CdP User’s Manual February 2017 (second edition revised and expanded)

170
CdP User’s Manual February 2017 (second edition revised and expanded)

LIST OF PRINCIPAL SYMBOLS

A m2 Wetted cross-sectional area


b m Total solid floor length
B m Overall waterway (open water) width between
river banks or canal banks (across the flow)
Bcl m Clear or effective waterway (open water) width
at gate section or chamber (across the flow)
Bstructure m Overall waterway (open water) width of
structure between abutments (across the flow)
Cc X Contraction coefficient
Cd X Discharge coefficient
Ce X Effective discharge coefficient
Cv X Velocity coefficient
d m Depth of cut-off wall or pile
d/s X Downstream
d1 m U/s vertical cut-off wall depth
d2 m D/s vertical cut-off wall depth
d50 mm Average particle size or diameter of the channel
bed material of which 50 % of the mixture is
finer by weight. Sieve size through which 50%
of the material passes by weight
D m Depth of scour measured from channel bed
level
D m Depth of cut-off wall or pile, the influence of
which has to be determined on the neighboring
pile of depth d.
DFR m Depth of the application point of the resultant
hydrostatic force acting on a gate
E m (specific) energy
Ef1 m Estimated specific energy head in super-critical
flow range of hydraulic jump as expressed in
relation to the canal or channel bottom
Ef2 m Estimated specific energy head in sub-critical
flow range of hydraulic jump as expressed in
relation to the canal or channel bottom
Emin m Calculated minimum specific energy for a set
discharge rate
E1 m Calculated specific energy head per weight of
the fluid in super-critical flow range of hydraulic

171
CdP User’s Manual February 2017 (second edition revised and expanded)

jump as expressed in relation to the canal or


channel bottom
E2 m Calculated specific energy head per weight of
the fluid in sub-critical flow range of hydraulic
jump as expressed in relation to the canal or
channel bottom
f X Drowned flow reduction factor
f X Lacey silt factor
FoS X Factor of safety
Fr1 X Froude number of super-critical reach of
hydraulic jump (ratio of inertial forces to
gravitational force in a system)
Fr2 X Froude num of sub-critical reach of hydraulic
jump (ratio of inertial forces to gravitational
force in a system)
g m/s2 Gravitational acceleration
Ge X Exit gradient
hes m End sill height of the hydraulic structure in
relation to d/s channel or canal bed elevation
hj m Height of the hydraulic jump
h0 m U/s water head in relation to weir sill elevation
h2 m D/s water head in relation to weir sill elevation
Hd max m Maximum dynamic head (with flow)
Hs max m Maximum hydrostatic head (without flow)
H0 m Total u/s energy head in relation to the weir sill
elevation
H1 m Total energy head u/s of jump in relation to the
weir sill elevation
H2 Total d/s energy head in relation to the weir
crest
HL m Total energy head loss
HE0 m Total u/s energy elevation including kinetic
energy (energy grade line)
HE1 m Total energy elevation including kinetic energy
(energy grade line) at u/s jump formation
HE2 m Total energy elevation including kinetic energy
(energy grade line) at d/s jump formation
L m Flow wise length of crest or protective layer
n X Manning’s roughness or friction coefficient
Nsc X Number of side contraction of a weir
q m3/s.m Unit discharge rate per m width of structure,
stream or channel
Q m3/s Discharge rate in channel
Qcrop m3/s Discharge rate corresponding to irrigation
water need or water duty
172
CdP User’s Manual February 2017 (second edition revised and expanded)
3
Qmax m /s Flood discharge rate in channel
Qmin m3/s Low water flow rate in channel
R m Lacey’s normal or mean scour depth measured
from high flood level
sr X Specific gravity of rock
S m/m Channel slope
t m Thickness
u/s X Upstream
V m/s Fluid velocity
Va m/s Flow velocity of approach
V2/2g m Kinetic energy head due to flow velocity
V1 m/s Velocity of super-critical flow
V2 m/s Velocity of sub-critical flow
V3 m/s Flow velocity d/s hydraulic structure
w m Under-flow gate opening height
wcrop m Under-flow gate opening height producing,
after contraction, a flow satisfying the field
water requirements
y m Depth of flow
ygate m Depth of water from water surface to centre of
a closed gate
yn m Normal depth of flow
ys m Water depth with the presence of a structure
y1 m Super-critical initial depth of flow
y2 m Sub-critical sequent depth of flow
Z m Elevation of datum
Zc m Crest elevation to cistern elevation (drop
height)
γ N/m3 Specific or unit weight
ΔE m Mechanical head loss over the structure
Δh m Head loss (water depth change) over structure
in relation to the weir sill
Φ % Hydrostatic pressure force
φt % Residual hydrostatic or dynamic pressure force
at toe of glacis

173
CdP User’s Manual February 2017 (second edition revised and expanded)

174
CdP User’s Manual February 2017 (second edition revised and expanded)

GLOSSARY
Abutments: walls that flank the edge of a weir or other hydraulic structure, and
which support the river banks on each side of the weir.
Accretion: process by which particles carried by the flow of water are deposited
and accumulate (opposite of erosion).
Afflux: rise in water level.
Aggradation: general or progressive rise of the bed level of a channel by the
accumulation of sediments (silting) (opposite of degradation).
Angle of internal soil friction: steepest angle of descent of a granular material
relative to the horizontal plane to which a material can be piled without
slumping.
Apron: a layer of stone, concrete or other scour protection placed on the
channel bed in the vicinity of a hydraulic structure.
Backfill: soil placed behind a wall.
Backwater effects: effects in sub-critical flow that flow conditions in one
location have on flow conditions farther upstream (in particular, the water
surface elevation u/s of a weir).
Baffle pier: block placed in intermediate position across the stilling basin.
Bank: the edge of a river or stream. Note that left and right refer to the river
viewed looking downstream.
Bank protection: works to protect a bank from erosion or undermining by
scour.
Boil: concentrated outflow of seepage water, for example through a crack
channel or a hole in the sub-soil (sand boil is a type of boil which carries out
sand out of the substrate).
Braided river: alluvial river having two or more channels that form a braided
pattern and whose size, length and transverse pattern tend to vary
considerably in successive floods.
Broad crested weir: weir with a crest section of significant length or thickness
measured in the direction of flow.
Centroid: the centre of mass (or gravity) of a geometric object of uniform
density.
Channel: natural open watercourse that contains and conveys water.
Chute blocks: blocks placed at the entrance of the stilling basin to form a
serrated device.
Control section: control section of a measuring structure is located where
critical flow occurs and sub-critical, tranquil, or streaming flow passes into
super-critical, rapid, or shooting flow.
Crest (of weir): top part of weir. The level of the crest, its length and its cross-
sectional shape determine the discharge (flow) characteristics of the weir.
175
CdP User’s Manual February 2017 (second edition revised and expanded)

Critical angle of repose: steepest angle of descent of a granular material


relative to the horizontal plane to which a material can be piled without
slumping.
Critical depth: depth representing the minimum specific energy (occurs when
the Froude number equals 1.0).
Cross-regulator or weir: an artificial obstruction or constriction in any
watercourse that results in increased water surface level upstream or used to
measure the discharge.
Cut-off wall: barriers constructed vertically either of reinforced concrete,
masonry or of steel sheet pile, and provided at the bottom of the structure to
protect it by extending the line of seepage against scours and possible piping
due to excessive exit gradients of the seepage flow below the foundations.
Degradation: the processes of progressive lowering or drop in the bed of a
channel by erosion (scouring) (opposite of aggradation).
Depth of flow: vertical distance from the bed of a channel to the water surface.
Design discharge rate: peak flow at a specific location defined by an
appropriate return period to be used for design purposes.
Discharge rate: flow rate expressed in volume per unit time (typically m3/s). In
this Manual, the word “flow” is used to mean flow rate or discharge.
Discharge intensity: discharge per unit length of weir or stream (see also unit
discharge).
Drowned: in the context of weir hydraulics, a weir is said to be drowned (or
drowned out) when the downstream water surface elevation rises to the point
where it begins to affect flow over the weir.
End sill: vertical, stepped, sloped, solid or dentated wall constructed at the
downstream end of the stilling basin.
Exit gradient: hydraulic gradient in the sub-soil at the location of the exit point.
Energy grade line or total energy line: profile of the free flow surface flowing
along the structure or channel specifying the total energy head available at any
point of the structure or channel. The graphical representation of the elevation
of the line at any point is the sum of the datum head, the flow depth and the
velocity head. In uniform flow (Manning’s equation), the energy grade line
(total energy line) is parallel to the hydraulic grade line (water surface).
Erosion: process by which material forming the bed or banks of channel is
removed by the action of flowing water or waves (opposite of accretion).
Filter: granular or fabric material placed between the erosion protection
revetment (riprap, blocks etc.) and the underlying soil surface to prevent soil
movement into and through the revetment. Riprap and concrete block layer
should have a filter placed under it in all cases.

176
CdP User’s Manual February 2017 (second edition revised and expanded)

Flood bank: embankment, usually earthen, built to prevent or control the


extent of flooding.
Flow: flow rate or discharge.
Freeboard: height of the top of a bank, flood bank, or structure above the level
of the water surface.
Froude number: dimensionless parameter representing the ratio of the velocity
wave relative to inertia and gravity forces in a fluid, taking the value of unity for
critical flow. It is an indication of the energy in a channel at a certain point.
Geotextile: permeable synthetic fabric filter that prevents migration of the fine
soil particles through voids in the structure, that distributes the weight of the
armor units to provide more uniform settlement and that permits relief of
hydrostatic pressures with the soils.
Glacis: downstream sloping face of a weir, between the weir crest and the
stilling basin.
Gradation: size distribution of a particular layer (riprap etc.) or aggregate.
Material that is well graded has a uniform distribution of sizes, within a given
minimum and maximum range.
Grade: inclination or slope of a stream channel, energy etc. usually expressed
in terms of the ratio or percentage of number of units of vertical rise or fall per
unit of horizontal distance.
Head (of water)1: the height of water surface elevation above a datum (such as
the weir crest).
Head loss: the drop in water surface elevation across a weir or other hydraulic
structure.
Heel: portion of the footing extending behind the wall (under the retained soil).
Hydraulic depth: average depth for a river cross-section.
Hydraulic grade line: surface or profile of water flowing in an open channel
specifying the sum of the static pressure and the elevation head Z at any point
of the structure or channel. The difference between the energy grade line and
the hydraulic grade line is equal to the dynamic (velocity) head v2/2g. In uniform
flow, the hydraulic grade line is parallel to the water surface and is also
represented by the bed of channel.
Hydraulic gradient: slope of the hydraulic grade line (quotient of the difference
in head between the two points and the distance between those points).
Hydraulic head: difference in head between two points.
Hydraulic jump: abrupt rise in water surface elevation when flow changes from
a super-critical to a sub-critical state, with associated dissipation of energy.

1 There are more technically precise definitions of head, making the distinction between static head, velocity head
and total head.

177
CdP User’s Manual February 2017 (second edition revised and expanded)

Invert level: level of the lowest point of a cross-section of a natural or artificial


channel or a structure.
Jet: water exiting an orifice.
Kinetic energy: energy that a system possesses as a result of its motion relative
to a reference frame.
Launching apron: apron of riprap or other material that subsides as scour
occurs to prevent scour undermining a structure.
Modular flow: condition in which flow is able to discharge freely over a weir,
resulting in a unique relationship between flow rate and referenced upstream
water surface elevation (Upstream head).
Modular limit: the maximum submergence ratio for which a weir will operate
with critical depth flow in the throat reach.
Nappe: over falling sheet of water passing over a weir crest and plunging into
the stilling basin. Term normally only applied where the jet is not in contact
with the weir structure (i.e. there is an air gap between the underside of the
nappe and the downstream face of the weir).
Normal depth: water depth corresponding to the minimum energy level in
conditions of normal flow.
Non modular flow: condition in which flow is not able to discharge freely over
a weir, with the downstream water surface elevation influencing the upstream
level (i.e. drowned flow).
Pitching: regularly sized and shape stones or concrete blocks placed in an
ordered fashion as protection against erosion.
Profile: flow profile represents the surface curve of the flow
Rating curve: a plot of water head against flow rate for a channel, weir or other
hydraulic structure (also called a stage-discharge curve).
Regulator: hydraulic structure for controlling water heads or division of flow.
Regime channel: state in which a channel or river formed in erodible material
which has reached a stage of virtual equilibrium (has adjusted its gradient and
cross-section to an equilibrium condition) with no long-term aggradation or
degradation (accretion or erosion), but which may be subject to meandering.
Retaining wall: any constructed wall that restrains soil or other material at
locations having an abrupt change in elevation.
Riprap: broken rock or cobbles placed on a random fashion as protection
against the action of water.
Roller: the large-scale turbulence region of a stilling basin.
Scour (local): scour that results directly from the impact of individual structural
elements (pier, abutment etc.) on the flow and occurs immediately in the
vicinity of those elements.

178
CdP User’s Manual February 2017 (second edition revised and expanded)

Scour (natural): erosion resulting from the shear forces associated with flowing
water or wave action.
Sediment: erodible material forming bed or banks of channel, which may be
eroded or deposited depending on the prevailing flow conditions.
Short crested weir: streamline curvature above the weir crest has a significant
influence on the head-discharge relationship of the structure.
Side weir: weir installed in a channel to divert part of the approach flow into a
separate spill channel.
Sill: top of an embedded structural member on which a gate rests when in
closed position.
Siltation: the deposition of sediment.
Specific energy: for a given cross-section, total mechanic energy per unit weight
of water as expressed in relation to the channel bottom; it is the sum of the
water depth and the velocity head, provided the streamlines are straight and
parallel.
Seepage: water which flows through the sub-soil, as a consequence of the
upstream existing hydraulic head.
Stilling basin: an energy dissipater comprising a basin in which a hydraulic jump
occurs.
Steady flow: flow with streamlines parallel and hydraulic characteristics
remaining constant for the time interval under consideration.
Streamline: path followed by a molecule of water.
Sub-critical flow: the flow in a channel at less than critical velocity.
Super-critical flow: the flow in a channel at greater than critical velocity.
Tailwater level: the water surface elevation downstream of a hydraulic
structure.
Toe: portion of footing which extends in front of the front face of the stem
(away from the retained earth).
Uniform flow: flow of water in a channel in which the depth and the velocity
remain constant along the channel.
Water level: the elevation of the water surface.
Wing wall: a wall on a weir or other hydraulic structure that ties the structure
into the river bank.

179
CdP User’s Manual February 2017 (second edition revised and expanded)

180
CdP User’s Manual February 2017 (second edition revised and expanded)

REFERENCES
Bureau of Reclamation, 1987. Design of Small Canal Dams. United States
Department of the Interior. Denver (Colorado), United States.

Chow, V. T., 1959. Open Channel Hydraulics. Mac Graw Hill, United States.

dePatoul, C., 1981. Design Manual of Small Scale Irrigation Structures. Besut
Irrigation Project. Besut, Malaysia.

dePatoul, C., 1990. Manual de Diseño para Obras Hidráulicas y Edificios Rurales.
Proyecto de Consolidación de la Reforma Agraria CORASUR. Choluteca,
Honduras.

dePatoul, C., 1992. Critères de Design pour les Ouvrages Hydrauliques. Projet
National d'Infrastructure Rurale. Conakry, Guinée.

Garg, S. K., 2007. Irrigation Engineering and Hydraulic Structures. Delhi, India.

ILRI, 1989. Discharge Measurement Structures. Publication 20. Wageningen,


Netherlands.

IIT Kharagpur. Version 2 CE; Modules 1 to 4. India.

Mott MacDonald & Partners, 1985. Design of Canal Structures. Working Papers,
Cambridge, United Kingdom.

P-ARBP, 2013. Irrigation Design Manual. Landell Mills, Kabul, Afghanistan.

P-ARBP, 2013. Construction Management and Supervision Manual for the Panj-
Amu River Basin Programme (P-ARBP). Landell Mills, Kabul, Afghanistan.

United States Department of Transport, Federal Highways Department.


Hydraulic Engineering Circular 14 (third edition). National Highway Institute,
United States.

Glenn Moglen:
https://www.youtube.com/watch?v=KH575pF7xUc&list=PLKHS1dE68m6_pzg
xcBp_8XeukzkeTe2X8&index=1

181

You might also like