You are on page 1of 110

BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE i

GOVERNMENT OF BALOCHISTAN
BALOCHISTAN COMMUNITY IRRIGATION AND AGRICULTURE PROJECT

DESIGN MANUAL
PART 6 – IRRIGATION STRUCTURES

TABLE OF CONTENTS

1 INTRODUCTION ................................................................................................. 1
1.1 Scope of Irrigation Structures Design Manual ................................................................... 1
1.2 Design Considerations ...................................................................................................... 1
2 ENERGY LOSSES THROUGH STRUCTURES .................................................. 3
2.1 Losses due to Bridge or Aqueduct Piers ........................................................................... 3
2.2 Losses through Trash Racks............................................................................................. 4
2.3 Losses through Transitions ............................................................................................... 5
2.3.1 General.................................................................................................................. 5
2.3.2 Losses between a Trapezoidal Channel and a Conduit ........................................ 5
2.3.3 Losses through Horizontal Channel Transitions.................................................... 6
2.3.4 Losses through Vertical Channel Transitions ........................................................ 6
2.4 Losses in Bends of Closed Conduits ................................................................................. 8
2.5 Energy Losses through Friction ......................................................................................... 9
3 REGULATING STRUCTURES .......................................................................... 13
3.1 Introduction ...................................................................................................................... 13
3.2 Discharge Characteristics ............................................................................................... 13
3.3 Checks............................................................................................................................. 14
3.3.1 Purpose ............................................................................................................... 14
3.3.2 Design Considerations ........................................................................................ 14
3.4 Check Drops .................................................................................................................... 15
3.5 Turnouts .......................................................................................................................... 15
3.5.1 Purpose ............................................................................................................... 15
3.5.2 Design Considerations ........................................................................................ 15
3.6 Division Structures ........................................................................................................... 16
3.6.1 Purpose ............................................................................................................... 16
3.6.2 Design Considerations ........................................................................................ 16
3.6.3 Time Division Structures ..................................................................................... 16
3.6.4 Flow Division Structures ...................................................................................... 18
3.7 Gates for Small Division Structures ................................................................................. 19
3.8 Canal Protection at Regulators........................................................................................ 20
4 DROPS AND CHUTES ...................................................................................... 31
4.1 Introduction ...................................................................................................................... 31
4.2 Construction Materials ..................................................................................................... 32
4.3 Vertical Drops .................................................................................................................. 32
4.3.1 Hydraulic Design ................................................................................................. 32
4.4 Chutes ............................................................................................................................. 36
4.4.1 Description .......................................................................................................... 36
4.4.2 Design Considerations ........................................................................................ 36
4.5 Stability and Structural Considerations ............................................................................ 37
5 AQUEDUCTS .................................................................................................... 41
5.1 Introduction ...................................................................................................................... 41
5.2 Construction Materials ..................................................................................................... 41
5.3 Aqueduct Configuration ................................................................................................... 42
5.4 RCC Trough Aqueduct - Hydraulic Design ...................................................................... 45
5.5 RCC Trough Aqueduct - Structural Design ..................................................................... 46
PAGE ii BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

5.6 Steel Pipe Aqueducts ...................................................................................................... 50


5.7 Design of Columns and Foundations .............................................................................. 51
5.7.1 Masonry Columns ............................................................................................... 51
5.7.2 Reinforced Concrete Columns ............................................................................ 53
5.8 Abutments ....................................................................................................................... 54
6 INVERTED SYPHONS ...................................................................................... 65
6.1 Purpose and Description ................................................................................................. 65
6.2 Hydraulic Design.............................................................................................................. 65
6.3 Design Features .............................................................................................................. 66
7 ROAD CROSSINGS: CULVERTS AND BRIDGES ........................................... 71
7.1 Introduction ...................................................................................................................... 71
7.2 Culverts ........................................................................................................................... 71
7.2.1 Hydraulic Characteristics .................................................................................... 71
7.2.2 Design Considerations for Concrete Pipe Culverts ............................................. 72
7.2.3 Design Considerations for Box Culverts ............................................................. 78
7.3 Bridges ............................................................................................................................ 79
7.3.1 Road Bridges....................................................................................................... 79
7.3.2 Footbridges ......................................................................................................... 80
8 CROSS DRAINAGE STRUCTURES ................................................................. 91
8.1 Introduction ...................................................................................................................... 91
8.2 Cross Drainage Culvert Structures .................................................................................. 91
8.2.1 Design Process for Pipe Culverts ....................................................................... 91
8.2.2 Design Process for Box Culverts ........................................................................ 92
8.3 Superpassages ................................................................................................................ 93
8.4 Overchutes ...................................................................................................................... 94
8.5 Drain Inlets ...................................................................................................................... 94
9 ESCAPES .......................................................................................................... 99
9.1 General ............................................................................................................................ 99
9.2 Side Channel Spillways ................................................................................................... 99
9.2.1 General................................................................................................................ 99
9.2.2 Design Considerations ........................................................................................ 99
10 SOCIAL STRUCTURES .................................................................................. 101
10.1 Community Participation in Site Selection ..................................................................... 101
10.1.1 Involvement of Farmers ..................................................................................... 101
10.1.2 Involvement of Women...................................................................................... 101
10.1.3 Involvement of Other Groups ............................................................................ 101
10.2 Wuzu Structures ............................................................................................................ 101
10.3 Animal Drinking Troughs ............................................................................................... 101
10.4 Washing Structures ....................................................................................................... 102

BIBLIOGRAPHY
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE iii

GOVERNMENT OF BALOCHISTAN
BALOCHISTAN COMMUNITY IRRIGATION AND AGRICULTURE PROJECT

DESIGN MANUAL

CONTENTS OF OTHER PARTS

Part 1 Site Investigations


Part 2 Flood Estimation
Part 3 Weirs
Part 4 Infiltration Galleries
Part 5 Irrigation Canals
Part 6 Irrigation Structures
Part 7 Flood Protection Structures
Part 8 Potable Water Supply Systems
Part 9 Structural Design Criteria
Part 10 Draughting Standards
Part 11 Value Engineering
Part 12 Selected Drawings

Annex 1 Monthly Rainfall Data

DISCLAIMER

This Design Manual was prepared under the Balochistan Community Irrigation and Agriculture Project (BCIAP) for
the design of schemes constructed under the Project. While every effort to check for mistakes in this manual has
been made, no liability for the use of this Manual for any other purpose can be accepted by BCIAP, or the Project’s
Consultants.

No credit is claimed here for original research or thought. As far as possible all reference material has been quoted
and acknowledged in the appropriate places.
PAGE iv BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

GOVERNMENT OF BALOCHISTAN
BALOCHISTAN COMMUNITY IRRIGATION AND AGRICULTURE PROJECT

DESIGN MANUAL

Conversion Factors

Length
1 inch = 25.4 mm
1 foot (12 inches) = 0.3048 m
1 mile (5280 ft) = 1609 m

Area
1 ft2 = 0.093 m2
1 acre (43,560 ft2) = 0.4047 hectares (4047 m2)
1 sq. mile (640 acres) = 259 hectares

Volume
1 ft3 = 0.028 m3
35.315 ft3 = 1 m3 (=1,000l)
1.0 Imp. gallon (=0.16 ft3) = 4.546 l
1.0 US gallon = 3.785 l

Discharge
1 cusec (ft3/s) = 0.028 cumecs (m3/s)
1 Imp. gallon/minute = 0.076 l/s

Weights
1 lb = 0.454 kg
2.2 lb = 1.0 kg
1 ton (US) = 907.2 kg (0.907 tonnes)

Force
0.2248 lbf = 1 N (0.1020 kgf )
0.06852 lbf/ft = 1 N/m (0.1020 kgf/m)
145.0 lbf/in2 = 1 N/mm2 (10.20 kgf/cm2)

Moment
0.7376 lbf ft = 1 Nm (0.1020 kgf m)

Useful Data
Density of Water = 1,000 kg/m3 = 62.4 lb/ft3
Nominal weight of reinforced concrete = 23.6 kN/m3 (2,400 kg/m3) = 150 lb/ft3
Nominal elastic modulus of concrete = 14 kN/mm2 (140 x 103 kg/cm2) = 2 x 106 lb/in2
Co-efficient of Linear expansion of concrete = 10 x 106 per oC =5.5x106peroF
Acceleration of gravity, g = 9.806 m/s2 = 32.3 ft/s2
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 1

1 INTRODUCTION

1.1 Scope of Irrigation Structures Design Manual


This Part of the Design Manual is intended to provide a guide to the engineering design of
hydraulic structures commonly required for perennial and flood irrigation schemes. It does not
cover the design of headworks, which are covered in Part 3: Weirs and Part 4: Infiltration
Galleries. It is based to some extent on the BMIADP Perennial Irrigation Canals and Canal
Structures Design Manual published in December 1991.

The structures covered by this Part of the Design Manual include:


• regulating structures (outlets and division structures);
• conveyance structures (drops, aqueducts and inverted syphons);
• protective structures (cross drainage structures and escapes); and
• social structures (wuzu, animal drinking structures and washing structures).

1.2 Design Considerations


Designs of irrigation structure take into account hydraulic, stability and structural design
considerations.

Hydraulic design provides for:


• adequate discharge capacity;
• adequate built-in overflow capacity;
• adequate energy dissipation measures where required;
• structural proportioning to minimise hydraulic headloss were required.

Stability design provides adequate structural dimensions so that for most soil foundation
materials, the structure will:
• resist sliding and overturning;
• prevent percolation water removing foundation materials;
• prevent undermining of the structure;
• provide foundation pressures less than the maximum allowable bearing pressure.

Structural design provides appropriate concrete thicknesses and reinforcement steel patterns for
structural members to resist bending moment, thrust and shear stresses imposed by loads on the
structure.

Stability and Structural design criteria are discussed in Part 9: Structural Design Criteria.
Hydraulic design criteria common to a range of structures covering discharge characteristics and
energy dissipation are included in Part 3: Weirs, while energy losses through structures are
covered here in Chapter 2.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 3

2 ENERGY LOSSES THROUGH STRUCTURES

2.1 Losses due to Bridge or Aqueduct Piers


Bridge or aqueduct piers form obstructions to the flow and cause an increase in upstream
depth. This increase in upstream depth, equal to the head loss through the bridge, is properly
defined as the increase in water level just upstream of the bridge caused by introducing the
bridge piers into the stream. Experimental work by Yarnell1 lead to the following empirical
equation:

Head loss = d3 K Fr32 (K + 5Fr32 - 0.6) (+ 15 4) [m]

Where:
d3 = depth of flow downstream of bridge [m]
K = co-efficient for pier shape
Fr3 = Froude Number downstream of bridge (V3 / (gd3)0.5)
V3 = velocity of flow downstream of bridge [m/s]
 = ratio of pier width to clear span between piers

Note: If the piers are on a raised crest, the values of d3 and V3 relate to the depth and
velocity of flow on the crest, immediately downstream of the piers.

Values for the co-efficient K are given in the following table.

Table 2.1 Co-efficient for Pier Shape


Pier Shape Co-efficient, K
Semi circular nose and tail 0.90
Lens shape nose and tail 0.90
Twin-cylinder piers with connecting diaphragm 0.95
Twin-cylinder piers without connecting diaphragm 1.05
900 triangular nose and tail 1.05
Square nose and tail 1.25

Worked Example

For semi-circular piers, a flow depth of 3.3 m and flow velocity of 0.9 m/s, typical head losses
would be as follows:
K = 0.90
▪ for a bridge with clear spans of 10.0 m and piers widths of 1.0 m ( = 0.10) the head loss
would be 6 mm; and
▪ for a bridge with clear spans of 5.5 and pier widths of 1.0 m ( = 0.18) the head loss would
be 11 mm.

Note: if bridge abutments project into the waterway head losses will be slightly higher.

1 Open Channel Flow, Henderson. Also, Open Channel Hydraulics, Ven Te Chou
PAGE 4 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

2.2 Losses through Trash Racks


For flow through trash racks, the head loss due to the resistance of the rack may be
expressed in terms of the velocity head of the approach flow 2:

h = c V2/2g [ft]
c =  {s / (s+b)}2 sinx 

Where:
V = upstream velocity of flow [ft/s]
g = gravitational constant 32.2 [ft/s2]
c = loss co-efficient
s = thickness of rack bars [ft]
b = clear distance between bars [ft]
 = angle of inclination of the bar from the horizontal
L = length of the rack bar [ft]

Values for the coefficient  and exponent x are given in the following table for various forms of
trash rack bar.

Table 2.2 Values of Co-efficient 


Form of Trash Rack Bar Value of  Value of x
Square nose and tail
- with sharp corners, L/s = 10 7.1 1.0
- with sharp corners, L/s = 12 6.2 1.0
- with slightly round corners, L/s = 8 to 11 6.1 1.0
Semi-circular nose and tail, L/s = 7 5.6 1.5

Worked Example

Determine the head loss due to a round bar trash rack placed at an angle of 600 to the
horizontal in the water way, for an upstream flow velocity of 6ft/s. The bars are ¾” in diameter
placed at 3” centres.
s = ¾”
b = 2¼”
 = 600
 = 5.6
x = 1.5
Loss coefficient c = 0.28
Head loss = 0.16ft (2”)

Note: It is wise to assume that the screen is not kept clear, and becomes partially blocked. If
50% blockage were assumed (ie effectively s = 1.875” and b = 1.125”), then c = 1.76, and the
head loss is 0.98 ft (12”).

2 Open Channel Hydraulics, Ven Te Chow


BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 5

2.3 Losses through Transitions

2.3.1 General
Energy losses through transitions include for losses through pipe culverts and syphons and
other conduits, as well as through warped transitions such as occur between a lined and
unlined section of canal and approaching gated regulators.

The general expressions for transition losses are;

For exit losses: He = Ce (V2 - V3)2


2g

For inlet losses: Hi = Ci (V1 – V2)2


2g
Where:
Hi, He = energy loss through transition [m]
C i, C e = coefficient of head loss in transition
V1 = velocity (in canal) upstream of transition [m/s]
V2 = velocity in constriction [m/s]
V3 = velocity (in canal) downstream of transition [m/s/]

Values for the head loss co-efficients for different transitions are given in the following
sections and apply for flows in the constricted section having Froude numbers of less than
0.5.

2.3.2 Losses between a Trapezoidal Channel and a Conduit


For transitions from a trapezoidal channel to a submerged circular shaped conduit or to a
rectangular shaped structure with a free water surface, the loss coefficients given in Table 2.3
will apply. The transitions are shown on Figure 2.1. Note: these coefficients apply for Froude
numbers of the constricted flow of less than 0.5.

Table 2.3 Energy Loss Coefficients for Transitions from Open Canals to Conduits
Type of Transition Type of Conduit
Circular, Rectangular (free
Submerged water surface)
Ci Ce Ci Ce
Conduit protruding from canal slope 0.65 1.00 0.50 1.00
Conduit connected to straight head wall 0.55 1.10 0.50 1.00
Conduit connected to simplified straight line 0.50 0.65 0.30 0.60
transition
Conduit connected with rounded ends to 0.25 0.50
straight head wall
Conduit connected to simplified straight line 0.20 0.40
transition with vertical walls
Conduit connected to transition with round to 0.40 0.60
rectangular 6D long pipe transition (width
2D, height D)
Warped transition 0.10 0.20
PAGE 6 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

2.3.3 Losses through Horizontal Channel Transitions


Losses though warped transitions such as occur between a lined and unlined section of canal
and approaching and leaving (gated) cross regulators or falls depend largely on the
sharpness or angle of the transition.

a Expanding Section
Ideally, for a diverging (expanding) section the angle of expansion should be between 1:4 (in
which case the exit loss co-efficient is 0.27) and 1:3 (in which case the exit loss coefficient is
0.41). Expansions which are more gradual that 1:4 rarely result in savings in head
commensurate with the extra construction cost.

For sharper transitions, head losses increase as shown on Figure 2.2. Transitions sharper
than 1:2 will result in trash being caught in back eddies and are not recommended.

Given the angle of divergence, the exact form of the side walls is not a matter of great
importance, provided that they follow reasonably smooth curves without sharp corners.

b Contracting Section
For converging (contracting) sections head losses are smaller than for expansions and it is
usual for the approach walls angle of convergence to be much sharper. For a warped
transition approach with a converging angle of 1:2, a head loss coefficient of 0.10 is
applicable. For a semi-circular vertical wall connecting to a trapezoidal channel a head loss
coefficient of 0.25 is applicable.

Worked Example

For cross regulators with semi circular upstream vertical walls, and with downstream walls
that diverge at an angle of 1:3, for upstream and downstream flow velocities of 0.9 m/s, and
flow through the constriction of 1.6 m/s, then the transition head loss would be about 20mm.

In addition, head losses due to bridge piers (if provided) need to be added. Also, for cross
regulators with high crests, losses due to the vertical transitions could also be significant.

For design purposes, the head loss though a cross regulator should be about twice that
calculated to allow a reasonable safety factor.

2.3.4 Losses through Vertical Channel Transitions


As for horizontal transitions the energy loss coefficients due to bridge floors, cross regulator
crests or syphons that project into the channel depend largely on the sharpness of the
transition and are greater for an expending section that for a contracting section. The
coefficients are applicable for sub-critical flow and only accurate when the Froude number of
flow over the vertical transition or Ahump@ is less than 0.5.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 7

a Expanding Section
Energy loss coefficients depend largely on the slope of the transition. Adopted values for Ce
are given in Table 2.4.

Table 2.4 Loss Coefficients (vertical transition, expanding section)


Slope Co-efficient Ce
1:7 0.28
1:6 0.30
1:5 0.40
1:4 0.60
1:3 0.85
1:2.5 1.00
1:2 1.05
1:1 1.10

b Contracting Section
The energy loss across the upstream transition or contracting section is calculated using the
formula given below and not that in Section 2.2.1:

Hi = Ci V22 / 2g

Where:
Hi = energy loss through transition [m]
Ci = coefficient of head loss
V2 = velocity above the (raised) floor [m/s]

The coefficient of head loss (Ci) depends on the ratio water depth above floor (d2) / water
depth in the canal upstream (d1). For an abrupt, vertical transition the following values of Ci
are adopted.

Table 2.5 Loss Coefficients (vertical transition, contracting section)


d2/d1 Ci
1.0 0.00
0.9 0.01
0.8 0.05
0.7 0.10
0.6 0.20
PAGE 8 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

For a sloping vertical transition where the slope is 1V : 1.7H, or for rounded transitions, the
values for Ci given above should be multiplied by 0.6.

Worked Example

For a design flow depth of 4.67 m (15.32 ft) and a design flow velocity of 1.21 m/s (3.99 ft)
application of the above gives energy losses shown in the table below. For design purposes it
is suggested that the calculated head losses be increased by a safety factor (SF) of 1.5.

Height of Crest (hump) Head loss Design Head loss


above Channel Bed (with SF of 1.5)
(m) (m) (m)
0.25 0.001 0.001
0.50 0.003 0.004
1.00 0.011 0.017
1.50 0.032 0.047
2.00 0.083 0.125
2.50 0.193 0.290

For humps that project so high that the Froude number of the flow over the hump is greater
than 0.5 the equations given herein for transitions are not valid.

Note: using formulae for a Crump weir gives a head loss of 0.156 m for a 2.5 m high hump,
and 0.300 m for a 3.0 m high hump.

2.4 Losses in Bends of Closed Conduits


For sharp edged bends or elbows, the loss will be calculated using:

H = C V2 / 2g

Where:
H = loss in bend [m]
V = velocity in conduit [m/s]
C = Loss coefficient

Values for the loss coefficient (C) for both rectangular and circular conduits are given in Table
2.6.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 9

Table 2.6 Loss Coefficients for Sharp Edged Bends


Angle of elbow (0) C for Rectangular C for Circular
Profile Profile
5 0.02 0.02
10 0.04 0.03
15 0.05 0.04
22.5 0.06 0.05
30 0.14 0.11
45 0.30 0.24
60 0.60 0.47
75 1.00 0.80
90 1.40 1.10

For the losses in rounded bends the energy loss will be lower depending on the radius of the
bend.

2.5 Energy Losses through Friction


Energy losses through friction are may be calculated using Manning=s formula:

Hf = L n2 V2

R4/3
Where:
Hf = loss due friction [m]
L = length of conduit [m]
V = velocity in conduit [m/s]
n = Manning=s coefficient
R = hydraulic radius of conduit , A/P [m]

For Manning=s An@, the following values apply 3.

Table 2.7 Manning’s “n”


Material Value of Manning’s “n”
Minimum Normal Maximum
Masonry (dressed Ashlar) 0.013 0.015 0.017
Brick in cement mortar 0.012 0.015 0.018
Old Concrete (depends on condition) 0.014 - 0.020
New Finished Concrete (depends on 0.011 0.013 0.015
quality of finish & formwork used)
Untreated Sprayed Concrete 0.016 0.019 0.023
Metal (smooth, unpainted) 0.011 0.012 0.014

3 Open Channel Hydraulics, Van Te Chow


PAGE 10 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 2.1 Headloss Co-efficients for Transitions


BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 11

Figure 2.2 Energy Loss in Channel Expansions


BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 13

3 REGULATING STRUCTURES

3.1 Introduction
Irrigation systems require flow regulating structures to allow timely and equitable supply of
irrigation water to farmers. Providing the right structure type of the right size at all points on
the scheme requires careful consideration and extensive discussions with farmers. This
section discusses the following flow regulating structures:
• Checks and Check Drops
• Turnouts
• Division Structures

3.2 Discharge Characteristics


Flows through the openings of a distribution structure are either determined from the formula
for flow over a rectangular weir, or the formula for flow through an orifice. Where gates are
provided, providing they can be lifted clear of the water level upstream of the structure, then
the weir formula is applicable.

These formula are described in Part 3: Weirs of the Design Manual, Chapter 5. However, for
convenience the formulae are summarised below.

a Flow over Rectangular Weir


For free flow conditions, the discharge is given by:

Q = 1.7 b h11.5 [in metric units]

Where:
Q = discharge [m3/s]
h1 = upstream depth of flow over the weir crest [m]
b = width of the weir [m]

For submerged (non-modular) flow, the discharge is reduced by multiplication by the


applicable reduction factor “f”.

b Flow through Orifices


Free flow through an orifice occurs as long as the roller of the hydraulic jump does not
submerge the section of minimum depth of jet which is located at a distance a/ downstream
of the face of the vertical gate. Free flow discharge is a function of the upstream water depth
and the gate opening and is given by the following formula:

Q = Cd Cv A 2g(h1 –  a) [in metric units]

Where:
Q = discharge [m3/s]
Cd = discharge co-efficient
Cv = velocity co-efficient
A = area of opening (= product of orifice opening “a” and orifice width “b”) [m2]
h1 = upstream water depth above orifice crest [m]
 = contraction co-efficient.

For most cases where the approach velocity is negligible, the co-efficient of velocity Cv can be
taken as 1.0. Values of Cd and  depend on h1/a, and for design can be taken from the
following Table 3.1.
PAGE 14 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Table 3.1 Values for Cd and  for Free Orifice Flow


H1/a  Cd
1.5 - 2.0 0.63 0.60
2.5 - 3.5 0.625 0.60
3.5 - 5.0 0.625 0.605
 5.0 0.62 0.61

Note: If the floor downstream of the (gate) orifice is depressed the free flow formula may
not be applicable. The free flow formula may be retained provided the floor level is
maintained at gate sill level (crest level) for a distance of at least 2a before dropping
to a lower level.

3.3 Checks

3.3.1 Purpose
Checks are used to regulate the canal water surface upstream of the structure and to control
the downstream flow. When a canal is flowing at partial capacity, checks may be operated to
backup the flow, and maintain sufficient water levels to command upstream offtakes. Checks
typically have a walkway spanning the structure to allow access to gates or stoplogs. Checks
may be separate structures, or they may be combined with other structures such as turnouts.
The spacing of check structures is determined by the need to command turnouts. Steeper
canal slopes usually require closer spacing between checks.

Under BCIAP, checks were provided along the main and branch channels of flood irrigation
schemes (eg Toiwar FIS), and designs were based on the guidelines given in USBR “Design
of Small Canal Structures”. Design considerations are summarised here.

3.3.2 Design Considerations


Checks need to withstand hydraulic forces imposed by water ponded upstream of the check,
and no water downstream. Under this loading condition, consideration of the following stability
aspects is required:
• Percolation (seepage) which could cause “piping”
• Scour
• Uplift
• Sliding

Provision of upstream and downstream cutoffs, and a sufficiently thick floor, guards against
stability failure.

The check length needs to be sufficient long to dissipate turbulence.

The flow velocity though check structures using stop logs should be limited to about 1.1m/s
(3.5ft/s). With gates the velocity may be increased to about 1.5m/s (5ft/s). The use of stop
logs is not recommended for flows greater than 1.4 m3/s (50 cusecs), where the width of the
stop log opening is more than 1.5m (5ft) or if the water depth is more than 1.8m (6ft).

Operating (walkway) decks should be at least 1m (3ft) wide.

In metric units, the headloss [m] through a check may be computed as 1.5 (v1-v2)2 / 2g where
v1 and v2 are the velocities of flow upstream of the check and though the check respectively
[m/s], while g is 9.806 m/s2.

Standard Check Structures taken from USBR “Design of Small Canal Structures” are given on
Figure 3.1 and Figure 3.2.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 15

3.4 Check Drops


These combine the function of a vertical drop with those of a check. The drop portion of the
structure dissipates excess head in the canal, while the check portion maintains the required
upstream water level.

The vertical drop through this type of structure in an unlined canal is generally limited to 1m (3
ft) for canal capacities of 2 m 3/s (70 cusecs) and less, and to 0.45m (1.5ft) for larger
capacities. For lined sections these structures have been used for vertical drops of up to
about 2.4m (8ft). For larger drops, economic and stability considerations are likely to mean
that chutes are more suitable.

Check drops are designed to meet the same requirements for percolation, stability, velocity
and safety as checks. In addition the structure must be long enough to contain the falling
water and to provide a stilling pool to control turbulence downstream.

Standard Check-Drop Structures taken from USBR “Design of Small Canal Structures” are
shown on Figure 3.3 and Figure 3.4.

3.5 Turnouts

3.5.1 Purpose
Turnouts are used to divert water from the supply channel to a smaller channel. The structure
usually comprises an inlet, a conduit to convey water through the bank of the supply channel
and an outlet transition. Gates are usually used in the inlet to control the flow.

Under BCIAP, turnouts were provided along the main and branch channels of flood irrigation
schemes (eg Toiwar FIS), and designs were based on the guidelines given USBR “Design of
Small Canal Structures”. Design considerations are summarised here.

3.5.2 Design Considerations


For structures not having a concrete outlet transition and discharging into an unlined channel,
the maximum velocity in the (pipe) conduit should be about 1.1m/s (3.5 ft/s). If an outlet
transition is provided, the maximum velocity in the (pipe) conduit may be increased to about
1.5m/s (5 ft/s). A minimum pipe diameter of about 0.45m (18 inches) is used.

The structure must provide a percolation path of sufficient length to prevent the removal of
foundation materials by seepage water (refer Lane’s Weighted Creep method in Part 3:
Weirs). It may be necessary to provide concrete collars to the (pipe) conduit within the bank to
increase the length of the percolation path.

Where there is excessive head across the structure, the conduit and outlet portions of the
turnout must be designed as a drop structure.

Turnout inlets should be set so that they do not protrude into the canal prism.

Parshall flume and weir structures are sometimes incorporated into the turnout designs to
measure flows. An alternative are constant head orifice structures developed by USBR to
both regulate and measure the flow.

A typical turnout taken from USBR “Design of Small Canal Structures” is shown on Figure
3.5.
PAGE 16 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

3.6 Division Structures

3.6.1 Purpose
Division structures are used to divide the flow from a supply channel or pipe among two or
more channels or pipes. If measurement is not required at the point of division, the flow
through the structure may be directed through the various outlets with gates or stop logs (time
division structures). If flow measurement is required, and the required head is available, weirs
may be used to proportion the flow (flow division structures).

Many time division structures are designed to be operated as flow division structures as well,
since while for much of the year farmers may use time division at the structure, at times of
high flow in the channel, they may divide the flow in order to make the flows more
manageable.

Under BCIAP a range of division structures have been used from simple field outlets
(nuccas), to gated time and flow division structures. For the perennial irrigation schemes
standard drawings have been prepared for flows up to 0.14m3/s (5 cusecs).

For flood irrigation schemes, larger flow division structures were adopted, with weir crest
levels and widths carefully selected to proportion the flow in accordance with traditional and
agreed water rights.

3.6.2 Design Considerations


For small division structures (flows less than 0.14m3/s, 5 cusecs), stability factors need not be
considered. For larger structures, it is necessary to check for percolation (seepage) under the
structure, scour, and uplift. Dissipation of turbulence also needs to be considered.

3.6.3 Time Division Structures

a Field Outlets (Nuccas)


In the early days of BMIADP the
Project used the standard
nuccah designed on the Mona
Reclamation Project in the
1950's and used extensively in
Sindh and Punjab. This structure
is usually constructed of
plastered brickwork and uses
circular reinforced concrete
panels as gates.

These structures have a number


of problems: bricks are not widely
available in Balochistan; the
structures are easily dismantled
and bricks stolen; the concrete Nucca Stucture
panels chip easily and once
chipped leak; and the circular shape of the gate opening requires quite a high structure to
incorporate it.

Not withstanding these deficiencies, the standard nucca is cheap, and under BCIAP it was
reintroduced for on-farm channels improved under the Command Area Development
component of the project. They are also suitable for main channels of small schemes where
the perennial flow is less than about 28 l/s (1 cusec).
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 17

b Time Division Structure for Flows less than 0.042 m3/s (1.5 cusecs)
For the main channels where the flow is more than about 28 l/s (1 cusec), an alternative
structure to the standard nuccah was developed and is shown on Figure 3.6. This structure
comprises a simple mass concrete box, with gated openings set in the sides to feed flow in
any one of three directions. On many minor perennial irrigation schemes identical structures
are required in large numbers, both as field outlets and division structures between main and
secondary channels. It is then worthwhile for a steel mould to be made up to form the four
sides of the structure and to cast in the gate frames. The structure is then constructed by first
laying the concrete base slab and then, with the base slab concrete still green, placing the
steel mould on the base slab and casting the four walls with the gate frames built in.

The gates used in these structures are simple steel sliding shutters, as described in
Section 3.7. Whilst these do leak, the leaks can be sealed with mud. They are also less
vulnerable to damage than the circular concrete nuccahs.

The size of the gate openings is calculated


using the weir equation given in Section 3.2.
The maximum gate width is 457mm (18
inches). The height of the structure
incorporates a 76mm (3 inch) freeboard above
the incoming channel. The plan dimensions
are determined to suit the gate openings.

The drawing on Figure 3.6 is suitable where


the drop to the main channel and the fields is
between 150mm, 6 inches (the minimum drop
recommended) and 230mm, 9 inches. Where
the drop to either the channel or the fields is
larger than this then the structure should be
used in conjunction with drop boxes, as shown
on the drawing on Figure 3.7. The sizes of the
drop boxes are determined as for drop
structures (see Chapter 4).

In the event that the drop is less than 150mm


(6 inches), and flow not modular (ie not free),
and where flows themselves do not have to be
divided in specific proportions, it is sufficient to Time Division Structure
ensure that the gates for each outlet should no
narrower than the bed width in the main channel.

c Time Division Structure for Flows less than 0.14 m3/s (5 cusecs)
A typical time division structure for flows more than 0.042 m3/s (1.5 cusecs) and less that 0.14
m3/s (5 cusecs) is shown on Figure 3.8.

In order to provide a firm foundation for the gate frames, the section of the structure which
incorporates the gates is made from mass concrete. The rest of the structure can be more
easily and cheaply made from stone or concrete block masonry, depending on which has
been adopted for the scheme. The structure shown incorporates both left and right hand
offtakes though it can obviously be adapted to have only two outlets in any of the three
directions shown, by blanking off one wall.
PAGE 18 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

The design procedure for the structure is as follows:


• Gain a good understanding of the flow division requirements for the structure and
ascertain the design flow for the structure.
• Using the weir equation to size the gates (widths and heights).
• The sizes of the upstream section of the structure will be determined as follows:
Width: The larger of either the top width of the channel plus 0.6m (2 ft), or the sum of
the width of the gates set across the structure, plus 0.3m (1 ft) between the gates and
0.3m (1 ft) between the edge of the gates and the side walls.
Length: To suit the gates with 0.3m (1 ft) between gates, 0.45m (1.5 ft) between the
last gate and the downstream end of the structure and a length equal to the width of
the structure between the upstream gate and the headwall of the structure.
Height: The height of the structure will normally be the height of the incoming canal,
plus 150mm (6 inches) freeboard and a 0.3m (1 ft) deep silt trap.
• Wherever possible the outlet channels should be at least 150mm (6 inches) below the
incoming channel to ensure that the flow through the gates is not submerged.
• The dimensions of the outlet boxes will be designed to suit the dimensions of the
gates and the outflowing canal, except in the case where the drop across the
structure is greater than 0.3m (1 ft), in which case they will be in accordance with the
dimensions for drop structures.

3.6.4 Flow Division Structures

a Flow Division Structures for Flows less than 0.14m3/s (5 cusecs)


The time division structure shown on Figure 3.8, without the gates, would form a suitable flow
division structure or, with the gates, a combined flow and time division structure. The
structure will only work adequately as a flow division structure if the following three conditions
are met:
• The size of the openings are adjusted to be exactly proportional to the water rights
allocated to each of the downstream channels.
• Flow over the weir on every opening must be free flowing (i.e. flow must not become
submerged by the downstream water level backing up. For this reason it is
recommended that the drop across the weir for these small structures should be at
least 150mm (6 inches).
• The incoming velocity immediately upstream of the outlets should be less than 1 m/s
(3 ft/s).

The water depths over the weirs of any division structure are effectively governed by the
height of the incoming channel sides (banks). Hence, to determine the width and number of
gates required, one first calculates the maximum allowable head of water over the offtake cills
(h) and then determines the required width and number of gates which will pass the design
flow without causing the water upstream to back up.

An alternative flow division structure is shown on Figure 3.9. This structure closely follows the
traditional log flow division structures and may therefore be more acceptable to some groups
of farmers. However, where multiple outlets are required, the structure quickly becomes
excessively wide and expensive.

Trials have been carried out in order to determine sensible dimensions for flow division
structures of the type shown on Figure 3.9. The trials were carried out under BMIADP on
schemes at Duki Viala, Thal and Tor Morgha. As a result of these trials, the following
guidelines are proposed:
• The maximum velocity of approach to such structures should be 1m/s (3 ft/s).
• The inlet walls should be splayed at 45o in order to reduce the area of dead water.
• The level of the weirs should be set at 150mm (6 inches) above the incoming channel
bed level. There should also be a minimum of 150mm (six inches) drop across the
structure in order to achieve modular flow across the weirs.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 19

• It is recommended that a silt trap of at least 150mm (6 inches) be provided.


• The piers between successive ungated weir outlets should be 150 - 230 mm (6-9
inches) wide.
• An obstruction/baffle less than one third the width of the structure should be placed
0.3 m (1 ft) downstream of the end of the incoming channel to break up the incoming
jet. If no baffle is provided, a jet of flow tends to pass down the centre of the structure,
with slack water (and in some cases eddies of reversing flow) occurring along the
sides. The aim of the baffle is to break up this jet. However, it is important that the
baffle is no larger than specified otherwise this tends to produce two jets of water
along either side of the basin, with slack water in the middle.

b Flow Division Structures for Large Flows


The design of Flow Division structures for larger flows follows the same principles for the
smaller structures to determine weir levels and widths. However it is necessary to check for
percolation (seepage) under the structure, for scour, and for uplift. Sliding may also have to
be checked. Dissipation of turbulence also needs to be considered.

3.7 Gates for Small Division Structures


For the small division structures for flows less than 0.14 m 3/s (5 cusecs), simple steel gates,
manufactured from 3mm (1/8th inch) thick steel plate, as shown on Figure 3.10 are cheap,
easy to fabricate, robust and the manufactured tolerances are such that, for the low operating
heads involved, they can be sealed by forcing a small amount of mud pushed into the gate
grooves.

The gates should be hot dip galvanised after fabrication. The maximum practical size is 450
mm (18 inches) wide by 450 mm (18 inches) deep. Where a larger open area is required, it is
suggested that multiple gates are used. It is proposed that wherever possible one of the four
standard sizes shown in Table 3.2 is adopted.

Table 3.2 Standard Gate Sizes


Width (mm & inches) Height (mm & inches)
450 (18) 450 (18)
450 (18) 300 (12)
450 (18) 230 (9)
300 (12) 300 (12)
300 (12) 230 (9)
300 (12) 150 (6)
230 (9) 230 (9)
230 (9) 150 (6)

On distribution structures not every opening needs to be provided with a gate. Indeed it would
be dangerous to do so since someone might decide to put in all the gates, causing the
structure to overtop. If one or two gates are always to be left open, then gates for these need
not be provided.

Unless fixed to the structure the gates would be easily stolen. The gates are therefore fixed to
the structure by means of a steel chain, one end of which is welded to the gate handle and
the other is fixed to a steel ring embedded into the concrete of the structure. The chain must
be long enough to enable the gate to be used in more than one of the gate frames.
PAGE 20 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

3.8 Canal Protection at Regulators


For unlined canals, stone protection is required both upstream and downstream of regulators.
This may be calculated as detailed in Chapter 8 of Part 3: Weirs. As a guide, typical
dimensions of stone pitching required at regulators (and at other canal structures) are given in
Table 3.3.

Table 3.3 Canal Stone Pitching Protection at Regulators


Approx Canal Downstream Protection Upstream Thickness
Capacity Protection of
Pitching
(m)
(m3/s) (cusecs) Turndown Bed Sides Bed &
Depth (1) Sides
(m) (m) (m) (m)
0.10 3.5
0.16 5.6 0.20m on
0.25 8.8 0.40 1.0 1.0 0.15m
0.39 13.4 1.0 gravel
0.60 21.2 backing
0.90 31.8
0.50 2.0 4.0
1.50 53
2.90 102
4.20 148
3.0 6.0 2.0
7.00 247 0.30m on
10.00 353 0.15m
10.60 374 0.60 gravel
4.0 8.0 3.0
20.00 706 backing
33.00 1,165
50.00 1,765 6.0 12.0 4.0
80.00 2,824
Note: A 1V:2H sloping “turndown” to the depth indicated is recommended at the end of the bed and side
protection downstream of the structure, in addition to the lengths indicated above .
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 21

Figure 3.1 Standard Check Structures (Q less than 70 cusecs)


PAGE 22 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 3.2 Standard Check Structures (Q greater than 70 cusecs and less than 105 cusecs)
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 23

Figure 3.3 Standard Check Drop Structures (Q less than 70 cusecs)


PAGE 24 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 3.4 Standard Check Drop Structures (Q greater than 60 and less than 100 cusecs)
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 25

Figure 3.5 Turnout


PAGE 26 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 3.6 Standard Outlet Suitable for Discharge less than 1.5cusecs (Drop: 6” to 9”)
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 27

Figure 3.7 Standard Outlet Suitable for Discharge less than 1.5cusecs (Drop: 9” to 6ft)
PAGE 28 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 3.8 Time Division Structure Suitable for Discharge less than 5.0cusecs
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 29

Figure 3.9 Typical Flow Division Structure


PAGE 30 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 3.10 Gate Details


BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 31

4 DROPS AND CHUTES

4.1 Introduction
Drop structures are required on both lined and unlined channels to dissipate excess head.
Where channels are constructed on steeply sloping ground the velocity must be not so high to
cause scour in an earthen channel, or disruptive flow in a lined channel. Where the channel
slope is less than the ground slope, drop structures will be required to maintain the channel
at, or near, ground level.

There are three main types of drop structures, as shown on Figure 4.1. The choice of drop
type is generally dictated by economic considerations, and the following provides some
guidance.
• Vertical Drop. This is the simplest form of drop structure. For an unlined canal the
vertical drop is suitable for drops of up to about 1 m (3 ft) for canal capacities of up to
about 9.9 m3/s (350 cusecs) 4, and for drops of up to about 0.46 m (1.5ft) for larger
capacities. For lined sections, vertical drops are suitable for drops of up to about 2.4
m (8ft). About 50% of the excess energy is dissipated as the water strikes the basin
floor and the remaining head is lost in a hydraulic jump, which is usually submerged,
and in residual turbulence downstream.
• Glacis Type Drop. Energy dissipation for the glacis drop is by means of a hydraulic
jump which forms at the bottom of the glacis. Glacis drops are common for larger
canals and for drops in the range from about 1 m (3ft) to 4.6 m (15ft).
• Pipe Drop. Here, the pipes are inclined to start with and then either horizontal or
sloping slightly upwards. The energy is dissipated by means of a hydraulic jump
formed in the pipes. However, a downstream outlet structure is often also required to
dissipate residual energy. Pipe drops are generally used for smaller flows than for
glacis type drops, but again for drops in the range from about 1m (3ft) to 4.6 m (15ft).

Figure 4.1 Drop Structure Types

Vertical Drop Glacis Drop

Well Drop Pipe Drop

Most drops on BCIAP irrigation schemes are small, being between 0.3 m (1ft) and 2.4m (8 ft)

4The limit of 350 cusecs (15 m 3/s is recommended for a Sarda type vertical drop. For the USBR type
vertical drop, with a shorter stilling basin, a limit of about 70 cusecs (2 m 3/s) is recommended.
PAGE 32 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

for flows up to about 0.14 m3/s, 5 cusecs (perennial irrigation schemes), and between 0.3m
(1ft) and 1 m (3 ft) for flows of up to about 14 m3/s, 500 cusecs (flood irrigation schemes). For
these conditions, the vertical drop is likely to be the most economical.

Drop structures are often combined with other structures, particularly outlets and checks.

The design methodology given below is suitable for the design of drop structures standing
alone, or for the stilling basins of other structures which incorporate a drop.

Rectangular section chutes are usually used where the drop is greater than about 4.6m
(15ft), and where the water is conveyed over long distances and along grades that may be
flatter than those for drops but yet steep enough to maintain supercritical velocities. The
decision as to whether to use a chute or a series of drops is based on a hydraulic and
economic study of the two alternatives. From a hydraulic viewpoint, drops should not be so
closely spaced to prevent uniform flow between consecutive structures. Very broadly, the
minimum distance between drops should be about 60m (200ft).

4.2 Construction Materials


Under BCIAP, for chutes and for drops the base is generally constructed of concrete and the
walls of stone masonry. For larger drops, and for stilling pools at the end of the chutes, the
concrete base may be protected with rock armouring. One alternative, used for the flood
irrigation schemes constructed under BCIAP where a large number of almost identical drops
were required (eg Chandia and Toiwar FIS), was for the whole structure to be constructed of
reinforced concrete.

For drops in lined channels, expansion joints are required between the structure and the
channel lining upstream and downstream.

4.3 Vertical Drops

4.3.1 Hydraulic Design

a Choice of Drop Size

Where a drop structure is required to take a


channel over a drop in the ground surface, the
location of the drop is obvious. However,
where a series of drops are required to
maintain a channel near ground level on
steeply sloping ground, it is recommended
that, subject to the limitations of requirements
to command adjacent land, the drops are
located so that the bed of the channel is
always in cut. This is in order to avoid
settlement of canals built on poorly compacted
fill.

For an unlined canal vertical drops of up to


about 1m (3ft) are recommended for canal
capacities of up to 10m3/s, 350cusecs (for a
Sarda type vertical drop: 2m3/s, 70cusecs for a
USBR type vertical drop), and up to about
0.5m (1.5ft) for larger capacities. For lined
sections vertical drops may be up to about
2.4m (8ft) for steeply sloping ground. For Vertical Rectangular Drop
gradual slopes, large drops will result in
excessive cut, and despite the saving in structure costs, will result in an expensive
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 33

conveyance system. The optimum drop for gradually sloping land is likely to be around 0.6m
to 1m (2 to 3 ft).

For a lined channel, significant savings can be made in the cost of an irrigation conveyance
system if larger drops are replaced with a series of smaller steps (< 9"). The savings come
about from avoiding the cost of constructing a complete stilling basin and reducing the
amount of cut and fill between bigger drops. The limitation on the use of this type of drop is
that they should ideally be not less than (say) 60m, 200ft apart, and are therefore only an
option for fairly mildly sloping ground.

b Hydraulic Design Methods


Six standard methods of design for the stilling basin length downstream of the drop are
compared. All are essentially empirical, although some incorporate elements of basic
hydraulic theory. The six methods are briefly described below:
• Newton's Laws of Motion. These describe the arc created by a falling jet of water.
The method can therefore be used to calculate the distance along the stilling basin
that the water will fall, but not the overall length of the basin required to dissipate
sufficient energy.
• USBR Method. USBR 5 give dimensions for two standard vertical drops; for flows
less than 2m3/s (70 cusecs), and for flows greater than 1.7m3/s (60 cusecs) and less
than 2.8m3/s (100 cusecs). These are discussed in Section 3.2: Check Drops.
• SOGETHA. This method, developed by the French, is quoted by FAO 6. The length
of basin is related to the drop height only, and appears quite short. It would be
satisfactory for low approach velocities only.
• Varshney, Gupta & Gupta. An empirically based design quoted by Vashney 4. It
relates the stilling basin length to the depth over the weir and the height of the drop.
• Sarda Type Fall. An empirically based Indian design, quoted by Vashney 7. It relates
the stilling basin length to the critical depth upstream and the height of the drop.
• ILRI Method. This method is quoted by Bos 8. The distance from the fall at which the
water strikes the stilling basin floor is calculated from a formula which relates flow and
drop height. The length of the stilling basin then has to be determined from
considerations on the formation of an adequate hydraulic jump.
• VDF Fall. Another Indian method quoted by Vashney 6. It relates the length of the
basin to the height of the drop and the upstream depth of water.

5/ USBR ADesign of Small Canal Structures @ United States Department of Interior, Bureau of Reclamation.

6/ Kraatz, DB & Mahajan, IK ASmall Hydraulic Structures.@ Irrigation and Drainage Paper No 26, FAO. 1975.

7/ Vashney, RS, Gupta, SC & Gupta, RL ATheory and Design of Irrigation Structures.@ Nem Chand & Bros,
Roorkee, UP, India, 1988

8/ Bos, MG ADischarge Measurement Structure.@ International Institute for Land Reclamation and Improvement
(ILRI), Publication No 20. Wageningen, The Netherlands.
PAGE 34 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

c Design of Vertical Drop using Method Quoted by Vashney 4.


For a vertical drop the following dimensions for the stilling pool are given in Section 8.4 of
“Theory and Design of Irrigation Structures” by Varshney, Gupta and Gupta:

Lsb = 5.0 (H x h1) 0.5 [metric units]


X = 0.25 [H x h1) 2/3 [metric units]

Where:
Lsb = stilling basin length [m]
X = stilling basin depth [m]
H = head loss over drop [m]
h1 = upstream depth of flow over crest [m]

The design parameters are shown on Figure 4.2.

Figure 4.2 Symbols to Describe Flow Over a Vertical Drop

U/S FSL

D/S FSL
H
h1

Stilling Basin X

Lsb
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 35

d Design of Sarda Type Vertical Drop


Length of Crest
Since fluming causes unfavourable conditions, the length of the crest (bc) is kept equal to the
bed width. For unlined channels, to allow for possible expansion in the future, the crest length
may equal the bed width + flow depth.

Stilling Basin Dimensions


The stilling basin length and depth are given by the following equations:

Lsb = 3.8 dc + 0.415 + H [metric units]

X = dc / 3 [metric units]

dc = (q2/g) 1/3 [metric units]

Where:
Lsb = the length of the stilling basin ([m]
dc = the critical depth of flow [m]
H = the height of the drop [m]
X = the height of the end sill [m]
q = Q / bc (ie flow per unit width) [m2/s]
g = gravitational constant (9.806) [m/s2]

e Standard USBR Vertical Drop Designs


Drop designs for two standard vertical drops, for flows less than 2m 3/s (70 cusecs), and for
flows greater than 1.7m3/s (60 cusecs) and less than 2.8m 3/s (100 cusecs), in accordance
with USBR recommendations are given in Section 3.2: Check Drops.

f Standard Small Vertical Drop Designs Adopted under BCIAP


Standard details for drop structures used on the BCIAP perennial irrigation schemes, which
are sized in accordance with the recommendations for a Sarda type drop, are shown on
Figure 4.3. This standard design is suitable for flows of 0.03 to 0.57m3/s (1 to 20 cusecs) and
for drops of 0.3 to 2.0m (1 to 6 ft).

For larger structures, a different


design is required. The Sarda
type drop with a rectangular crest
is recommended for use for
discharges up to 15 m3/s (530
cusecs). The design is suitable
for both lined and unlined
channels, subject to adequate
protection being provided to
unlined channels upstream and
downstream of the drop
structure, and stability aspects
being taken into consideration.
Vertical Drops at Toiwar FIS
PAGE 36 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Worked Example

Design a vertical drop for an unlined channel with a design discharge of 530 cusecs (15.0
cumecs), a bed width of 35 feet (10.67m) and a flow depth of 3.27ft (1.0m), design.

Using the equations for a Sarda type vertical drop (Section 4.3.1d), and selecting a
(maximum recommended) drop of 3 ft (0.91m), stilling basin dimensions are as follows:

Crest length = 35ft (10.7m)


q = 15/10.7 = 1.40 m2/s
dc = 2
(1.40 / g )1/3 = 0.59
Length of basin = 3.8 dc + 0.415 + 0.91 = 3.6 m
Depth of basin = dc/3 = 0.20m

Using the equations in Section 4.3.1 c


Length of basin = 5.0 (0.91 x 1.0) 0.5 = 4.8 m
Depth of basin = 0.25 (0.91 x 1.0) 2/3 = 0.23m

Adopt
Length of basin = 5.0 m (16.4ft)
Depth of basin = 0.25m (0.8ft)

4.4 Chutes

4.4.1 Description
A chute structure may consist of an inlet, a chute section, an energy dissipator and an outlet
transition. The chute section may be a pipe or an open channel section. Chutes are suited to
conveying water over longer distances, over flatter slopes, and through greater changes in
grade than drops. The inlet portion of the structure transitions the flow from the channel
upstream to the chute section. If the upstream channel is unlined, the inlet should provide a
control to prevent racing and scouring of the upstream channel. This is achieved by
combining a check, a weir or a control notch with the inlet. The inlet should have cutoffs as
required to provide a sufficient length of percolation path as computed by Lane’s weighted
creep method. The chute section generally follows reasonably closely to the original ground
surface, avoiding deep cut or fill, and connects to the energy dissipator at the downstream
end. Stilling pools or baffled outlets are used as energy dissipators. As outlet transition is
used when it is necessary to transition the flow between the energy dissipator and the
downstream channel. If it is necessary to provide tailwater for the energy dissipator, the water
surface at the outlet must be controlled. The transition may be used to provide backwater by
raising the floor of the transition. Alternatively, a control may be built into the outlet transition.

4.4.2 Design Considerations

a Transitions
Transitions in an open channel chute must be designed to prevent formation of waves which
can be troublesome as they travel through the chute section and energy dissipator. As a
general consideration the maximum deflection angle in the water surface should be about 300
(inlet section) and 250 (outlet section).

The transitions should be symmetrical about the centerline of the chute.


BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 37

b Chute Section
The usual section for an open channel chute is rectangular, and this was adopted for all
chutes built under BCIAP.

For chutes less than about 9m (30ft) long, friction in the chute may be neglected, and
Bernoulli’s (energy balance) equation is used to compute the flow variables at the bottom of
the chute. The energy balance equation, given below, is solved by trial and error.

2 2
V1 V
y1 + Z + = y2 + 2
2g 2g

Where:
y1 = depth of flow upstream [m]
Z = elevation above datum [m]
V 1 (2) = flow velocity upstream of chute (at bottom of chute) [m/s]
g = gravitational constant, 9.806 [m/s2]

For chutes longer than about 9m (30ft), friction losses are included, and the equation
becomes:

2 2
V1 V
y1 + Z + = y 2 + 2 + hf
2g 2g

Where hf is the headloss due to friction in the reach (ie the friction slope times the length of
the reach). The friction slope may be calculated using Manning’s equation with the
appropriate value for Manning’s “n” (see Table 2.4 in Chapter 2).

c Stilling Basin
The stilling basin ensures that energy of flow is dissipated in a hydraulic jump. Under BCIAP
horizontal USBR type stilling basins were provided. These structures are the same, but on a
much smaller scale, as the stilling basins provided downstream of weirs, and their design is
covered in Part 3: Weirs of this Design Manual.

4.5 Stability and Structural Considerations


Drop and chute structures are designed to:
• Resist sliding and overturning
• Prevent percolation water removing foundation material
• Prevent undermining of the structure
• Provide foundation pressures less than the maximum allowable bearing pressures.

Unless the drops are gated, or have provision for stop logs, of these stability considerations,
only percolation and scour are likely to be critical. For small lined channels, the channel lining
protects against both scour and percolation failure. For larger lined channels, it would be
prudent to assume that the lining fails and provide cuts offs.

For drops in unlined channels, cutoffs are required to control percolation and undermining
(scour). These cutoffs need to extend well into the berms and embankments either side of the
structure, as well as underneath the structure. In addition, stone or gabion protection is
required both upstream and downstream.

For chutes, in addition to percolation and scour, sliding may be a problem. To increase the
PAGE 38 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

resistance of the chute section to sliding, cutoffs are used to “key” the structure into the
foundation.

Concerning structural aspects, structure thickness and reinforcement steel patterns (for
reinforced concrete constructions) need to resist bending moment, thrust and shear stresses
imposed by loads on the structure. These loads include uplift pressures and lateral earth and
water pressures.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 39

Figure 4.3 Standard Drop Structure used for BCIAP PISs


PAGE 40 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 41

5 AQUEDUCTS

5.1 Introduction
Aqueducts are required to convey water across river beds (nullahs), or rough ground where a
channel would be vulnerable to landslips. Aqueducts can have an overall length of just a few
feet, or a thousand feet or more.

In most instances where flow


has to be conveyed across a
river, a choice has to be made
whether to use an inverted
syphon beneath the river or an
aqueduct over it. The aqueduct
will almost certainly be the more
expensive option. However,
aqueducts are not vulnerable to
blockages which are a severe
problem with inverted syphons
(see Chapter 6). An aqueduct is
therefore usually the preferred
technical solution.

This section describes the steps Aqueduct, Sinjori PIS


involved in locating and
designing an aqueduct and makes recommendations on best practice based on past
experience on the BMIADP and BCIAP schemes.

5.2 Construction Materials


Aqueducts comprise four main components: the trough, columns, abutments and
foundations. In Balochistan, aqueduct troughs are usually made from in-situ reinforced
concrete. Alternative construction methods and materials include fabricated steel troughs,
steel pipes, precast concrete sections, fibre glass troughs, etc. Of these, fabricated steel
pipes are a viable alternative to concrete and are particularly suited to high aqueducts (10m,
33ft above bed level) where casting troughs in-situ becomes expensive, and precast concrete
troughs are too heavy to handle easily. Fibre glass and plastic alternatives are not considered
sufficiently tough to be durable.

Up to the early 1950's, stone masonry was the preferred material for columns and abutments
for road, rail and water bridges throughout Pakistan. Since 1950, reinforced concrete has
gained favour for columns, due to the reduction in size of the columns that is possible.
However, scour damage to concrete columns can be a problem in fast flowing rivers,
requiring expensive steel sheeting protection.

Aqueduct abutments generally have to be massive to prevent erosion of banks in the vicinity
of the structure. Masonry, being cheaper than concrete, and the preferred material of local
contractors in Balochistan, is therefore generally used for abutments.

The foundations to columns and abutments are always of in-situe concrete, reinforced in the
case of RCC columns and plain for masonry columns and abutments.
PAGE 42 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

5.3 Aqueduct Configuration

a Location
The aqueduct must be located in a stable section of the river, where it can command the
irrigation scheme and be commanded by the irrigation works upstream. A stable section of
river is one that is not meandering, liable to change its course, degrading or agrading.
Particular care should be taken to ensure that the aqueduct is not likely to be outflanked by
the river, resulting in possible damage to the irrigation channel leading to or from it.

b Overall Length
In many instances, the overall length of the structure will be obvious as it will have to span
between well-defined banks. However, when spanning large rivers with less well-defined
banks, it may well be worth while determining the regime width of the river. Lacey's empirical
regime equation for stable alluvial channels is, in metric units:

P = 4.75 Q 0.5

Where:
P = the wetted perimeter of the river [m]
Q = is the dominant discharge [m3/s].

The equation may be converted to ft-s units as follows:

P = 2.67 Q 0.5

where P is in feet and Q in ft3/s.

In the context of Balochistan, this equation may be applied to natural rivers in alluvial material
where the dominant discharge is taken as the mean annual flood.

Where the natural river bed width is considerably wider than the regime width P determined
from the above equation, it may be possible to consider shortening the length of the aqueduct
and training the river towards the aqueduct. However, this is unlikely to result in cost saving
due to the cost of the training works, which will usually comprise stone and gabion protected
earthen embankments. The safer option is likely to be to construct a longer aqueduct.

Where the river is narrower than the regime width P then it is likely that either the channel
width is confined by rock or hard strata, or that the banks are ill defined. Where the latter is
the case, the aqueduct length should be set at the regime width, P, and river training works
provided both upstream and downstream of the aqueduct.

c Span Length
For short spans, the cost of the supporting columns makes the overall cost of an aqueduct
unacceptably high. For reinforced concrete (RCC) aqueducts, as the span length increases,
the overall cost of the columns decreases, but the cost per unit length of the troughs
increases as more structural steel is required. Balancing these two considerations, indicates
that the optimum length of each span is approximately 7.6m (25 ft). Similarly, the optimum
length of span for a fabricated steel pipe aqueduct is approximately 12.2m (40ft).

For single spans, the maximum length is about 9.1m (30ft).


BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 43

For multiple span structures, it is common for the troughs to be designed as simply supported
and for joints to be provided between each span. This is wasteful as joints are expensive and
are also prone to leak. By designing the aqueduct as a continuous span over several
columns, the moments and hence the steel requirements can be significantly reduced.

It is therefore recommended that single spans be designed as simply supported, two span
aqueducts as continuous across the central support, and where three or more spans are
required, that the structure be divided into three span units. It is not recommended that more
than three spans are made contiguous, due to the need to allow for thermal movements.

d Aqueduct Height
The height of the aqueduct above the river bed level must take into account all of the
following:
• the need to command cultivated land;
• the maximum design flood level; and,
• vehicle access.

Peak flood flows can be estimated using the methodology given in Part 2: Flood Estimation,
and flood depths calculated using Manning’s equation, as described in Appendix 10 to Part 2.

The soffit of the aqueduct should be 0.9m (3 ft) clear of the 1 in 50 year flood and, preferably,
also clear of the 1 in 100 year flood. If not, the lateral design loads on the structure will
significantly increase and serious consideration will need to be given to alternative structures
such as inverted syphons.

It is necessary to check whether the introduction of piers into the river will themselves create
an appreciable rise in the flood water level at the aqueduct site. This occurs when the flow in
the river is sub-critical (Froude number, Fr < 1), since for supercritical flow, the effects are
carried downstream rather then upstream. The increase in flow depth upstream of the piers
may be determined using Yarnell’s empirical method given in Section 2.1.

e Foundation Level
The foundation level of the columns and abutments is usually determined on the basis of
scour. The scour depth may be calculated using Lacey's equation.

The design scour depth below bed level (D) is given by:

Design scour depth (D) = XR – Y [metric units]

Where:
X = scour factor dependent on type of reach (see Table 5.1 below)
Y = design depth of flow [m]
R = 1.35 (q2/f)1/3
q = the maximum discharge per unit width [m2/s]
f = Lacey’s silt factor
PAGE 44 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Table 5.1 Scour Factors

Type of Reach Mean Value of


Scour Factor "X"

Straight 1.25
Moderate bend (most transitions) 1.50
Severe bend (also Shank protection at spurs) 1.75
Right angled bend (and pier noses and spur heads) 2.00
Nose of Guide Banks 2.25

For channels where the bed material size is well known, the Lacey silt factor (f) may be
calculated from the formula:

f = 1.76 D50

Where:
D50 = the sieve size through which 50% of the material passes by weight [mm].

Alternatively, the silt factor is given in Table 5.2 below for various soil types.

Table 5.2 Lacey’s Silt Factor

Soil Type Lacey's Silt Factor "f"

Large boulders and shingle 20.0


Boulders and shingle 15.0
Boulders and gravel 12.5
Medium boulders, shingle and sand 10.0
Gravel and bajri 9.0
Gravel 4.75
Coarse bajri and sand 2.75
Heavy sand 2.0
Fine bajri and sand 1.75
Coarse sand 1.5
Medium sand 1.25
Standard silt 1.0
Medium silt 0.85
Fine silt 0.6
Very fine silt 0.4
Clay 5.0
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 45

5.4 RCC Trough Aqueduct - Hydraulic Design


As with channels, the capacity of an aqueduct is calculated using Manning=s equation:

V = R2/3 x S½ / n Metric units [m/s]

V = 1.49 x R2/3 x S½ / n Ft-s units [ft/s]

Where (in metric units):


S = river slope (dimensionless)
R = hydraulic radius, = A/P [m]
A = cross-sectional area [m2]
P = wetted perimeter [m]

The discharge (Q) is given by:

Q = VxA [m3/s]

The aqueduct should be capable of passing the same design flows as the channels, and the
design should ensure smooth transitions at either end. This is usually achieved by making the
depth of flow in the aqueduct the same as that in the channels and the overall height of the
aqueduct can then be the same as that of the channels.

Water overtopping at an aqueduct is likely to do


less damage than over topping of the irrigation
channel on an earthen embankment, and a
slightly lower freeboard for the central span(s)
of the aqueduct is advisable.

Aqueducts are physically more difficult (and


hazardous) to desilt than channels. Therefore,
the slope of the aqueduct should be slightly
greater than that of the upstream channel, so
that the velocity of flow in the aqueduct is
greater and the aqueduct therefore less likely to
silt up. However, problems of the levels of the
upstream and downstream channels and the
height of the aqueduct required over the river
may prevent this. The slope of the aqueduct
should however never be less than the slope of
the upstream channel.

For an efficient structural design, it is normal


for aqueducts to have a breadth to depth ratio
Aqueduct, Kunara
of near unity.

Four standard sections for aqueduct troughs covered most of the flow ranges used on both
BMIADP and BCIAP, and are given in the following table.

Table 5.3 Standard Aqueduct Trough Sizes


Type Width (m, ft) Overall Height (m, ft)
A 0.46 (1.5) 0.61 (2.0)
B 0.61 (2.0) 0.76 (2.5)
C 0.76 (2.5) 0.91 (3.0)
D 0.91 (3.0) 1.07 (3.5)
PAGE 46 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

5.5 RCC Trough Aqueduct - Structural Design

a Design Standards
Pakistan does not have its own design standards for the design of reinforced concrete
structures. British or American standards are therefore used. For the structural design of the
RCC aqueducts, this manual has adopted the British design code, BS 8110 "Structural Use of
Concrete", 1985.

b Trough Dimensions
Reynolds and Steedman9 give the minimum cover to reinforcing steel for Class A concrete
(nominal mix proportions 1:1.5:3) as 50mm (2 inches) where the structure is subject to driving
rain, alternate wetting and drying or to freezing when wet. Assuming two layers of main
reinforcement and links, with : inch (#6) main reinforcement and d inch (#3) links, then the
minimum practical thickness for the walls and bed of the trough is 230mm (9 inches).

BS 8110 (Clause 3.4.1.6) proposes that the slenderness of concrete beams without lateral
support should be such that the span is less then 60 times the width of the beam. For a 7.6m
(25 ft) span and 230mm (9 inch) thick beam, the span is 33 times the breadth which is
therefore acceptable.

The dimensions for the four standard troughs adopted for BCIAP perennial irrigation schemes
are shown on Figure 5.1.

c Loads
Loading on an aqueduct trough comprises the deadload of the concrete and the live load of
the water flowing through the trough. The water should be taken as full to the top of the
trough.

Table 5.4 gives the design loadings for the four standard trough sections. The loads assume
a density of concrete of 152 lb/ft3 and a density of water as 64 lb/ft3, and are factored up by
1.4 for dead loads and 1.6 for live loads, as required by BS 8110.

Table 5.4 Loads on Standard Aqueduct Troughs

Trough Type Channel Channel Factored Dead Factored Live Total Factored
Width (ft) Height Load (lb/ft) Load (lb/ft) Load (lb/ft)
(ft)

A 1.5 2.0 1,117 300 1,417

B 2.0 2.5 1,357 500 1,856

C 2.5 3.0 1,570 749 2,319

D 3.0 3.5 1,915 1,048 2,963

9/ Reynolds, CE & Steedman, JC AReinforced Concrete Designer=s Handbook@ 10th Edition. E & FN Spon Ltd,
London, 1988.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 47

d Bending Moments
For a simply supported span, the bending moment M at the centre of the span is:

M = w le2 / 8

Where:
w = load per unit length; and
le = the effective span, taken as the distance between the centre lines of the
columns

For double and triple continuous spans, the bending moments, in terms of the co-efficient
wle2, are reproduced from Reynolds and Steedman10 on Figure 5.2(a). Table 5.5 gives the
maximum bending moments for the four standard trough sections using the loadings from
Table 5.4, the coefficients given in Figure 5.2(a) and assuming a span length of 7.6m (25 ft).

e Reinforcement Design
Table 5.5 gives the moments and required reinforcement for the four standard troughs
calculated according to BS 8110. The design calculation assumes that the aqueduct trough
comprises two beams (the two side walls) each taking half the load. The effect of the trough
floor has been neglected. Reinforcement details for the standard aqueduct troughs for 7.6m
(25ft) long spans are shown on Figure 5.3(a) and Figure 5.3(b).

10/ Reynolds, CE & Steedman, JC AReinforced Concrete Designer=s Handbook@ 10th Edition. E & FN Spon Ltd,
London, 1988.
PAGE 48 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Table 5.5 Bending Moments and Steel Requirements in Aqueduct Troughs

Trough A Trough B Trough C Trough D

Single Span (Simply Supported)

Mid span Moment (lb.ft) 110,500 145,000 180,000 231,000

As (in2) 0.74 1.60 1.71 1.92

Bars 2 no #6 2 no 8# 4 no 6# 3 no 8#

Continuous Over Two Spans

Mid span Moment (lb ft) 62,000 81,200 100,800 130,000

As (in2) 0.41 0.90 0.95 1.07

Bars 3 no 4# 2 no 4# 2 no 4# 2 no 4#
1 no 8# 1 no 8# 2 no 6#

Centre support Moment (lb ft) 110,500 145,000 180,000 231,500

As (in2) 0.74 1.60 1.71 1.92

Bars 3 no 4# 2 no 8# 2 no 8# 2 no 4#
2 no 4# 2 no 8#

Continuous Over Three Spans*

Mid outer spans Moment (lb ft) 70,000 92,800 115,200 148,200

As (in2) 0.47 1.02 1.09 1.23

Bars 3 no 4# 2 no 4# 2 no 4# 2 no 4#
1 no 8# 1 no 8# 2 no 6#

Mid centre span Moment (lb ft) 22,100 29,000 36,000 46,300

As (in2) 0.15 0.32 0.34 0.38

Bars 2 no 4# 2 no 4# 2 no 4# 2 no 4#

Supports Moment (lb ft) 88,400 116,000 144,000 185,200

As (in2) 0.59 1.28 1.36 1.53

Bars 3 no 4# 2 no 8# 2 no 8# 2 no 8#
2 no 4# 2 no 4#

Notes: As = Area of steel required


2 no 6# = 2 number 6/8 (3/4 inch) bars.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 49

f Shear
The shear force V at either end of a simply supported beam may be calculated as follows:

V=kwl

Where:
k = coefficient, which for a simply supported beam is 0.5
w = the load per unit length
l = the length of the beam

The k values for double and triple continuous spans are reproduced from Reynolds and
Steedman in Figure 5.2(b).

The shear stress v is then determined from the following equation:

v = V / (b d)

Where:
V = shear force
b = the breadth of the beam
d = the effective depth of the beam (ie distance from the compressive face to the
centre line of the tensile reinforcement).

As with the determination of the tensile steel requirements, it is not proposed to provide the
details of the calculations for shear reinforcement here, since different codes will provide
somewhat different methods of analysis.

BS 8110 requires that only nominal shear links be provided for each of the standard troughs
under each of the continuity conditions described, and d inch links at 8 inch centers satisfies
the requirements.

g Deflection
BS 8110 (Clause 3.4.6.1) states that the deflection of a beam will not be excessive if the ratio
of the span to the effective depth is not greater than:
• for simply supported beams: 20
• for continuous beams: 26

For trough types A to D, the ratio of span to effective depth is 10, 8.3, 7.1 and 6.2
respectively, which is satisfactory.
PAGE 50 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

h Joints
Standard joint details are shown on Figure 5.4. The detail shown is suitable for expansion
joints between simply supported and two and three span continuous spans. Joints are also
required at the end of each aqueduct.

The actual opening of the joint required to accommodate movement is difficult to evaluate,
although the maximum theoretical space can be calculated on the basis of thermal movement
from the following equation:

S=ktL

Where:
S = the joint movement [ft or m]
k = thermal coefficient of expansion for reinforced concrete (about 6.5 x 10-6 per
F; 11.7 x 10-6 per C)
t = maximum temperature variation [F or C]
L = joint interval [ft or m]

For a 22.9m (75ft) joint interval, and an 80F temperature range, a joint movement of 21mm
(0.47 inches) can be anticipated. In practice, when the trough is full, the movement will be
considerably less than this since the water will reduce the temperature range the concrete is
subjected to.

The joint detail incorporates a 9 inch centre bulb PVC waterstop to be placed in the centre of
the joint. This must be wired to the reinforcement to hold it in place when the concrete is
being poured. The concrete must be placed with care and properly vibrated to ensure that it
passes right round the water stop. The joint is filled with a compressible filler and sealed on
exposed surfaces with a cold applied rubber bitumen sealant. Rubber bitumen sealants are
capable of movement of up to 50% of their thickness. Considerable care must be taken on
site to ensure that the joints are properly fabricated. There is nothing worse than an aqueduct
where all the joints leak.

5.6 Steel Pipe Aqueducts


Steel pipe aqueducts have the advantage of easy quality control and speed of erection.
Typical spans are about 12m (40ft), and the higher cost of the steel pipes may be offset by
the reduced number of piers. Pipe sections may have to be transported in shorter lengths and
welded on site. Alternatively, flanges for bolted connections may be provided. Pipe
connections should be avoided at points of maximum bending moment (at centre spans).

As for culverts and inverted syphons, it is wise for the aqueduct to be large enough to allow
cleaning and maintenance, and this requires a minimum pipe diameter of about 0.9m (3ft).

The hydraulic capacity of the pipe the may be determined using Manning=s equation and
should confirm that the pipe is capable of carrying the maximum design flow without the pipe
becoming full. The maximum design flow should be taken as the bank full capacity of the
channel immediately upstream of the aqueduct.

As for RCC aqueducts, the bed slope of a pipe aqueduct should be the same or, preferably,
greater than that of the upstream channel so that the velocity of flow in the aqueduct is
greater than in the upstream channel, and the aqueduct unlikely to silt up.

In the structural design of the aqueduct and choice of wall thickness, allowance should be
made for corrosion. This is likely to be in the range from 0.1mm/annum to 0.3mm/annum.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 51

Expansion joints are required to cope with thermal expansion, and need to be provided at
intervals of about 61m (200ft) on straight sections, as well as at bends and at either end of the
pipe aqueduct. The amount of thermal expansion may be estimated using the formula given
in Section 5.5h, and using the coefficient of expansion for steel (about 6.7 x 10-6 per F; 12.0
x 10-6 per C).

The pipes should be allowed to move longitudinally over the piers; this may be facilitated by
wrapping the pipes in building paper and coating the pier supports with two coats of bitumen
paint.

5.7 Design of Columns and Foundations

5.7.1 Masonry Columns


The masonry column shown on the drawing on Figure 5.5 is suitable where the height of the
aqueduct from the underside of the trough to the base of the foundation is 3.7m (12ft) or less
and where the velocity of flow is less than 4.6m/s (15 ft/s) in a 1 in 100 year flood.

Using the nomenclature shown on Figure 5.6, the design of the column may be carried out as
follows:

(1) Determine the lateral loads for the worst design conditions. These will be when the
river is in full flood and the wind loading on the trough is in the same direction as the
river flow on the piers. It also assumes that the river has scoured down to the base of
the foundation.

It is recommended that the wind loading be based on a wind speed of 80 miles per
hour, which will produce a pressure on the side of the trough of 1.7kN/m 2 (35 lb/ft2).
The wind load is then the frontal area of one span of the trough times the pressure
(P1).

The water load on the submerged potion of the pier (P2) may be calculated from the
following equation, which is given in both imperial and metric units:

P2 = K A V 2 [lb]

P2 = 0.51 K A V 2 [kN]

Where:
K = pier shape coefficient for water loading as given in Table 5.6.
A = the area of the submerged part of the pier normal to the flow [ft2 or m2]
V = the velocity of flow [ft/s or m/s].

Table 5.6 Pier Shape Coefficients for Water Loading

Shape of Pier Nose Coefficient, K

Square ended piers 1.50


Circular pier or pier with semi circular ends 0.66

Pier ends formed by equilateral arcs of a circle 0.45

Piers with triangular cut ends, with an angle of 60o


between the faces 0.70
As above but with 90o between the faces 0.90
PAGE 52 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

(2) Determine safety against overturning. The overturning moment is given by the
following equation:

Overturning moment = (P1 h1) + (P2 h2)

Where:
P1, P2 = wind, water load on trough and pier respectively
h2 = distance from foundation to center of water load
h1 = distance from foundation to center of trough wall.

It should be noted that the water load due to flowing water may be assumed to act
half way up the pier and not one third of the way up as would be the case for static
water pressure.

The stabilising force comprises the weight of one span of the trough plus the weight
of the pier. It is recommended that the trough be assumed to be empty, to produce
the worst case. The stabilizing moment about the downstream toe of the foundation is
the sum of the weights of the trough and the pier, times half the length of the base.

The factor of safety against overturning may be determined as follows:

Factor of safety against overturning = Sum of the stabilizing moments


Sum of the overturning moments

It is recommended that the factor of safety against overturning should be 2.0.

(3) Determine the factor of safety against sliding. Since the design assumes that the
river has scoured out the river bed material to the base of the foundation, the sliding
force must be resisted purely by the friction between the base slab and the ground.
The factor of safety against sliding may be determined from the following formula:

FoS = (W 1 + W 2 – U) tanφ

(P1 + P2)

Where:
U = uplift forces, zero in this case
W 1, W 2 = vertical forces due to trough & pier
P1 + P2 = horizontal forces due to wind and flowing water
φ = the shearing resistance of the soil, which may be taken as 30 for
gravel and sandy soils (see Part 1: Site Investigations).

It is recommended that the factor of safety against sliding be 1.5 for granular soils and
2.0 for cohesive soils.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 53

(4) Bearing Capacity. It is necessary to check that both the resultant force of the
overturning and stabilising moments occur within the middle third of the base and that
the bearing pressure is within acceptable limits.

The eccentricity E of the resultant force of the overturning and stabilizing moments
from the centre of the base is determined from:

E = (P1 h1) + (P2 h2)

(W 1 + W 2)

The eccentricity must be less than one sixth of the length of the pier base slab in
order that tensile forces do not occur under the heal of the foundation slab.

(W 1 + W 2) 6E( W 1 + W 2 )
P max / min = 
B.L B. L 2

The maximum and minimum pressures under the base slab may be found as follows:
The minimum pressure under the heal must be greater than zero, whilst the
maximum must be less than the safe bearing capacity (refer Part 1: Site
Investigations). For loose gravel or sand and gravel, the safe bearing capacity is
about 290 kN/m2, 30 tonnes/m2 (2.8 tons /ft2).

(5) Earthquake loading


Much of Balochistan is prone to earthquakes. Earthquakes impart accelerations to
structures which usually increase the effective loadings. The Geological Survey of
Pakistan Publication "Earthquake Risk in Quetta" suggests that for an operational
basis earthquake (that which has a high expectancy of occurring during a structures
life) a factor of 0.25 to 0.35 to the gravitational force should be applied horizontally to
the structure. However, it is not recommended that this loading condition be
considered to occur simultaneously with the 1 in 100 year flood flow. Given that
reduced factors of safety are usually allowed for earthquake conditions, it is unlikely
that the forces from the above earthquake loading will be more onerous than those
resulting from the 1 in 100 year flood. It is not therefore considered necessary to
check the effects of earthquake loading on aqueducts.

5.7.2 Reinforced Concrete Columns


The reinforced concrete column and foundation shown on Figure 5.7 and Figure 5.8 are
suitable where the height of the aqueduct from the underside of the trough to the base of the
foundation is 6.1m (20ft) or less, and where the velocity of water is 4.6m/s (15 ft/s) or less in a
1 in 100 year flood.

The stability of the column is based on the same criteria as described for the masonry column
in Section 5.7.1. In addition, the strength of the column has to be checked to resist:
• bending and shear at just above the base of the column
• bending and shear across the toe of the base.
PAGE 54 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Most codes will require that the effective height of the column is determined. For this the
column may be considered to be restrained against direction and rotation at the base and to
have no restraint at the top. According to BS 8110, the column's effective height is then twice
the actual height. A 6.1m (20ft) high column, with a width of 0.61m (2ft), must be considered
slender and additional moments due to the eccentricity of the load must therefore be
considered in the design. The design loading condition therefore becomes:
• overturning moments due to:
(a) the wind force on the trough
(b) the water pressure on the pier
(c) the eccentricity of the load of the trough and the pier
• the vertical load on the base from:
(a) the weight of the trough
(b) the weight of the pier
(c) in the case of the forces on the base, the weight of the base itself.

Concrete is susceptible to erosion by abrasion from stones and gravel being thrown against it
by flood flows in the river. Where it is considered that the velocities and nature of the bed load
in the river are such that the column is likely to be eroded away, the concrete of the column
must be protected. One means of achieving this is shown on the drawing on Figure 5.8.
Here, the nose of the column has been extended to incorporate a masonry nose, which is
fully bonded into the main concrete column. Another method of protection is to wrap the
bottom section of the column in a steel sheet.

5.8 Abutments
Two sizes of abutment are shown on the drawings. The smaller size shown on Figure 5.4 is
suitable for heights up to 3.6n (12 ft), while the larger size shown on Figure 5.7 and Figure
5.8 is suitable for heights up to 6.1m (20ft).

Both abutments are stone masonry with mass concrete bases. The main difference between
the structures being the size of the steps which are set to suit the overall height of the
structure. The abutments are short and incorporate a return which makes them inherently
stable. Stability calculations are not therefore required for these short abutments.

However, should it be necessary to design a longer abutment wall, or walls with greatly
different heights, then they should be checked for overturning, sliding, bearing pressure and
structural integrity at all levels. The loads will be as for the columns, except for the lateral
earth pressure from the backfill, and any surcharge loading.

The depth of the abutment should be the same as that for columns, except where the
orientation of the river is such that greater or lesser scour is expected on either bank of the
river.

The drawings show a reinforced concrete transition flume being placed on top of the
abutments, between the channel and the aqueduct. This flume is intended to both act as a
hydraulic transition between the shape of the channel and the aqueduct and as a short
launching section, which will accommodate a small amount of settlement of the backfill
behind the abutment. It is however, very important, that the soil behind the abutment is
properly compacted.

The drawings do not show any protection around the abutments. However, the backfill to the
banks around the abutments is liable to scour as a result of the river currents and additional
eddies created by the abutments. It is therefore recommended that stone pitching or gabion
protection be provided on the banks upstream and downstream of the abutments, extending
to below scour depth.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 55

Figure 5.1 Standard RCC Aqueduct Trough Dimensions

Type A 9” 1’-6” 9”

2’

9”

9” 2’ 9”
Type B –
6”
-”

2’ – 6”

9”

9” 2’ – 6” 9”
Type C

3’

9”

9” 3’ 9”
Type D

3’-6”

9”
PAGE 56 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 5.2(a) Moment Coefficients for Simply Supported and Continuous Spans
(Note: all spans loaded equally)

0 0

0.125

0 0.125 0

0.070 0.070

0 0.100 0.100 0

0.080 0.025 0.080

Figure 5.2(b) Shear Force Coefficients for Simply Supported and Continuous Spans
(Note: all spans loaded equally)

0.50 0.50

0.375 0.375

0.625 0.625

0.400 0.400

0.600 0.500 0.500 0.600


BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 57

Figure 5.3(a) Standard Aqueduct Trough Details


PAGE 58 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 5.3(b) Standard Aqueduct Trough Details (cont)


BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 59

Figure 5.4 Standard Single Span Aqueduct


PAGE 60 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 5.5 Standard Double Span Aqueduct and Masonry Column


BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 61

Figure 5.6 Design Loads on a Masonry Aqueduct Pier


PAGE 62 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 5.7 Triple Span Aqueduct with RCC Column


BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 63

Figure 5.8 Triple Span Aqueduct: Details


PAGE 64 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 65

6 INVERTED SYPHONS

6.1 Purpose and Description


Inverted syphons are used to convey water by gravity under (drainage) channels, rivers, roads
or other structures. A syphon is a closed conduit designed to flow full and under pressure.
The structure should operate without excess head when flowing at design capacity. A typical
arrangement for a syphon is shown on the drawings on Figure 6.1 and Figure 6.2.

While usually cheaper than aqueducts, syphons have the disadvantage that they are easily
blocked by silt and other debris. They should therefore only be used as a last resort and never
on channels carrying a large silt load.

If access to the pipeline is required for inspection and cleaning, the minimum diameter of the
pipeline should be about 0.9m (3 ft). However, for very small discharges this could result in an
excessively large pipe being selected thereby increasing the cost, and reducing low flow
velocities within the pipe.

Transitions are usually required at the inlet and outlet of a syphon to reduce head losses and
prevent canal erosion in unlined channels.

Pipe collars to prevent piping due to water moving along the outside of the pipe are not
usually needed, but this may be confirmed.

Washout structures may be provided at the low point of relatively long inverted syphons to
permit draining of the pipe for inspection, and to wash out accumulated sediment.

Canal free board should be increased upstream of a syphon by 50% to allow for back up of
flow due to blockage. Trash racks are often placed upstream of the syphon to prevent entry of
trash into the syphon.

Overflow spillways and waste ways may be provided upstream of a syphon to prevent
damage to the channel bank in the event of the syphon becoming blocked.

6.2 Hydraulic Design


The hydraulic design for a syphon is primarily concerned with the head losses through the
various components of the structure. The allowable head loss will depend on the relative
elevations of the upstream and downstream canals. The length of the pipeline will depend on
design flow conditions in the river. The inlet and outlet chambers should be set sufficiently
back from the river banks so that they will be affected by bank erosion.

Having selected a trial pipe size, the total head loss for the design flow should be calculated
to ensure that it is less than the available head loss. An economical design will be one in
which the available head balances the head loss for the design flow plus at least 10% as a
safety factor. Under BCIAP, a factor of safety of 30% was used. Where the design head loss
is greater than the available head, a larger pipe should be selected, or possibly, multiple
pipes.

Particularly for long syphons, the pipe diameter will determine the syphon cost and this
therefore should be as small as hydraulically possible. However, the need for access for
inspection and cleaning should be considered, and may result in a minimum pipe diameter of
0.9m (3ft) being selected.
PAGE 66 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Head losses in a syphon comprise three components:

a Trash Rack Losses


Head loss through the trash racks may be calculated as shown in Section 2.2. For the trash
rack shown in Detail R on Figure 6.2, for channel velocities of up to 1.8m/s (6 ft/s), the head
loss would be about 50mm (2 inches). However, for design purposes, it should be assumed
that the trash rack is partially blocked. It is therefore recommended that for this type of trash
rack a head loss of at least 100mm (4 inches) be allowed.

b Transition Losses
Head losses due to transitions is dependent on the change in velocity head through the
transition, and may be calculated as detailed in Section 2.3 using the appropriate loss energy
coefficients for submerged flow, which are likely to be about 0.5 for the inlet and 0.65 for the
outlet. Losses flowing into and out of each inspection chamber (if provided) must also be
allowed for.

c Pipe Friction Losses


The headloss due to friction in the pipe can be determined using Manning=s equation as
detailed in Section 2.5, or, more accurately, the Colebrook-White equation. The latter is more
complex to use, but results have been tabulated by HR Ltd11. Table 6.1 gives pipe friction
losses per 100 feet of pipe, for poor quality concrete pipes, for various flow velocities and for
a range of pipes sizes based on the HR Ltd tables.

Table 6.1 Friction Losses in Concrete Pipes

Flow Velocity Headloss (ft) per 100ft length of pipe


(ft/s)
24" dia pipe 30" dia pipe 36" dia pipe

1 0.02 0.02 0.01

2 0.10 0.07 0.06

3 0.22 0.16 0.12


4 0.38 0.29 0.19

5 0.60 0.45 0.36


6 0.85 0.65 0.52

6.3 Design Features


The syphon pipeline needs to be set below scour depth as calculated for a design flood of 1 in
50 years using Lacey=s method.

The recommended maximum spacing for manholes along the pipeline is about 76m (250 ft).
The roof of each manhole should be fitted with a hinged, lockable, steel hatch to facilitate
access for inspection and cleaning. To prevent any ingress during floods, the top of the each
manhole in the river bed should be set either 0.6m (2ft) above the flood water level for a 1 in
50 year flood, or 0.6m (2ft) above the water level inside the manholes during operation,
whichever is the higher.

11/ HR Ltd ATables for the Hydraulic Design of Pipes and Sewers@ Hydraulics Research Ltd, Wallingford, UK.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 67

The joint between the manholes and the pipe is made with thickened mass concrete infill,
both to ensure that the pipe is well supported and to provide good bond between the pipe and
the wall. A flexible (spigot and socket) joint should be fitted either side of each manhole to
accommodate differential settlement between the pipe and the manhole.

A trash rack is required on the inlet to prevent large debris entering the syphon; either in the
flow or from children trying to throw things in. The trash rack shown in Detail R on Figure 6.2
has bars at 76mm (3 inch) spacings in one direction only. It has been found that this is about
the ideal spacing as it is close enough to prevent large material entering the syphon, but not
too close that it would quickly clog with small pieces of material and cause the upstream
channel to overtop. For large trash racks, steel flats should be used instead of bars, as the
thinner but wider steel flats will provide a stronger trash rack for the same head loss.
PAGE 68 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 6.1 Syphon General Arrangement


BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 69

Figure 6.2 Syphon Details


PAGE 70 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 71

7 ROAD CROSSINGS: CULVERTS AND BRIDGES

7.1 Introduction
Many irrigation structures incorporate a culvert or a bridge in order to facilitate vehicular
access within the scheme. For example, turnouts may have a culvert to take the offtaking flow
through the parent canal bank, and checks may incorporate bridge or walkway decks.

Where a road cross a canal at some distance from any other structure, it will be necessary to
provide an independent culvert or bridge. Under BCIAP, road crossings comprised either
concrete pipe or box culverts to convey canal water under the road, or concrete slab bridge
decks supported on concrete or masonry piers.

The choice between culvert and bridge will normally be made on economic grounds: for small
flows pipe culverts are almost always cheaper. Concrete box culverts are suitable for larger
flows (say more than 100 cusecs, about 3 m3/s).

Culverts may be designed to flow full, while bridge decks are always kept clear of the free
water surface. Culverts fall into two main categories: box or rectangular culverts and circular
or pipe culverts.

7.2 Culverts

7.2.1 Hydraulic Characteristics


The head losses though a culvert depends on whether the cuvert is submerged, in which
case losses are as for an inverted syphon with appropriate losses for submerged transitions,
or whether the water surface is free in which case flow is as for flow in an open channel with
appropriate losses for open surface transitions and barrel friction. Losses due to submerged
or free water surface transitions are described in Section 2.3.2.

The characteristics of flow through culverts is well described by Chow12 which where the
following description was taken.

A culvert will flow full when the outlet is submerged, and also when the outlet is not
submerged but the headwater is high and the barrel is long. According to laboratory
investigations, the entrance of an ordinary culvert will not be submerged if the headwater is
less then a certain critical value, designated by H *, while the outlet is not submerged. The
value of H * varies from 1.2 to 1.5 times the height of the culvert depending on the entrance
geometry, barrel characteristics, and approach conditions. For a preliminary analysis, the
upper limit H * = 1.5d may be used, where d is the height of the culvert, because computations
have shown that, where submergence is uncertain, greater accuracy could be obtained by
assuming that the entrance is not submerged.

Laboratory investigations also indicate that a culvert, usually with a square edge at the top of
the entrance, will not flow full even if the entrance is below headwater level when the outlet is
not submerged. Under these conditions, the flow entering the culvert will contract to a depth
less then the height of the culvert barrel in a manner very similar to the contraction of flow in
the form of a jet under a sluice gate. This high velocity jet will continue through the barrel
length, but its velocity will be reduced slowly as the head is gradually lost by friction. If the
culvert is not sufficiently long to allow the expanding depth of flow below the contraction to
rise and fill the barrel, the culvert will never flow full. Such a culvert is considered hydraulically
short. Otherwise the culvert is hydraulically long, for it will flow full like a pipe.

12/ Chow, VT AOpen Channel Hydraulics.@ McGraw-Hill Book Company. 1959.


PAGE 72 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Whether a culvert is hydraulically short or long cannot be determined by the length of the
barrel alone. It too depends on other characteristics, such as slope, size, entrance geometry,
headwater, entrance and outlet conditions, etc. A culvert may become hydraulically short,
that is it may flow partly full, even when the headwater is greater than its critical value.

For practical purposes, culvert flow may be classified into six types, as shown on Figure 7.1.
The identification of each type of flow may be explained according to the following:

A. Outlet Submerged ................................................................. Type 1

B. Outlet Unsubmerged
1. Headwater greater than critical value
a. Culvert hydraulically long .............................. Type 2
b. Culvert hydraulically short ............................. Type 3
2. Headwater less then the critical value
a. Tailwater higher than critical depth ............... Type 4
b. Tailwater lower than critical depth
1. Slope sub-critical .................................... Type 5
2. Slope super-critical ................................. Type 6

Flow Types 1 & 2


Flow types 1 and 2 are “culvert full“ flow, and head losses may be determined for the
submerged transitions and for friction loss in the pipe barrel, as for inverted syphons (see
Section 6.2).

Flow Types 3, 4, 5 & 6


For Type 3 flow, the culvert acts like an orifice. For Type 4, 5 and 6 flows, the entrance is not
sealed by water and it acts like a weir. As shown on Figure 7.1, Type 4 flow is subcritical
throughout the barrel length. Type 5 flow is subcritical and, hence, the control section is at the
outlet. Type 6 flow is supercritical and, hence, the control section is at the entrance.

For free water surface flow, head losses may be determined for the free water surface
transitions and for friction losses in the culvert barrel. Alternatively, flow may be estimated
using the charts in Figure 7.2(a) for box culverts and Figure 7.2(b) for pipe culverts. These
charts are applicable to culverts having square edged entrances.

7.2.2 Design Considerations for Concrete Pipe Culverts

a Hydraulic Design
A typical arrangement for a canal road pipe culvert is shown on Figure 7.3

Pipe culverts are generally designed to flow full, with the pipe soffit submerged by at least
0.20m (0.7ft) both upstream and downstream. The pipe normally has a fall of 0.05m (0.2ft).

Design head loss through concrete culverts is normally 100mm (4inches), but can be reduced
to 50mm (2 inches) if necessary by using a larger culvert. Actual headloss is estimated for the
submerged transitions and for friction loss in the pipe barrel. Values are tabulated in Table
7.1.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 73

Table 7.1 Concrete Pipe Culvert Design Head Loss


Pipe Diameter Flow Capacity (m3/s) Total Head Loss (m)
(m) V = 0.7 m/s V = 1.0 m/s V = 0.7 m/s V = 1.0 m/s
0.30 0.05 0.07 0.06 0.13
0.375 0.08 0.11 0.06 0.12
0.45 0.11 0.16 0.05 0.11
0.60 0.20 0.28 0.05 0.10
0.75 0.31 0.44 0.05 0.09
0.90 0.45 0.64 0.04 0.09
1.05 0.61 0.87 0.04 0.09
1.20 0.79 1.13 0.04 0.09
2 X 0.75 0.62 0.88 0.05 0.09
2 x 0.90 0.89 1.27 0.04 0.09
2 x 1.05 1.21 1.73 0.04 0.09
2 x 1.20 1.58 2.26 0.04 0.09
3 x 0.90 1.34 1.91 0.04 0.09
3 x 1.05 1.82 2.60 0.04 0.09
3 x 1.20 2.38 3.39 0.04 0.09
Note: Assumes 10m total pipe length

For unlined channels, flow velocities in excess of 1.0 m/s are not recommended as this is
likely to result in scour downstream. If the available headloss through the culvert is
substantially greater than 100mm (4 inches), then pipe velocity can be more than 1.0 m/s
(3.3ft/s) if provision for energy dissipation is provided downstream. This most suitable outlet in
this case would be the USBR Type 6 impact basin, which is described in the USBR
publication “Hydraulic Design of Stilling Basins and Energy Dissipators”.

b Pipe Strength, Bedding and Cover


Concrete pipes are available in several strength classes and there are a number of bedding
alternatives. The choice of pipe class and appropriate bedding is affected by various factors
including:
• design loading on the road
• depth of cover to the pipe (it is often practicable to raise the road in order to provide
minimum cover)
• availability of the stronger pipes (often only low grade pipes are locally available)
• trench conditions (these can usually be specified)
• cost.

For most roads within an irrigation scheme, which will be subjected to loading from
agricultural plant and trucks, it is generally unnecessary to design for main road loading.
Culverts which pass under main roads should be designed individually using the appropriate
loading for the location (normally based on BS153 Type HB Loading).

Table 7.2 gives the strength requirements for four classes of concrete pipe. The ultimate load
quoted is, in effect, the minimum permissible value of ultimate strength. Ultimate strength is
defined as the load per unit length which the pipe can sustain without collapse. The proof load
quoted is that which the pipe can sustain without developing a crack exceeding certain
specified dimensions.
PAGE 74 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Table 7.2 Minimum Crushing Test Loads (kilograms per linear meter of effective length)
Nominal Standard pipes Extra Strength Pipes
bore of Class S Class L Class M Class H
pipe Proof Ultimate Proof Ultimate Proof Ultimate Proof Ultimate
(mm)
150 2010 2380
225 2010 2380
300 2010 2380 2380 2980
375 2010 2380 3120 3900 3720 4650
450 2010 2380 3570 4460 4170 5210
525 2010 2380 3870 4840 4610 5760
600 2010 2380 4610 5760 5510 6870
675 2010 2380 5060 6320 6100 7620
750 2010 2380 3780 4840 5360 6700 6550 8190
825 2010 2380 4170 5210 5810 7250 6990 8740
900 2010 2380 4610 5760 6850 8560 8630 10790
975 2010 2380 4910 6140 7300 9110 9220 11530
1050 2010 2380 5210 6500 7740 9670 9820 12280
1125 2010 2380 5350 6690 8330 10410 10710 13390
1200 2010 2380 5800 7250 8780 10980 11160 13940
1350 2010 2380 6400 8000 9680 12100 12360 15430
1500 2010 2380 6990 8780 10560 13200 13400 16750
1650 2010 2380 7580 9480 11750 14690 14880 18600
1800 2010 2380 8330 10240 12650 15800 16070 20090
Notes:
1. Pipes of nominal bores larger than 1800 mm are also available for any agreed strength.
2. For information on the application of the above pipe strengths for design purposes, reference may be made
to the national Building Studies Special Report 37, ‘Loading charts for the design of buried pipelines’, or to
‘Simplified tables of external load on buried pipelines (revised edition)’. Both are published by HMSO.
3. Un-reinforced pipes are not normally available in sizes and strength classes to the right of and below the
shaded cells in the Table.
4. Source: BS 556, Part 2, 1972.

For safe design it is necessary to ensure that the computed total design load, W e does not
exceed the test strength multiplied by the appropriate bedding factor, Fm, and divided by a
factor of safety, Fs.

W e  W T Fm / Fs

If the ultimate load in Table 7.2 is used for W T, then Fs = 1.25. If the proof load is used, Fs =
1.00.

However, for reinforced pipes it is necessary to design for both conditions; that is,

W e  W T(ult) Fm / 1.25
and
W e  W T(proof) Fm / 1.00

In the paper “Simplified tables of external loads on buried pipelines” (HMSO, 1969), values of
W e have been tabulated for an range of pipe diameters depths of cover for three loading
conditions: main roads, light roads and fields. Adopting the light roads case, safe depths of
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 75

cover have been computed and these are presented in Table 7.3 and Table 7.4. Table 7.3
applies only to single pipes in “narrow trench” conditions. Multiple pipes and single pipes for
which the trench width is greater than the specified width are covered in Table 7.4 The four
pipe bedding alternatives with appropriate values of Fm are shown on Figure 7.4.

Table 7.3 Safe Depths of Cover for Single Pipes in Narrow Trenches Assuming
Light Road Loading
Pipe Class Bd Safe range of depth and cover (m)
diameter (m) Class of Bedding
(m) C B A1 A2
0.30 S 0.75 * 1.0 to 4.3 0.9 to 7.6 0.9 to 7.6
L - - - -
M - - - -
H 1.0 to 4.1 0.9 to 7.6 0.9 to 7.6 0.9 to 7.6
0.375 S 1.05 * 0.9 to 3.0 0.9 to 5.3
L - - -
M 0.9 to 4.3 0.9 to 7.6 0.9 to 7.6
H 0.9 to 6.5 0.9 to 7.6 0.9 to 7.6
0.45 S 1.15 * 1.1 to 2.2 0.9 to 4.2
L - - -
M 0.9 to 4.6 0.9 to 7.6 0.9 to 7.6
H 0.9 to 6.6 0.9 to 7.6 0.9 to 7.6
0.60 L 1.35 - - -
M 0.9 to 4.4 0.9 to 7.6 0.9 to 7.6
H 0.9 to 6.0 0.9 to 7.6 0.9 to 7.6
0.75 L 1.50 * 0.9 to 4.2 0.9 to 7.3
M 0.9 to 4.3 0.9 to 7.6 0.9 to 7.6
H 0.9 to 6.5 0.9 to 7.6 0.9 to 7.6
0.90 L 1.90 * 0.9 to 3.5 0.9 to 5.6
M 0.9 to 4.1 0.9 to 7.1 0.9 to 7.6
H 0.9 to 6.0 0.9 to 7.6 0.9 to 7.6
1.05 L 2.05 * 0.9 to 3.5 0.9 to 5.7
M 0.9 to 4.2 0.9 to 7.1 0.9 to 7.6
H 0.9 to 6.2 0.9 to 7.6 0.9 to 7.6
1.20 L 2.30 * 0.9 to 3.4 0.9 to 5.4
M 0.9 to 4.1 0.9 to 7.0 0.9 to 7.6
H 0.9 to 6.2 0.9 to 7.6 0.9 to 7.6
Notes:
1 Class C bedding is not recommended for pipe diameters greater than 0.30 m.
2 Pipe strength classes L and M (0.30m diameter) and L (0.375m, 0.45m and 0.60m diameter) are
not available.
3 * indicates pipe is unsuitable.
4 Bd = assumed trench width. The tabulated values of safe depth of cover are valid only if the trench
width is less than or equal to Bd. Where the trench width exceeds Bd refer to Table 7.4.
5 Light loading is defined as a group of two wheels, 0.9 m apart, each having a static weight of
70kN, with an impact factor of 1.5, both wheels acting simultaneously.

The minimum depth of cover to the pipe is 0.90m (3ft). If this is impracticable, then
consideration should be given to:
• surrounding the pipe in concrete;
• using smaller pipes (for example by using 2 x 0.60 m diameter pipes instead of a
single 0.90 m diameter pipe)

The maximum depth of cover considered in the tables is 7.6m (25 ft). In many instances a
larger depth may be acceptable, but this should be checked when the situation arises.
PAGE 76 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Selection of Pipe Strength Class and Bedding


It can be seen from Table 7.3 and Table 7.4 that, in certain circumstances, there could be a
choice of three pipe strength/bedding combinations. For example, a 0.75 m (2.5ft) diameter
single pipe culvert with a maximum depth of cover of 6.0 m (20ft) could be:
• a Class H pipe on Class B bedding;
• a Class M pipe on Class A1 bedding; or
• a Class L pipe on Class A2 bedding.

Generally the Class B (Granular) bedding is preferred because it is the simplest. However, if
Class H pipes are not available or the contractor is unable to manufacture them, then a
concrete bedding may be necessary.

Table 7.4 Safe Depths of Cover for Pipes in Wide Trenches and Multiple Pipes
Assuming Light Road Loading
Pipe Class of Safe range of depth of cover (m)
dia (m) Pipe Class of bedding
C B A1 A2
0.30 S * 1.0 to 2.7 0.9 to 4.1 0.9 to 5.6
L - - - -
M - - - -
H 1.0 to 2.5 0.9 to 3.7 0.9 to 5.3 0.9 to 7.0
0.375 S * 0.9 to 3.0 0.9 to 4.2
L - - -
M 0.9 to 3.7 0.9 to 5.4 0.9 to 7.1
H 0.9 to 4.6 0.9 to 6.4 0.9 to 7.6
0.45 S * 1.2 To 2.2 0.9 to 3.4
L - - -
M 0.9 to 3.6 0.9 to 5.2 0.9 to 6.9
H 0.9 to 4.3 0.9 to 6.1 0.9 to 7.6
0.60 L - - -
M 0.9 to 3.4 0.9 to 5.0 0.9 to 6.6
H 0.9 to 4.3 0.9 to 6.0 0.9 to 7.6
0.75 L * 0.9 to 3.2 0.9 to 4.5
M 0.9 to 3.2 0.9 to 4.8 0.9 to 6.3
H 0.9 to 4.2 0.9 to 5.9 0.9 to 7.6
0.90 L 1.1 to 2.0 0.9 to 3.3 0.9 to 4.6
M 0.9 to 3.6 0.9 to 5.2 0.9 to 6.9
H 0.9 to 4.8 0.9 to 6.7 0.9 to 7.6
1.05 L 1.2 to 2.0 0.9 to 3.2 0.9 to 4.5
M 0.9 to 3.6 0.9 to 5.1 0.9 to 6.8
H 0.9 to 4.8 0.9 to 6.6 0.9 to 7.6
1.20 L 1.2 to 1.9 0.9 to 3.1 0.9 to 4.3
M 0.9 to 3.5 0.9 to 5.1 0.9 to 6.7
H 0.9 to 4.7 0.9 to 6.5 0.9 to 7.6
Notes:
1 Class C bedding is not recommended for pipe diameters greater than 0.30 m.
2 Pipe strength classes L and M (0.30m dia) and L (0.375m, 0.45m and 0.60m dia) are not
available.
3 * indicates pipe is unsuitable.
4 Light loading is defined as a group of two wheels, 0.9 m apart, each having a static weight of 70
kN, with an impact factor of 1.5, both wheels acting simultaneously.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 77

c Transitions
Transitions are required at both the inlet and outlet of the culvert barrel. An accelerating water
velocity usually occurs at the inlet, and a decelerating velocity at the outlet. Transitions reduce
head losses and prevent canal erosion by making the velocity changes less abrupt. Concrete
or stone masonry transitions are usually used. One example is shown on the pipe culvert road
crossing on Figure 7.3. Concrete or stone masonry transitions are required for all culverts
with a 0.9m (3ft) pipe diameter or larger, and for all culverts where the flow velocity is 1.1m/s
(3.5ft/s) or more discharging into an earthen channel. For smaller pipe culverts and lower flow
velocities stone protected earthen transitions are acceptable.

d Pipe Collars
Pipe collars may be required to reduce seepage flows along the outside of the pipe, thereby
preventing removal of soil particles at the point of emergence (piping). This should be
checked using Lane’s seepage path length method, particularly for any culvert where the
canal water surface is significantly higher than the potential point of relief of the percolating
water. Pipe collars may also be necessary to discourage rodents from burrowing along the
pipe.

e Canal Erosion Protection


Canal erosion protection is required in earthen channels, and usually takes the form of dry
stone pitching placed upstream and downstream of the transitions. In addition, for large flows,
cutoffs may be provided to the transition structures extending below scour depth.

The thickness of the stone pitching is 0.20m (8 inches) on 0.15m (6 inches) gravel backing for
flows up to about 1m3/s (35.3 ft3/2). For flows larger than this, the thickness of the stone
pitching is 0.30m (12 inches) on 0.15m (6 inches) gravel backing.

As a result of model tests, the US Corps of Engineers13 have developed an equation for
determining the length of downstream protection required from a pipe culvert:

L = 1.7 D (Q / D5/2) + 8 (ft - s units)

Where:
L = length of downstream protection from the pipe outlet to the end of the
protection works (ie including the length of any concrete or masonry
transition) [ft].
D = diameter of the pipe culvert, or in the case of a box culvert, the width
of the culvert [ft]
Q = Design discharge of the culvert [cusecs].

The total length of the upstream protection (concrete or stone masonry transition plus stone
pitching) should be about half the length of the downstream protection.

13/ US Corps of Engineers AHydraulic Design Criteria@


PAGE 78 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

7.2.3 Design Considerations for Box Culverts

a General
A typical arrangement for a canal road box culvert on an unlined canal is shown on Figure
7.5.

Box culverts can be single or multiple, with square or rectangular section. They can be
constructed from reinforced concrete, or comprise concrete floor and roof stabs with masonry
walls. The minimum height of the culvert should be about 0.76 m (2.5 ft) since anything
smaller than this is probably more easily constructed using concrete pipes.

The width of the culverts should be equal to the width of the irrigation canal, with a minimum
width of 0.6m (2 ft).

A typical road crossing box culvert for a lined small perennial irrigation canal as used on
BCIAP is shown on Figure 7.6. The structure is a simple masonry culvert with RCC roof and
base slabs to take the traffic load and small wing walls.

The length of the culvert should normally be sufficient to allow a 4m (13.1ft) width of road,
which is sufficient for single line traffic across the culvert.

It is important that the culvert is founded on firm foundations.

Compacted earth ramps are provided on either side of the culvert to raise the road from
natural surface level to culvert deck level. It is important that the ramp is built of suitable fill,
well compacted, to 95% of maximum compaction.

b Hydraulic Design
Reinforced concrete or masonry box culverts are used for larger canals, generally for flows in
excess of about 3 m3/s (106 ft3/s).

Box culverts are normally designed for free flow, that is with the water surface below soffit
level. A minimum freeboard of 0.10m may be adopted. Headloss across the culvert can be
calculated by summing entry and exit losses, plus friction loss in the culvert which may be
determined using Manning’s equation.

Box culverts can be designed to flow full, in which case the soffit both upstream and
downstream should be drowned by at least 0.20m (8 inches) to prevent “gulping”.

It is recommended that the culvert invert is, in any case, not depressed more than d/4 below
the bed level, where d = depth of flow in the culvert.

c Structural Design
Structural design needs to take into account loads due to water, soil, self weight, and
vehicles, as appropriate, as well as the strengths of the materials.

For the standard box culverts used as road crossings for the BCIAP perennial irrigation
canals (see Figure 7.6), the culvert reinforced concrete deck has been designed to resist
vehicle loads as specified in British Standard 153 Part 3A "Specification for Steel Girder
Bridges". This loading comprises an HA uniformly distributed loading of value shown on
Figure 7.7, acting together with a knife edge load of 2,700 lb/ft across the width of the bridge,
positioned at the worst location. The bridge deck is designed to resist bending, shear and
punching shear. The two reinforcing details shown are for clear spans of up to 0.61m (2 ft)
and for 2 to 4 ft (0.61 to 1.22m). This will be satisfactory for most perennial irrigation
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 79

channels. It is recommended that the reinforced concrete deck is topped by a 75mm (3 inch)
thick concrete wearing course.

Short wingwalls (say less than 3.0m, 10ft) should not require a movement joint at the junction
with the culvert. For longer walls consideration should be given to providing a joint, and all
walls over 6.0m (20ft) long should have a joint. Movement joints at intervals along the culvert
are necessary for long culverts. A single continuous length of up to 15m (50ft) is probably
acceptable. For lengths in excess of this joints at intervals of 10 to 15m (33 to 50ft) are
required.

d Scour and Piping


As for pipe culverts flexible protection (stone pitching) and cutoffs may be required to control
scour. Also, and particularly if the culvert barrels are provided with stop logs, checks must be
carried out for “piping”, and possibly for differential hydro-static loads between barrels and
flotation.

7.3 Bridges

7.3.1 Road Bridges


Under BCIAP, independent road
bridges were only required on the
flood irrigation schemes for
crossing the relatively large canals.
Otherwise it was cheaper to use
pipe or box culverts.

Standard designs for road bridges


were not developed under BCIAP.
However, for completeness of this
design manual, standard bridge
decks designed as simply
supported reinforced concrete
structures crossing clear spans of
1.0m, 2.0m, 2.5m 3.0m and 4.0m
are shown on Figure 7.8. Multiply
span structures may be used to RCC Road Bridge, Toiwar FIS
accommodate larger spans.

The roadway width is generally 4.0m (13.1ft), or 6.0m (19.7ft) for major roads. A 1 in 40
camber should be provided, at least 0.35m (1.1ft) wide kerbs either side, and also 0.90m (3ft)
high hand railing comprising 50mm (2 inch) diameter GI pipe. Drainage pipes (100mm
diameter) are provided at mid span in decks of 2.5m (8ft) clear span and over.

Traffic loading has been taken as HA to BS 153, with additional reinforcement around the
hand railing bases.

The bridge deck is supported by a continuous rubber strip bearing at either end, with one end
fixed by means of dowel bars and one end free.
PAGE 80 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

The minimum structural free board between design water level in the canal and the bridge
deck soffit is given in Table 7.5.

Table 7.5 Bridge Deck Soffit Freeboard


Design Discharge Minimum Freeboard
(m3/s) (m)
0 – 1.5 0.20
1.5 – 3.0 0.30
> 3.0 0.40

Flexible stone protection in unlined canals upstream and downstream of bridges may be as
for regulators (see Section 3.8).

7.3.2 Footbridges
Footbridges or reinforced concrete walkways are often provided at regulating structures for
operation of gates or stop logs.

They are normally between 0.90m (3 ft) and 1.5m (5 ft) wide depending on their use, and are
sometimes provided with hand railing on one or both sides. They are generally designed as
simply supported reinforced concrete slabs (maximum span 4.5m, 15ft) with a live loading of
4kN/m2 based on the gross plan area.

The minimum deck soffit freeboard should be as for road bridges.

Standard walk ways designed as simply supported reinforced concrete structures crossing
clear spans of 1.0m, 2.0m, 2.5m 3.0m and 4.0m are shown on Figure 7.9. Multiply span
structures may be used to accommodate larger spans.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 81

Figure 7.1 Types of Culvert Flow


PAGE 82 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 7.2(a) Chart for Estimating Flow through Rectangular Culverts with square edged
entrances

Figure 7.2(b) Chart for Estimating Flow through Pipe Culverts with square edged entrances
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 83

Figure 7.3 Typical Road Crossing Pipe Culvert for an Unlined Canal
PAGE 84 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 7.4 Pipe Bedding Alternatives


BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 85

Figure 7.5(a) Typical Road Crossing Box Culvert for an Unlined Canal
PAGE 86 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 7.5(b) Typical Road Crossing Box Culvert for an Unlined Canal (cont.)
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 87

Figure 7.6 Typical Road Crossing Box Culvert for a Lined Perennial Irrigation Canal (BCIAP Standard Drawing)
PAGE 88 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 7.7 Imposed Loads from Vehicles


BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 89

Figure 7.8 Standard Bridge Decks

2” GI Pipe
Hand railing 0.5” bars at spacing 0.5” bars at 6”
as for bar A centres, EW
UPVC
drainage
pipe

1 in 40
0.15m

0.35m S/2

Bar A

Suggested Bridge Deck Dimensions & Principal Reinforcement


Bridge Clear Deck Width Deck Thickness Main Reinforcement (Bar A)
Span (S) (X) Bar Diameter Spacing
(m) (m) (m) (inches) (inches)
1.0 4.0 or 6.0 0.30 0.5 4.0
2.0 4.0 or 6.0 0.30 1.0 6.0
2.5 4.0 or 6.0 0.35 1.0 6.0
3.0 4.0 or 6.0 0.35 1.0 6.0
4.0 4.0 or 6.0 0.40 1.0 6.0
PAGE 90 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 7.9 Standard Walkways

2” GI Pipe Hand
railing
(as required)
3/8” dia bars at
6” spacings Bar A

Varies
(Typically 0.9m – 1.5m)

Suggested Walkway Thickness & Principal Reinforcement


Walkway Clear Deck Thickness Main Reinforcement (Bar A)
Span (X) Bar Diameter Spacing
(m) (m) (inches) (inches)
1.0 0.10 3/8 6.0
2.0 0.15 3/8 6.0
2.5 0.20 3/8 4.0
3.0 0.20 0.5 4.0
4.0 0.20 6/8 6.0
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 91

8 CROSS DRAINAGE STRUCTURES

8.1 Introduction
The need for cross drainage structures results from the flow of drainage or storm runoff water
from the high side of the canal to the low side. To protect the canal from such flows, cross
drainage structures are provided.

While the alignment of a canal usually follows natural ground contours, it is often necessary to
cross natural drainage channels. In crossing large natural drainage channels or nullahs, the
canal flow is usually conveyed in an aqueduct over, or in an inverted syphon under, the
drainage channel. For small drainage flows, it may be cheaper for the drainage flow to be
passed under, over or into a canal using one of the following cross drainage structures:
• culvert
• superpassage
• overchute
• drain inlet

Culverts are suitable for carrying small drainage flows under a canal which is in fill. Culverts
are unsuitable for large cross drainage flows which could be transporting boulders and other
debris which would block the culvert. Superpassages can take much larger cross drainage
flows than culverts, are not vulnerable to blocking by debris, and are suitable where the canal
is in cut allowing drainage flow to pass over the canal. Overchutes, also convey drainage
water over the top of a canal. Rectangular overchutes are able to transport a considerable
amount of debris, whereas the pipe overchute, like a culvert, is subject to blockage. If the
design flow for a drainage area is very small, it may be desirable to allow the drainage water
in the canal through a drain inlet.

8.2 Cross Drainage Culvert Structures


Culverts fall into two main categories, box or rectangular culverts and circular or pipe culverts.

When used to carry small drainage flows under a canal which is in fill, the hydraulic
characteristics and design considerations are similar to those for road crossings (see Section
7.2), except that loading due to vehicles need not be considered. Also, for cross drainage
culverts under canals, the outlet is likely to be free-flowing, with the culvert rather than
downstream conditions providing a restriction on the flow. Hence the flow will be Types 3 to 6
rather than Types 1 or 2 (see Figure 7.1).

Figure 8.1 and Figure 8.2 show typical designs for a pipe culvert and for a box culvert as
used on BCIAP for cross drainage.

8.2.1 Design Process for Pipe Culverts


The design process for a cross drainage pipe culvert is as follows:
• Estimate the design drainage flow based on local observations of flood levels and
survey data for channel size and slope. The flood estimation method described in
Part 2: Flood Estimation is inappropriate for micro-catchments such as where one
might need to construct a culvert.
• Carry out the hydraulic design of the pipe culvert. Select an appropriate trial pipe size.
The minimum diameter which should be considered is 0.6m (2ft) in order to allow for
cleaning. There should also be enough room to fit the pipe under the channel without
having to set the invert of the pipe below natural ground level so as to ensure that free
flow conditions exist.
• Determine the length of pipe from the geometry of the canal embankment.
• Determine the maximum allowable head of water upstream of the culvert. This must
take into account a freeboard of at least 0.6m (2ft) against water overtopping the
PAGE 92 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

canal embankment. Also, the head of water on the upstream side of the embankment
should not be sufficient to cause a piping failure of the embankment. It is suggested
that, given the short duration of such flows, the hydraulic gradient from the point of
intersection of the upstream water level with the upstream face of the embankment
and the downstream toe of the embankment should not be greater than 1 in 6. If
necessary provide concrete collars to increase the seepage path.
• Using the graph on Figure 7.2(b), determine the flow the culvert can accommodate
for the allowable head upstream of the culvert. If this is greater than the cross
drainage design flow then the design is satisfactory. If the capacity of the pipe is less
than the design flood flow then try a larger pipe, or a box culvert.
• Design the inlet and outlet transitions. Concrete or stone masonry transitions are
required for all culverts with a 0.9m (3ft) pipe diameter or larger, and for all culverts
where the flow velocity is 1.1m/s (3.5ft/s) or more.
• Design the inlet and outlet downstream protection works. The length of the protection
may be in accordance with Section 7.2.2e (Canal Erosion Protection), but gabion
mattresses should be used instead of stone pitching.
• Design the pipe bedding material. Granular bedding under the pipe as shown on
Figure 8.1 is adequate unless the canal embankment is very high. The clear cover to
culvert from the bed of the canal should be 0.3m (1ft), otherwise the culvert should
perhaps be encased in concrete. For further details of bedding requirements refer to
Section 6.6 of Part 4: Infiltration Galleries.

8.2.2 Design Process for Box Culverts


The design process for a cross drainage box culvert is as follows:
• Estimate the design drainage flow based on local observations of flood levels and
survey data for channel size and slope. The flood estimation method described in
Part 2: Flood Estimation is inappropriate for micro-catchments such as where one
might need to constructing a culvert.
• Carry out the hydraulic design of the box culvert by selecting an appropriate box
culvert size to suit the following criteria:
(1) the height of the culvert should be small enough to fit beneath the channel,
without being buried below natural ground level, in order to ensure that free
flow conditions exist.
(2) The minimum height of the culvert should however be 0.76m (2.5 ft) since
anything smaller than this is probably more easily constructed using concrete
pipes.
(3) It is recommended that the width of the cross drainage culvert be kept at
1.2m (4 ft), since larger spans will require a heavily reinforced top slab, and
smaller spans would make cleaning difficult. Where a single box of maximum
possible height is insufficient to carry the design flood flow, then multiple
boxes or a small aqueduct may be used.
• Determine the length of culvert from the geometry of the canal embankment.
• Determine the maximum allowable head of water upstream of the culvert. This must
take into account a freeboard of at least 0.6m (2ft) against water overtopping the
canal embankment. Also, the head of water on the upstream side of the embankment
should not be sufficient to cause a piping failure of the embankment. It is suggested
that, given the short duration of such flows, the hydraulic gradient from the point of
intersection of the upstream water level with the upstream face of the embankment
and the downstream toe of the embankment should not be greater than 1 in 6. If
necessary provide concrete collars to increase the seepage path.
• Using the graph on Figure 7.2(a), determine the flow the culvert can accommodate,
per foot width of culvert, for the allowable head upstream of the culvert. Multiply this
by the width of the culvert to obtain the total flow the culvert can carry. If this is
greater than the cross drainage design flow then the design is satisfactory. If the
capacity of the culvert is less than the drainage design flow, then try a higher or
multiple box culvert.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 93

• Design the inlet and outlet transitions. Concrete or stone masonry transitions are
required for all culverts with a 0.9m (3ft) pipe diameter or larger, and for all culverts
where the flow velocity is 1.1m/s (3.5ft/s) or more.
• Design the inlet and outlet downstream protection works. The length of the protection
may be in accordance with Section 7.2.2e (Canal Erosion Protection), but gabion
mattresses should be used instead of stone pitching.
• No bedding, apart from a firm foundation, is required for box culverts.

8.3 Superpassages
Superpassages can take much larger cross drainage flows than culverts, are not vulnerable
to blocking by debris, and are suitable where the canal is in cut allowing drainage flow to pass
over the canal. Superpassages are often used where the canal is passing around a steep
hillside and many nullahs have to pass over the channel.

Figure 8.3(a) and Figure 8.3(b) show various standard arrangements for the superpassages
constructed under BCIAP. The main differences between them arise from how downstream
protection is provided for. The drawings also leave the final alignment of the wingwalls and
any embankments to be determined by engineers on site once construction has started since
this is the easiest time to do it. The design process for superpassage is as follows:
• Estimate the cross drainage flow at the location. This will normally have to be done
based on local observations of flood levels and survey data for channel size and
slope. The flood estimation method described in Part 2: Flood Estimation is
inappropriate for micro-catchments such as where one might consider constructing a
superpassage.
• Determine the size of the irrigation channel conduit, as follows:
(1) The minimum height of the conduit should be 0.76m (2.5ft) in order to allow
cleaning. Where the channel depth is greater than 0.76m (2.5ft), the depth of
the conduit should be equal to the channel depth.
(2) For rectangular channels, the width of the conduit should be equal to the
width of the channel, with a minimum width of about 0.61m (2 ft).
(3) For trapezoidal channels, the width of the conduit should be the bed width of
the channel plus the height of the channel, again subject to a minimum width
of 0.61m (2ft). A transition should be provided from the channel to the
conduit.
(4) Headloss across the culvert can be calculated by summing entry and exit
losses, plus friction loss in the culvert barrel which may be determined using
Manning’s equation.
• The length of the superpassage required to carry the design cross drainage flow is
difficult to determine as the geometry of the structure is complex. Given that most
superpassages are on steeply sloping land, it is likely that the level top slab of the
superpassage will tend to act as a weir, and it is therefore recommended that the weir
equation be used to determine the capacity of the structure.
• Having determined the design depth of flow and length of the superpassage, the
height of the wall above the top of the roof slab of the superpassage should be
selected to provide a minimum of 150mm (6 inches) of freeboard.
• Upstream and downstream guide walls need to be constructed, as required according
to the local topography, to guide the cross drainage flow to and over the
superpassage. Where the guide walls become unduly long, it is recommended that
they be extended by stone pitched earthen guide bunds.
• Upstream and downstream bed protection needs to be provided to ensure that the
structure is not undermined by local scour.

For protection where the land is not excessively steep, 0.46m (1.5ft) thick gabion mattress
launching aprons are recommended. Gabion launching aprons will launch into a scour hole of
depth equal to two thirds of their length. Hence the 3.6m (12ft) downstream apron will launch
into an 2.4m (8ft) hole and the 1.8m (6ft) upstream apron will launch into a 1.2m (4ft) hole; both
of which should be adequate for most circumstances. For small cross drainage flows, the
PAGE 94 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

upstream gabion apron may be replaced with stone pitching.

Where the hill side on the downstream side of the channel is steep, it is recommended that
the wall and downstream gabion are stepped down (see Figure 8.3(b)). The steps will
dissipate the energy resulting from the fall in stages, avoiding the need for costly protection
works at the base of the fall. The length and proportions of the protection works have to be
adjusted to suit site conditions.

8.4 Overchutes
Overchutes convey drainage water over the top of a canal and are sometimes used when the
canal is in cut. Rectangular overchutes are able to transport a considerable amount of debris,
whereas a steel pipe overchute, like a culvert, is subject to blockage.

The minimum freeboard to the overchute soffit should be as for bridges, and as given in
Table 8.1 below.

Table 8.1 Overchute Soffit Freeboard


Design Discharge Minimum Freeboard
(m3/s) (m)
0 – 1.5 0.20
1.5 – 3.0 0.30
> 3.0 0.40

A mild slope of the cross drainage channel is usually accompanied by a mild slope and slow
velocity in the overchute, permitting the use of a standard concrete or stone masonry inlet and
outlet transitions, such as those described for culverts.

A steep slope of the cross drainage channel may be accompanied by a steep slope and fast
velocity in the overchute. If the overchute has an exit velocity exceeding 6m/s (about 20ft/s),
an energy dissipator is required at the outlet to dissipate the excess head.

The cross drainage flow at the location will normally be estimated based on local observations
of flood levels and survey data for channel size and slope. The flood estimation method
described in Part 2: Flood Estimate is inappropriate for micro-catchments.

The design of the conduit and piers is generally as for aqueducts, and is covered in Chapter
5.
Note: Caution should be used to avoid a slope at or near the critical slope, as critical flow is
rather unstable. A small change in energy will cause a relatively large change in water
surface. To avoid excessive water surface undulations from unstable flow, if flow in
the chute is designed to be supercritical, the supercritical slope adopted should
exceed the critical slope by about 20%, with computations based on a rough “n” (eg
0.015 in lieu of 0.013 for a concrete section). For the hydraulic design of chutes
carrying supercritical flow refer to “Design of Small Canal Structures”, USBR.

8.5 Drain Inlets


Drain inlets are structures used to carry relatively small amounts of storm runoff or drainage
water into the canal. Drain inlets are of three basic types:
• Pipe section;
• Rectangular concrete section; or
• Stone pitched blanket

Under BCIAP stone pitched inlets were proposed for the main flood channel at Toiwar, but
later cancelled. The design of inlets is not covered by this manual. The reader is referred to
“Design of Small Canal Structures”, USBR for guidance on the design of drainage inlets.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 95

Figure 8.1 Typical Cross Drainage Pipe Culvert used on BCIAP


PAGE 96 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 8.2 Typical Cross Drainage Box Culvert used on BCIAP


BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 97

Figure 8.3(a) Standard Superpassage used on BCIAP


PAGE 98 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 8.3(b) Standard Superpassage used on BCIAP (Alternative Details


BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 99

9 ESCAPES

9.1 General
An escape generally comprises an overflow structure, such as a side-channel spillway, in
combination with a drop or chute structure and a wasteway channel. Escapes allow automatic
release of water from an irrigation canal, and are perhaps particularly required for flood
irrigation schemes to where the flood flows that enter the irrigation channel can be quite
variable.

9.2 Side Channel Spillways

9.2.1 General
Side channel spillways are located
the canal bank with the spillway
crest parallel to the canal
alignment. As the canal water
surface rises above the crest, the
excess water is automatically
discharged into the side channel
from where it is conveyed to the
wasteway channel. The structure
may be provided with a slide gate
to allow for complete drainage of
the canal. A check structure may
also be provided a short distance
downstream to permit division of
the canal flow into the wasteway.

On BCIAP, a concrete overflow


spillways were provided in Side Spillway Escape, Toiwar FIS
conjunction with an orifice flow
regulating structure on the Toiwar
flood channel near the headworks, so that most of the excess flow diverted by the headworks
would spillover into the wasteway channel.

9.2.2 Design Considerations

a Overflow Spillway
The design capacity of the spillway and wasteway depend on its function. On flood irrigation
schemes, the capacity is based on the excess flow which may enter the irrigation channel
during a 30 year flood.

The spillway crest level should be set about 0.06m (0.2ft) higher than the design water level to
prevent unnecessary waste of water from normal wave action. To provide a reasonable safety
factor against overtopping, the maximum water surface encroachment on freeboard is
normally limited to:
• 50% - 75% of the canal lining freeboard; or
• 25% - 50% of the canal bank freeboard in an unlined canal section.
PAGE 100 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

The width of the spillway may be computed using the appropriate free flow formula, of the
type given below:

Q = C b h3/2 (m3/s)

Where:
Q = discharge [m3/s]
C = coefficient (=1.7 for broad crested weirs);
b = width of the spillway [m]
h = upstream depth of flow over the spillway [m]

b Energy Dissipation
An energy dissipation structure is required downstream of the spillway. The options include:
• a vertical drop for drops of up to about 2.4m (8ft), with canal protection for some
distance in the wasteway downstream. About 50% of the excess energy is dissipated
as the water strikes the basin floor and the remaining head is lost in a hydraulic jump,
which is usually submerged, and in residual turbulence downstream (refer Chapter 4
for design of the stilling basin for a vertical drop);
• a glacis type drop for drops of more than about 2.4m (8ft), where energy dissipation is
by means of a hydraulic jump which forms at the bottom of the glacis. A USBR stilling
basin is required downstream of the glacis, the design of which is covered in Part 3:
Weirs;
• other structures such as a pipe drop or baffled outlet, not covered by this Manual.

Flexible protection is required downsteam of the energy dissipation structure to prevent


erosion of the wasteway channel. This generally requires a length of stone pitching or gabions
(refer to Section 3.8 or to Part 3: Weirs).

c Structural and Stability Considerations


Structural and stability checks need to consider the following:
• sliding and overturning
• seepage (percolation) water removing foundation material
• scour and undermining of the structure
• foundation pressures which need to be less than the maximum allowable bearing
pressures.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 101

10 SOCIAL STRUCTURES

10.1 Community Participation in Site Selection


It is important to involve the community both in understanding how their irrigation system
works and in designing an improved system which meets their needs and which will be
accepted by the whole community. Experience has shown that without this schemes tend to
fail. The community needs to be involved in the design and selection of sites for social
structures such as:
 wuzu structures;
 road crossings;
 foot bridges;
 animal drinking troughs; and,
 washing structures.

The traditional division of labour in many villages makes the design and location of the first
two structures the primary concern of the men and the last the primary concern of the women.

10.1.1 Involvement of Farmers


It is important to take care in the alignment of the irrigation channel, and in the siting of social
and other structures, to involve farmers. Farmers are the primary users of the irrigation water
and their views should be incorporated into the design of the irrigation system. If properly
designed structures are not provided at appropriate locations, the community will modify or
sometimes even damage the conveyance system in order to meet their specific needs. For
example, a farmer wishing to cross a channel with his tractor and finding no properly
designed crossing where he needs it, is likely to still drive over the channel, damaging it in the
process.

10.1.2 Involvement of Women


It has often been forgotten that although women usually have no recognised shareholdings
and are not directly involved in irrigation, they are nevertheless users of the irrigation system.
Besides cooking, cleaning and washing, women’s tasks may include livestock husbandry and
kitchen gardening (ie small scale irrigation). The design team should therefore include women
members who can interview village women about their water needs and their preferences
regarding the design and location of social structures and alignment of the channel.

10.1.3 Involvement of Other Groups


One other group which may traditionally use water from the irrigation channel is migratory
livestock herders. It is obviously difficult to include these groups in the design process since
the farmers will object to their inclusion in discussions and they are only in the area for a short
period each year. However, the design team should be alert to the importance of providing,
say, animal drinking structures for the nomad=s animals. If they are not provided, the
nomads will instead block the channel with stones to raise the water level; leading to
overtopping and damage to the channel.

10.2 Wuzu Structures


Standard wuzu structure details (as constructed under BCIAP) are given on Figure 10.1. In
consultation with the community, wuzu structures should be placed adjacent to mosques for
use in ritual ablutions.

10.3 Animal Drinking Troughs


A standard animal drinking trough as constructed under BCIAP is show on Figure 10.2. One
important feature of the structure is that its ramp allows animals to approach the water=s
edge. This is particularly important for sheep which will not stand in deep water to drink but
cannot drink from water below the level of their feet. It is important to ask local people how
PAGE 102 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

many animals normally drink from a particular point on the channel at any one time so that the
structure can be made a suitable length.

10.4 Washing Structures


A standard washing structure is show on Figure 10.3. The structure comprises a number of
washing positions where the women can wash their clothes and a small steel gate which the
women can use to temporarily raise the water level in the channel. A foot bridge is provided
adjacent to the structure to aid access.
BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 103

Figure 10.1 Wuzu Structure (as constructed under BCIAP)


PAGE 104 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

Figure 10.2 Animal Drinking Trough (as constructed under BCIAP)


BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES PAGE 105

Figure 10.3 Washing Structure (as constructed under BCIAP)


PAGE 106 BCIAP DESIGN MANUAL PART 6: IRRIGATION STRUCTURES

BIBLIOGRAPHY

Chow, Ven Te 1978 Open Channel Hydraulics, Pub. Mc Graw Hill.

FAO 1975 Small Hydraulic Structures, FAO 26/1 and 26/2

Halcrow Group Ltd 1992 BMIADP. Design Manual for Perennial Irrigation Canals and
Canal Structures

Halcrow Group Ltd - Irrigation Manual. Drop Structure Design

Halcrow Group Ltd 1997 MRLCS. Hydraulic Design Criteria for Structures

ILRI 1978 Discharge Measurement Structures, Pub. International


Institute for Land Reclamation and Improvement,
Wageningen, Holland.

Khanna 1987 Practical Civil Engineers' Handbook, Pub. Engineers'


Publishers, New Delhi.

MMP 1986 Hydraulic Design of Canal Structures

USBR 1983 Hydraulic Design of Stilling Basins and Energy Dissipators.

USBR 1978 Design of Small Canal Structures

Varshney, Gupta & Gupta 1977 Theory & Design of Irrigation Structures

You might also like