You are on page 1of 136

International Training Workshop

Design and Evaluation of Pressurized


Irrigation Systems
March 3-7, 2009
Sponsored by
Islamic Development Bank
Prof. Dr. Muhammad Latif
Director
Cent re of Excellence in Wat er Resources Engineering
University of Engineering and Technology
Lahore - Pakistan
www.cewre.edu.pk
1
Unit I
SPRINKLER IRRIGATION SYSTEM AND ITS TYPES
INTRODUCTION
Sprinkler and trickle irrigations together represent the broad class of 'pressurized'
irrigation methods. In these systems water is conveyed through a network of pipes to a point in
the field where water is applied. The basic difference between surface irrigation and pressurized
irrigation is that in surface irrigation methods, soil surface of upper parts of the field is used to
transport water to lower parts of the field. Thus, surface irrigation methods are much more
affected by soil topography, infiltration and soil type as compared to the pressurized irrigation
method. Basic concepts about sprinkle irrigation and its type are discussed in this lecture.
Water is applied in droplets like rain by a sprinkler system. This method has been
developed since the early part of 20th century primarily from irrigation of lawns, orchards and
nurseries. Reduction in cost due to improvement in technology and development of light weight
aluminum pipe led its extensive use in irrigation of field crops and vegetables.
A pump is normally used to lift water from the source and pressurize it to throw into the
air for distribution by sprinklers. Other components of a sprinkle system include main pipeline,
sub-mains, laterals, valves and sprinkler heads which distribute water across the surface of the
field. A typical layout of a sprinkler system is shown in Fig. 1.1 whereas Fig. 1.2 shows different
components of a trickle irrigation system. Sprinkle irrigation has some advantages over surface
irrigation such as:
i. Sprinkler irrigation can be used on such undulating lands which are difficult or impossible
to irrigate by surface methods.
ii. Sprinkler irrigation is suitable for light and frequent watering whenever needed, such as for
germination and frost protection.
iii. Sprinkler irrigation may potentially provide better uniformity of application and high
application efficiency.
iv. More effective use of a small, continuous stream of water such as from springs or dug
wells.
Prof. Dr. Muhammad Latif, Director, Centre of Excellence in Water Resources Engineering,
Lahore
2
Figure 1.1 Primary components of a typical sprinkle irrigation system.
3
Figure 1.2 Components of a drip irrigation system.
Irrigation of steep and rolling topography is possible by sprinkling without producing runoff or
erosion.
v. Normally sprinkle irrigation is more efficient than surface irrigation which will result in
water saving and thus more land can be irrigated with the same water supply.
vi. Much less labour is required in sprinkle method as compared to surface methods
particularly with centre-pivot sprinkler system.
The sprinkle method has also some disadvantages which are mainly related to high cost, water
quality, climatic and environmental constraints. To off-set the cost, high value crops, vegetables
and young orchards are frequently irrigated by sprinkling. If it is unavoidable to use salty water
by sprinkling, crop damage may be reduced if good quality water is applied at the end of
irrigation to wash any salt deposited on the leaves. The use of sprinkler irrigation is less
favorable under extremely dry and windy conditions due to increased evaporative water losses
and poor distribution pattern. The efficiency may generally be improved by sprinkling during
nights and less windy periods.
TYPES OF SPRINKLE IRRIGATION SYSTEMS
There are many types of sprinkle systems and several versions of each type. Sprinkle
irrigation systems may be divided into two main groups: 1) set systems that operate when the
4
sprinklers are set at a particular position in the field, and 2) the continuous move systems that
operate while the system is moving in the field based on portability, systems may be further
classified as portable, semi-portable, semi-permanent and permanent systems. Centre-pivot
system and linear move system are examples of continuous moving system.
Portable System
A portable system has portable main lines, sub-mains, laterals and a portable pumping
plant. The entire system is designed to be moved from field to field or to different sites in the
same field. This system may be moved manually or by mechanical power. In manually moved
system called 'hand move system', the pipes are moved manually. This system has low initial
cost but labour cost is high. In the mechanically moved system, the lateral is mounted on wheels
and is moved as a unit instead of one pipe at a time. This system has high initial cost but labour
requirement is less.
Semi-Portable System
This system is similar to a fully portable system except that the location of water source
and pumping plant are fixed. Such a system may be used on more than one field where there is
an extended main line, but may not be used on more than one farm unless there are additional
pumping plants (see Fig. 1.3).
Figure1.3: Sprinkler with Portable Stand.
Semi-Permanent System
A semi-permanent system has permanent mains and sub-mains which are usually burried
with risers at suitable locations for connecting the laterals, which are portable. Water source and
pumping plants are stationary.
5
Permanent System
A fully permanent system commonly named as solid-set sprinkler system consists of
permanently laid mains, sub-mains, laterals, stationary water source and pumping plant. Mains,
sub-mains and laterals are usually burried. Sprinklers are permanently installed on each riser.
Such systems are costly and are suited to automation. This type of system is usually installed in
orchards, Golf courses and nurseries.
SET-MOVE SYSTEMS
Set-move sprinkle systems are periodically moved from one position (irrigation) to
another by hand or mechanically. These systems remain stationary as water is applied where they
are set. When the desired amount of water has been applied, water is turned off: the pipes are
drained and the system is moved to the next position. When the move is completed, water is
turned on and irrigation is resumed at the next position. This process is repeated until the entire
field has been irrigated. Common types of set-move systems include hand-move, tow-move,
side-roll and gun-type.
Hand-move systems
Components of the system are moved by uncoupling the pipes to the next position by
manual labour. This system became popular after the development of light weight aluminum
pipes. Most hand-move laterals are 50 to 150 mm (2 to 6 in) in diameter. Length of pipe sections
may be either 6, 9 or 12 m (20, 30 or 40 ft). These systems are least expensive but have high
labour requirement. An example of hand-move system is shown in Fig. 1.4.
Figure 1.4: Schematic View of Hand-Move Sprinkler Irrigation System.
Tow-move system
The long lateral of the tow-move sprinkle is mounted on wheels or skids so that the
system can be mechanically pulled to the next position. A tractor is normally used to pull the
6
lateral after draining water. This system is the least expensive mechanically move system. Tow-
move systems are not used extensively because moving the lateral is tedious and also damage
crop. This system has been generally used for forage and row crops.
Side-roll system
This system is very popular type of mechanically moved system. Sections of long lateral
pipes are supported on wheels with the lateral serving as axle of the wheels (see Fig. 1.5). Length
of the lateral may be as long as 800 m (about one half of a mile) and diameter is normally 100 to
125 mm (4 to 5 in). A gasoline engine is usually fixed in the center of the lateral to drive it to the
next position. Common wheel diameters used for side-roll systems vary from 1.2 to 1.9 m. The
wheel diameter must be large enough to allow the lateral to pass over the crop without damaging
it. Water is supplied either in the middle of the lateral or at the end of the system. The friction
losses within the lateral pipe are reduced when water is supplied at the center of the lateral.
Figure 1.5: A Side-Roll Sprinkler System.
Gun and Boom Sprinklers
Gun or giant sprinklers have 16 mm (5/8 in) or larger nozzles attached to long discharge
pipes. Boom sprinklers have 18 to 36 m long rotating arms and water is applied through nozzles
on the arms (See Figures 1.6 and 1.7). The anns, or booms are supported by a cable suspension
system mounted on a four-wheel trailer. Boom sprinklers apply more uniform water in smaller
droplets than gun sprinklers. But gun sprinklers are simpler to operate than booms and also less
7
expensive for the farmer. Discharge of gun and boom sprinklers vary from 65 to 78 l/s (l00 gpm
to 1250 gpm) operated at high pressures ranging from 480 to 896 kPa (70 to 130 psi). Wetting
diameter may be as large as 180 m.
Raingun Sprinklers
A typical stationary or traveling raingun sprinkler system consists of a pumping plant, flexible
hose, traveler assembly and gun sprinkler (see Figure 1.7). A stationary or fix gun irrigates at one
setting which is shifted to the next field. Whereas the traveling gun is mounted on a cart and it
moves continuously across the field while applying water. The assembly is pulled either by the
flexible pipe itself or by an anchor cable stretched across the field. The unit is moved to the next
irrigation position after irrigation is completed at the first setting. Both full and part-circle
sprinklers of large size can be used equally but the part-circle sprinklers are preferred as the cart
moves in relatively dry soil. The gun sprinklers have trajectory angles ranging between 18 to 32
degrees but for average conditions, trajectories between 23 to 25 degrees give satisfactory
results. The traveling system can be used on all field sizes and shapes. These systems can be
easily transported from field to field.
Gun and boom sprinklers can be used for most crops, but their application rates are high
and large sized water drops may compact the soil surface and cause surface runoff. Therefore,
these sprinklers are more suitable for coarse-textured soils.
Figure 1.6: A Gun Sprinkler with Rocker Arm.
8
Figure 1.7: Hose-Fed Traveling-Gun Sprinkler.
CONTINUOUS MOVE SYSTEMS
Laterals and sprinklers of the continuous move systems are constantly moving when
applying water. These systems have become more common due to low labour cost. Main types
of continuous moving sprinkler systems are center-pivot (CP), traveler and linear-move systems.
Center-pivot irrigation systems
The center-pivot sprinkler irrigation systems are the most popular sprinkle systems used
in many countries as shown in Fig. 1.8. A center-pivot system consists of a pipeline supported on
towers usually spaced about 61 m apart. The length of the lateral pipeline varies from
approximately 200 to 800 m but 400 m is the most common length. A center-pivot system of 400
m long lateral covers an area of 65 ha (160 acres). Water is introduced at the pivot and flows
outward through the pipeline, supplying to each of the individual sprinkler heads. Some of the
main advantages of center-pivot sprinklers are (Keller and Bliesner, 1990, p. 307)
i. Water delivery is simplified through the use of a stationary pivot point.
ii. Guidance and alignment are controlled at a fixed pivot point.
iii. Relatively high water application uniformities are easily achieved under the continuously
moving sprinklers.
iv. After completing irrigation cycle, the system is at the starting point for the next irrigation.
v. Achieving good irrigation management is simplified because accurate and timely
application of water is made easy.
9
vi. More accurate and timely applications of fertilizer and other chemicals are possible by
applying them through the irrigation water which is also possible with other sprinkler
systems as well.
vii. Flexibility of operation makes it feasible to develop electric-load-management schemes.
Figure 1.8: A Center-Pivot Irrigation System
Because this system irrigates a circular area, only about 79 percent of the area of a square
field is effectively irrigated by the basic unit (Figure 1.9). With an end-gun modification, an
additional five percent of the total area can be irrigated, leaving 16 percent of the area un-
irrigated. Not irrigating 16 percent of the area increases the cost per unit area of the irrigated
land. This problem can be overcome, to a large extent, by the addition of a 'corner-pivot' system
to the conventional center-pivot machine as shown in the above figure. The corner-pivot
assembly pivots about the outer tower and applies water at the corners of the field. This increases
the irrigated area to approximately 96 percent and reduces the investment per unit irrigated area
appreciably (Callies 1978). The corner-pivot unit is folded with the last tower of the main system
when not in use.
Center-pivot systems vary a great deal in their design. Some systems use only one type of
sprinkler head and vary the spacings between the sprinkler heads on the lateral to obtain
desirable application patterns, while others use constant spacings and vary the size of the
nozzles. Combining the two principles results in other designs, in which both the spacings
between the sprinkler heads and size of the nozzles are varied. As the perimeter of a circle is
10
proportional to the radial distance and the sprinkler heads at different distances from the pivot
point travel at different speeds, the time during which water is applied to a point is also different
along the lateral. To compensate for this and to achieve reasonably uniform depth of water, the
application rate is low near the pivot and increases toward the end tower. Such a distribution of
application rates may cause undesirable runoff near the end of the system on soils of low water
intake rate.
Figure 1.9 Area irrigated by the main system, corner-pivot assembly and endgun of a center-
pivot system (only one-fourth of field is shown).
11
Linear Move Systems
Linear-move systems have been developed in response to the problems of irrigating the
corners and runoff associated with the center-pivot system. Linear move systems have towers
with electric motors and alignment systems, just as do center-pivot systems. The lateral moves
continuously in a linear fashion across the field rather than rotating about a central pivot point.
Water is supplied to the lateral through a flexible hose hooked to a mainline or by a traveling
pumping plant that pumps water from an open ditch. Recently, a linear-move system that moves
continuously and obtains water from risers connected to a buried mainline by automatically
connecting and disconnecting itself from riser valves has been developed.
Linear moving systems may travel forth and back, i.e. they travel twice the length of the
field to complete each irrigation cycle. Hence, the sprinklers may operate continuously forth or
back, or for only half of the total distance traveled on both sides. Each method has its own
advantages and disadvantages.
If the sprinklers are operated continuously, the total operating time will be longer. This
will reduce system flow and application rate requirements. However, the area that is most
recently watered near the end of the field will be immediately re-watered each time on the move
of lateral in reverse direction. On the other hand, if the system is returned with the water shut off
(or empty) the operating time will be decreased accordingly.
The continuous versus part-time watering and the possibility for backtracking over
relatively dry rather than fully irrigated soil leads to following irrigation strategies.
1. Irrigate in only one direction and return empty (with the water off) as quickly as possible
to the starting point.
2. Irrigate while traveling at the same speed in both directions,
3. Irrigate half the field and quickly continue to the other end with the water off, then on the
return irrigate the other half of the field and quickly continue back to the starting point
with the water off, and
4. Proceed as in strategy 3 but continue as quickly as possible with the water on (instead of
off) during the last half of travel in either direction.
The purpose of strategy 3 is to eliminate returning across soil that has just been fully
irrigated. Otherwise it is like strategy 1. The purpose of strategy 4 is to reduce the hazard
associated with watering back across the area at the end that has just been fully irrigated.
However, both strategies 3 and 4 require more management and labour than 1 and 2.
COMPONENTS OF A SPRINKLE IRRIGATION SYSTEM
Primary components of all sprinkle systems are similar in many respects. They consist of
sprinkler with or without risers, laterals, main pipeline and pumping plant as shown in Fig. 1.1.
Different types of valves and couplings also form integral parts of any system. More detail of the
above components is discussed next.
12
Types of Sprinklers
The sprinklers also called sprinkler heads form an important component of any sprinkle
system. Water is sprayed like raindrops through these sprinklers. There are many types of
sprinklers available to suit different farming and climatic conditions. The sprinklers may be
either rotating or fixed type. Rotating sprinklers have three types: impact, reaction and gear-
driven sprinklers. Fixed-head sprinklers include spray-type and perforated pipes.
a) Rotating Head Sprinklers
i) Impact Sprinklers: Impact sprinklers have one or more nozzles that discharge jets of
water into the air. These jets are rotated in a start and stop manner by a spring loaded arm which
strikes (impacts) and then is bounced out of one of the jets. The spring returns the arm to strike
the jet again and the process is repeated. Several different nozzle types have been developed for
impact sprinklers including constant-diameter, constant discharge, and diffuse-jet nozzles.
Constant-diameter nozzles are the most commonly used with impact sprinklers. The discharge
from these nozzles is proportional to the square root of the operating pressure. Small single
nozzle sprinklers are designed at low pressures, while large multiple-nozzles require high
operating pressure. Usually, low cover larger areas and also have high application rates per
nozzle.
Constant-discharge nozzles are also used with impact sprinklers. These nozzle pressures
are associated with a small diameter nozzle, small wetted area and low water application rates.
On the other hand, large size nozzles require high operating pressures, are constructed so that as
long as the operating pressure exceeds a threshold value, changes in pressure do not affect
sprinkler discharge significantly. For example, constant discharge nozzles can be used to
minimize the variation in sprinkler discharge along laterals with fluctuating pressure caused by
undulating terrains.
Diffuse-jet nozzles are designed so that droplets are formed at a lower pressure than with
other impact nozzles. This is accomplished by using noncircular-shaped nozzle openings or
turbulence inducer at the orifice to spread (diffuse) the jet as it leaves the nozzle. Diffuse-nozzles
do not wet as large an area as do constant-diameter and constant discharge nozzles.
ii) Gear-Driven Sprinklers: Some rotating sprinklers are driven by a small water
turbine located in the base of the head sprinkler. These sprinklers are called gear-driven
sprinklers because the high rotational speed of the turbine is reduced through a series of gears.
Like impact sprinklers, gear-driven sprinklers have one or more jets that rotate around the
vertical axis of the sprinkler. Unlike the start and stop rotation of impact sprinklers, gear-driven
sprinklers rotate smoothly without the splash that occurs each time the arm of an impact
sprinkler strikes the jet.
13
iii) Reaction Drive Sprinklers: Small reaction drive sprinklers are normally rotated by the
torque produced by the reaction of water leaving the sprinkler. These sprinklers usually do not
wet as large an area as do impact or gear-driven sprinklers and usually operate at much lower
pressures (70 to 210 kPa or 10 to 30 psi). Some pertinent data about different sprinklers is
presented in Table 8.1 (Keller and Bliesner, 1990).
b) Fixed-Head Sprinklers
Fixed-head sprinklers depend on smooth and grooved cones, deflector plates, and slots to
produce full-or nearly full-circle sprays or several small holes that spray around the
circumference of the sprinkler. An example of a fixed-head spray-type sprinkler is a multi
streamlet-type fixed-head sprinkler. Many fixed-nozzle sprinklers that produce small droplets
and that operate at low pressures are currently available for center-pivot and linear-move
sprinkle systems.
Risers
Riser pipe connects the rotating or fixed sprinkler head to the lateral pipe. It may be a
fixed length of pipe depending upon the height of crop to be irrigated, or it may be collapsible
pipe. Pipes from 10 to 75 mm in diameter are usually used. A minimum length of about 75 mm
on small sprinklers and 1 m on large sprinklers generally provide the best flow pattern through
the sprinkler nozzle. For over-crop sprinkling in orchards or other tall crops, the riser length may
extend from 4 to 5m. Where high risers are used, quick-couplings are often provided to enable
uncoupling during the moving of the lateral. A tripod is sometimes used to keep the uncoupled
riser in upright position.
Laterals
The lateral allows transport of water from the main pipeline or from a sub-main to the
sprinklers through the risers. The lateral pipes usually available in portable lengths of 5, 6 or 12
m are regularly spaced on the mains or sub-mains. Buried permanent laterals are however, used
for orchards, tree nurseries, and for other special sites. Quick-coupled aluminum pipe is best for
most portable laterals. A rubber gasket in the female portion of these couplings has a U-shape.
The water pressure forces the outside of the 'U' to form a water tight seal. When the water is
turned off, the seal is broken and water drains from the pipe, making it easier to uncouple and
move the lateral line to next irrigation position.
Mains and Sub-mains
Main pipelines and sub-mains convey water from the pumping unit to the laterals. In
small systems, laterals may be directly connected to the main line without any sub-mains. Main
lines may be portable or permanent. Permanent mains are used where crops require full season
irrigation. Asbesttos-Cement (Ae), polyvinyl chloride plastic (PVC) and wrapped steel pipes are
used for buried pipelines. Although aluminum is generally used for portable systems, when
wrapped it is also used for buried pipelines. Portable mains are more economical when a sprinkle
14
system is to be used on a number of fields. Steel pipes are used for most permanent main lines
and for center-pivot laterals.
With water being supplied by the pipelines at desired pressures to each lateral and
sprinkler, the pipes must be strong enough to withstand expected operating and surge pressures
in case of water hammering. Buried pipes must resist overburden and dynamic surface loads,
while portable pipes must be tight and durable. PVC pipes are resistant to rust and corrosion
while steel pipes are immune to both of these. Exposure to saline or acid conditions can corrode
aluminum pipes.
Economical pipe selection: The selection of pipeline diameter is critical to the cost of
any sprinkler system. Pipeline cost consists of annual fixed cost of pipes (main, sub-mains and
laterals) and the annual operating cost of pumping water through them. The total cost further
depends on annual operation hours, costs of fuel, anticipated life and friction characteristics of
pipe material and annual interest rate. Head loss due to the friction is a function of pipe diameter.
If a small diameter pipe is used, operating cost will be more due to more friction loss but fixed
cost will be low. As the pipe diameter is increased, the operating cost will decrease but the fixed
cost will increase. The optimum pipe size is the one that minimizes the sum of the fixed- and
operating costs as shown in Fig. 1.10.
Figure 1.10 Relationship between pipe diameter, capital, operating and total costs.
15
Pumps
A pump is required for any sprinkle system. Centrifugal or deep-well turbine pumps are
commonly used. A pump may be stationary or it may be mounted on a movable assembly.
Pumps may be powered by electric motors or internal combustion engines using diesel, gasoline
or natural gas.
Valves and Accessories
Different types of valves, joints, couplings, pressure regulators and pressure gauges are used in a
properly designed sprinkle system. Main types of valves include pressure relief valves, check
valves, foot valves, drain valves, Tee and Y valves, etc. Pressure relief valves are used to relieve
excessive pressure surges. Check valves are used on the discharge side of the pump so that if the
pump is shut off it should maintain water in the pipeline above the pump. Foot-valve on the
bottom of the suction pipe maintains water in the pump when it is not in operation and thus keep
the pump in primed condition.
PERTINENT DATA OF SPRINKLER IRRIGATION SYSTEM
Tables 1 and 2 shows water saving by using sprinkler as compared to surface irrigation
method in Pakistan and India respectively. It is apparent from these tables that more than 50%
water may be saved by using sprinkler as compared to surface irrigation. Tables 3 and 4 show
cost of the piped irrigation system. It is apparent from Table 4 that almost half of the cost is for
the pipe including mains, sub-mains and laterals.
Tables 5 and 6 present results of evaluation of a sprinkler system at different pressures and the
same results are plotted in Figures 1.11 and 1.12. It is apparent from this data that uniformity of
the system depends on operating pressure of the system and fuel consumption increases when the
system is operated at higher pressure. A center-pivot system in maize crop and drip system on
different crops and vegetable are shown in Figures given at the end.
Table 1 Saving in irrigation water using sprinkler irrigation in Pakistan
Crop Water applied (cm) Water saving
(%) Furrow/Basin Sprinkler
Turnip 27.18 (F)* 9.4 65.3
Pea 19.00 (F) 7.9 58.5
Radish 33.00 (F) 11.7 64.5
Maize 25.5 (B)* 13.8 46.0
Average 58.60
* Furrow/Basin
16
Table 2 Saving in Irrigation Water Using Sprinkler Irrigation in India
Table 3. Comparative Costs of Piped Irrigation Systems
Piped Surface Method Sprinkler Conventional
hand-move
Micro-irrigation solid
installation
Area (ha) 1 1-2 2-3 1 1-2 2-3 1 1-2 2-3
Installation
cost
(US$/ha)
1700 1600 1400 2800 2700 2100 3950 3300 3000
Annual
maintenance
cost
(US$/ha)
85 80 70 140 135 105 200 165 150
Note: Average 1997 prices in Europe.
Source: (Phocaides, 2000)
Crop Water Applied (cm) Water
Saving (%)
Surface Sprinkler
Bajra 17.78 7.82 56
Jowar 25.40 11.27 56
Cotton 40.64 29.05 29
Wheat 33.02 14.52 56
Barley 17.78 7.82 56
Gram 17.78 7.82 56
Potato 60.00 30.00 50
Average 51.30
17
Table 4. Cost breakdown for piped irrigation systems
Component Parts Sophisticated
Installation
Simple
Installation
Control station
Mains, submains and manifolds
Fittings and other accessories
Laterals (pipes and emitters)
> 23%
10%
22%
45%
13%
21%
24%
42%
Source: (Phocaides, 2000)
Some Sprinkler Manufacturer Web Sites
www.senninger.com
www.rainbird.com
www.toro.com
www.orbitonline.com
www.rainforrent.com
www.valmont.com
18
19
20
21
1
Maize on Sprinkler System Maize on Sprinkler System
Grapes on Drip irrigation
22
Sugarcane on Drip
Chilies on Drip
19
Unit 2
IRRIGATION SCHEDULING
I. Irrigation Depth
Z
a x
W
100
MAD
d (2.1)
Where, d
x
is the maximum net depth of water to apply per irrigation; MAD is
management allowed deficit (usually 40% to 60%); W
a
is the water holding capacity, a
function of soil texture and structure, equal to FC - WP (field capacity minus wilting
point); and Z is the root depth.
For most agricultural soils, field capacity (FC) is attained about 1 to 3 days after a
complete irrigation.
The d
x
value is the same as "allowable depletion." Actual depth applied may be less if
irrigation frequency is higher than needed during peak use period.
MAD can also serve as a safety factor because many values (soil data, crop data, weather
data, etc.) are not precisely known.
Assume that crop yield and crop ET begins to decrease below maximum potential levels
when actual soil water is below MAD (for more than one day).
Water holding capacity for agricultural soils is usually between 10% and 20% by volume.
W
a
is sometimes called "TAW (total available water), "WHC" (water holding capacity),
"AWHC" (available water holding capacity).
Note that it may be more appropriate to base net irrigation depth calculations on soil
water tension rather than soil water content, also taking into account the crop type - this is
a common criteria for scheduling irrigations through the use of tensiometers.
II. Irrigation Interval
The maximum irrigation frequency is:
d
x
x
U
d
f (2.2)
Where, f
x
is the maximum interval (frequency) in days; and U
d
is the average daily crop water
requirement during the peak-use period.
The range of f
x
values for agricultural crops is usually:
0.25 < f
x
< 80 days (2.3)
20
Then nominal irrigation frequency, f', is the value of f
x
rounded down to the nearest
whole number of days.
But, it can be all right to round up if the values are conservative and if f
x
is near the next
highest integer value.
f' could be fractional if the sprinkler system is automated.
f' can be further reduced to account for non-irrigation days (e.g. Sundays), whereby f < f'
The net application depth per irrigation during the peak use period is d
n
= f'U
d
, which will
be less than or equal to d
x
. Thus, d
n
d
x
, and when d
n
= d
x
, f' becomes f
x
(the maximum
allowable interval during the peak use period).
Calculating d
n
in this way, it is assumed that U
d
persists for f' days, which may result in
an overestimation if f' represents a period spanning many days.
III. Peak Use Period
Irrigation system design is usually for the most demanding conditions:
Figure 2.1 A typical crop coefficient curve
The value of ET during the peak use period depends on the crop type and on the weather.
Thus, the ET can be different from year to year for the same crop type.
Some crops may have peak ET at the beginning of the season due to land preparation
requirements, but these crops are normally irrigated by surface systems.
When a system is to irrigate different crops (in the same or different seasons), the crop
with the highest peak ET should be used to determine system capacity.
21
Consider design probabilities for ET during the peak use period, because peak ET for the
same crop and location will vary from year-to-year due to weather variations.
Consider deficit irrigation, which may be feasible when water is very scarce and or
expensive (relative to the crop value). However, in many cases farmers are not interested
in practicing deficit irrigation.
IV. Leaching Requirement
Leaching may be necessary if annual rains are not enough to flush the root zone or if deep
percolation from irrigation is small (i.e. good application uniformity and or efficiency).
If EC
w
is low, it may not be necessary to consider leaching in the design (system
capacity).
Design equation for leaching:
w e
w
EC EC 5
EC
LR

(2.4)
where LR is the leaching requirement; EC
w
is the EC of the irrigation water (dS/m or
mmho/cm); and EC
e
is the estimated saturation extract EC of the soil root zone for a
given yield reduction value.
Equation 2.4 is taken from FAO Irrigation and Drainage Paper 29.
When LR > 0.1, the leaching ratio increases the depth to apply by 1/(1-LR); otherwise,
LR does not need to be considered in calculating the gross depth to apply per irrigation,
nor in calculating system capacity:
a
n
E
d
d : 0.1 LR (2.5)
a
n
LR)E - (1
d 0.9
d : 0.1 LR (2.6)
E
a
is the application efficiency in fraction.
When LR < 0.0 (a negative value) the irrigation water is too salty, and the crop would
either die or suffer severely.
Standard salinity vs. crop yield relationships (e.g. FAO) are given for electrical
conductivity as saturation extract.
Obtain saturation extract by adding pure water in laboratory until the soil is
saturated, then measure the electrical conductivity.
22
Here are some useful conversions: 1 mmho/cm = 1 dS/m = 550 to 800 mg/l (depending
on chemical makeup, but typically taken as 640 to 690). And, it can usually be assumed
that 1 mg/l 1 ppm, where ppm is by weight (or mass).
V. Leaching Requirement Example
Suppose EC
w
= 2.1 mmhos/cm (2.1 dS/m) and EC
e
for 10% reduction in crop yield is 2.5
dS/m. Then,
20 . 0
2.1 - 5(2.5)
2.1
EC 5EC
EC
LR
w E
w

(2.7)
Thus, LR > 0.1. And, assuming no loss of water due to application non-uniformity, the gross
application depth is related to the net depth as follows:
LR) - (1
d
) ( d d
n
n
d LR (2.8)
n
n
d 25 . 1
0.20) - (1
d
d (2.9)
Gross Application Depth
0.1 LR for ,
E
d
d
pa
n
(2.10)
Where, E
pa
is the design application efficiency (decimal; Eq. 6.9). And,
0.1 LR for ,
LR)E - (1
0.9d
d
pa
n
(2.11)
The gross application depth is the total equivalent depth of water which must be delivered
to the field to replace (all or part of) the soil moisture deficit in the root zone of the soil,
plus any seepage, evaporation, spray drift, runoff and deep percolation losses.
The above equations for d presume that the first 10% of the leaching requirement will be
satisfied by the E
pa
(deep percolation losses due to application variability). This presumes
that areas which are under-irrigated during an irrigation will also be over-irrigated in the
following irrigation, or that sufficient leaching will occur during non-growing season
(winter) months.
23
When the LR value is small (EC
w
< EC
e
), leaching may be accomplished both before and
after the peak ET period, and the first equation (for LR < 0.1) can be used for design and
sizing of system components. This will reduce the required pipe and pump sizes because
the "extra" system capacity during the non-peak ET periods is used to provide water for
leaching.
System Capacity
Application volume can be expressed as either Qt or Ad, where Q is flow rate, t is
time, A is irrigated area and d is gross application depth.
Both terms are in units of volume.
Thus, the system capacity is defined as (Eq. 5.4):
fT
Ad
K Q
s
(2.12)
Where,
Q
s
= system capacity, l/s (gpm)
T = hours of system operation per day (obviously, T< 24; also, t = fT)
K = coefficient for conversion of units (see below)
d = gross application depth, mm (in.)
f = time to complete one irrigation (days); equal to f' minus the days off
A = net irrigated area supplied by the discharge Q
s
, ha (acres)
Value of K:
Metric: for d in mm, A in ha, and Q
s
in lps: K = 2.78
English: for d in inches, A in acres, and Q
s
in gpm: K = 453
24
Notes about system capacity:
Eq. 2.12 (Eq 5.4 book) is normally used for periodic-move and linear-move sprinkler
systems.
The equation can also be used for center pivots if f is decimal days to complete one
revolution and d is the gross application depth per revolution.
For center pivot and fixed systems, irrigations can be light and frequent (d
applied
< d):
soil water is maintained somewhat below field capacity at all times (assuming no
leaching requirement), and there is very little deep percolation loss.
Also, there is a margin of safety in the event that the pump fails (or the system is
temporarily out of operation for whatever reason) just when MAD is reached (time to
irrigate), because the soil water deficit is never allowed to reach MAD.
However, light and frequent irrigations are associated with higher evaporative losses,
and probably higher ET too (due to more optimal soil moisture conditions).
Frequent irrigations correspond to a higher basal crop coefficient Kcb (due to more
favorable soil moisture conditions), and a higher wet soil surface evaporation
coefficient, Ks (due to more frequent wetting).
When a solid-set (fixed) system is used for frost control, all sprinklers must operate
simultaneously and the value of Q
s
is equal to the number of sprinklers multiplied by
q
a
. This tends to give a higher Q
s
than that calculated from Eq. 5.4.
Example: Determine the required system capacity for a sprinkler system with following data:
Area, A = 140 acres root depth, Z = 2 ft
T = 22 h/day U
d
= 0.21 in 1 day
MAD = 50 % C
u
= 79%
W
a
= 1.0 in/ft
1 2 1 0.5
100
Z MAD.W
d
a
x

Irrigation Interval, 4) f take, ve conservati be (to 4.76
0.21
1
U
d
f
d
x

gpm 720.7
22 4
1 140 453
f.T
Ad
Q


K
s
If f = 5 days gpm 577 Q
s

25
Unit 3
WATER DISTRIBUTION UNIFORMITY AND APPLICATION EFFICIENCY
Water Distribution Pattern from a Stationary Sprinkler
Sprinkler head is the most important component of a sprinkle
system because it distributes water over the land. Efficiency and
effectiveness of any sprinkle system depend how uniformly water is
sprayed from the sprinkler.
Water application rate and distribution pattern of a sprinkler
are a function of i) nozzle size and its angle, ii) nozzle pressure, iii)
diameter of throw, iv) wind, and v) sprinkler spacing on the lateral and
spacing of the lateral on the main. Water distribution characteristics of sprinkler heads are typical
and change with nozzle- size, shape, angle, and operating pressure. Water application beneath a
sprinkler varies with distance from the head. The pattern of this variation, called distribution
pattern, is usually consistent for a given pressure, nozzle geometry and wind. Pressure has a
significant effect on the distribution pattern. Typical distribution patterns beneath a stationary
impact sprinkler with fixed nozzle geometry for different operating pressures are given in Figure
3.1. If pressure is low, larger water drops fall near the sprinkler forming a 'donut-shaped'
distribution pattern. Under the normal operating pressure as recommended by the manufacturer a
triangular or elliptical shaped distribution pattern is obtained in which the depth of water
application is found maximum near the head and decreases toward the outer edge of the pattern.
Extremely high pressure produces too many fine water drops that fall finer water drops that fall
near the sprinkler distorting the desired distribution pattern (see Fig. 3.2).
Classification of Sprinklers and Applicability
Table 3.1 Classification of sprinkler heads based on operating pressure
Agriculture
sprinklers
(two nozzle)
Nozzle size mm Operating
pressure (bars)
Flow rate
(m
3
/h)
Diameter
coverage (m)
Low pressure
Medium pressure
High pressure
3.04-4.5x2.5-3.5
4.0-6.0x2.5-4.2
12.0-25.0x5.0-8.0
1.5-2.5
2.5-3.5
4.0-9.0
0.3-1.5
1.5-3.0
5.0-45.0
12-21
24.35
60-80
26
`
Figure 3.1 Water distribution pattern from a stationary rotating sprinkler head
(Source: USDA, SCS Handbook, 1960).
Precipitation Profiles
Typical examples of low, correct, and high sprinkler pressures (see Fig 5.5 below).
Figure 3.2 Water Profiles at different Pressures.
27
Analysis of Water Application under a Stationary Sprinkler
Agricultural sprinklers typically have flow rates from 4 to 45 lpm (1 to 12 gpm), at nozzle
pressures of 135 to 700 kPa (20 to 100 psi).
"Gun" sprinklers may have flow rates up to 2,000 lpm (500 gpm; 33 lps) or more, at
pressures up to 750 kPa (110 psi) or more.
Sprinklers with higher manufacturer design pressures tend to have larger wetted
diameters.
But, deviations from manufacturer's recommended pressure may have the opposite effect
(increase in pressure, decrease in diameter), and uniformity will probably be
compromised.
Sprinklers are usually made of plastic, brass, and or steel.
Low pressure nozzles save pumping costs, but tend to have large drop sizes and high
application rates.
Medium pressure sprinklers (210 - 410 kPa, or 30 to 60 psi) tend to have the best
application uniformity.
Medium pressure sprinklers also tend to have the lowest minimum application rates.
High pressure sprinklers have high pumping costs, but when used in periodic-move
systems can cover a large area at each set.
High pressure sprinklers have high application rates.
Rotating sprinklers have lower application rates because the water is only wetting a
"sector" (not a full circle) at any given instance.
For the same pressure and discharge, rotating sprinklers have larger wetted diameters.
Impact sprinklers always rotate; the "impact" action on the stream of water is designed to
provide acceptable uniformity, given that much of the water would otherwise fall far from
the sprinkler (the arm breaks up part of the stream).
The precipitation profile (and uniformity) is a function of many factors:
1. nozzle pressure
2. nozzle shape & size
3. sprinkler head design
4. presence of straightening vanes
5. sprinkler rotation speed
6. trajectory angle
7. riser height
8. wind
Straightening vanes can be used to compensate for consistently windy conditions.
28
Overlapping sprinkler profiles
Figure 3.3 Water distribution patterns for individual and overlapped sprinklers
(Source: USDA, SCS Handbook, 1960).
Set Sprinkler Uniformity & Efficiency
Sprinkler Irrigation Efficiency
1. Application uniformity
2. Losses (deep percolation, evaporation, runoff, wind drift, etc.)
29
It is not enough to have uniform application if the average depth is not enough to refill
the root zone to field capacity.
Similarly, it is not enough to have a correct average application depth if the uniformity is
poor.
Consider the following examples:
Fig. 3.4 Uniform, but average depth applied exceeds the soil water deficit
(too much deep percolation)
Fig. 3.5 Average depth is correct, but application is highly non-uniform, with
under-irrigation and DP
We can design a sprinkler system that is capable of providing good application
uniformity, but depth of application is a function of the set time (in periodic-move
systems) or "on time" (in fixed systems).
Thus, uniformity is mainly a function of design and subsequent system maintenance, but
application depth is a function of management.
30
Quantitative Measures of Uniformity
Traditional measurements of sprinkler irrigation uniformity only account for the aerial
distribution of water.
These measurements of uniformity do not account for redistribution of water within the
soil profile, redistribution due to foliar interception of water drops, and surface runoff
after the drops hit the ground.
Following are two commonly applied indicators of aerial water distribution:
Distribution uniformity, DU (Eq. 6.1):
|
|
.
|

\
|
=
depth avg
quarter low of depth Avg
100 DU (3.1)
The average of the low quarter is obtained by measuring application from a catch-can
test, mathematically overlapping the data (if necessary), ranking the values by magnitude,
and taking the average of the values from the low of all values.
For example, if there are 60 values, the low quarter would consist of the 15 values with
the lowest "catches".
Christiansen Coefficient of Uniformity, CU (Eq. 6.2):
|
|
|
|
.
|

\
|

=
=
n
1 j
n
1 j
j
) (z abs
0 . 1 100 CU
j
z
m
(3.2)
Where, z are the individual catch-can values (volumes or depths); n is the number of
observations; and m is the average of all catch volumes.
Note that CU can be negative if the distribution is very poor.
There are other, equivalent ways to write the equation.
These two measures of uniformity (CU & DU) date back to the time of slide rules (more
than 50 years ago; no electronic calculators), and are designed with computational ease in
mind.
More complex statistical analyses can be performed, but these values have remained
useful in design and evaluation of sprinkler systems. For CU > 70% the data usually
conform to a normal distribution, symmetrical about the mean value. Then,
31
|
|
.
|

\
|
~
depth avg
half low of depth avg
100 CU (3.3)
Another way to define CU is through the standard deviation of the values,
|
|
.
|

\
|
=

2
m
0 . 1 100 CU (3.4)
Where, o is the standard deviation of all values, and a normal distribution is assumed (as
previously)
Note that CU = 100% for o = 0
The above equation assumes a normal distribution of the depth values, whereby:

= 2/ n m - z (3.5)
By the way, the ratio o/m is known in statistics as the coefficient of variation.
Following is the approximate relationship between CU and DU:
CU ~ 100 - 0.63(100 - DU) (3.6)
Or
DU ~ 100 -1.59(100 - CU) (3.7)
These equations are used in evaluations of sprinkler systems for both design and operation.
Typically, 85 to 90% is the practical upper limit on DU for set systems DU > 65% and CU >
78% is considered to be the minimum acceptable performance level for an economic system
design; so, you would not normally design a system for a CU < 78%, unless the objective is
simply to "get rid of water or effluent" (which is sometimes the case). For shallow-rooted, high
value crops, you may want to use DU > 76% and CU > 85%.
32
Alternate Sets (Periodic-Move Systems)
The effective uniformity (over multiple irrigations) increases if "alternate sets" are used
for periodic-move systems (1/2 S
l
).
This is usually practiced by placing laterals halfway between the positions from the
previous irrigation, alternating each time.
The relationship is:
CU
a
~ 10\CU (3.8)
DU
a
~
10
\
DU (3.9)
Use of alternate sets is a good management practice for periodic-move systems.
The use of alternate sets approaches S
l
of zero, which simulates a continuous-move
system.
Uniformity Problems
From the various causes of non-uniform sprinkler application, some tend to cancel out
with time (multiple irrigations) and others tend to concentrate (get worse).
In other words, the "composite" CU for two or more irrigations may be (but not
necessarily) greater than the CU for a single irrigation.
1. Factors that tend to Cancel Out
Variations in sprinkler rotation speed
Variations in sprinkler discharge due to wear
Variations in riser angle (especially with hand-move systems)
Variations in lateral set time
2. Factors that may both Cancel Out and Concentrate
Non-uniform aerial distribution of water between sprinklers
3. Factors that tend to Concentrate
Variations in sprinkler discharge due to elevation and head loss
Surface ponding and runoff
Edge effects at field boundaries
33
Water distribution uniformity effectively increases after water infiltrates into the soil because
of root-zone redistribution from wetter regions to drier regions. This effect is usually greater
in "tight" clay soils than in sandy soils. Thus, the actual application uniformity in the root
zone tends to be a little better than the aerial distribution from the sprinklers, at least in the
absence of significant runoff.
System Uniformity
The uniformity is usually less when the entire sprinkler system is considered, because
there tends to be greater pressure variation in the system than at any given lateral
position.

+ ~ ) / P 1 (
2
1
CU CU system
n a
P (3.10)

+ ~ ) / P 3 1 (
4
1
DU DU system
n a
P (3.11)
Where, P
n
is the minimum sprinkler pressure in the whole field; and P
a
is the average
sprinkler pressure in the entire system, over the field area.
These equations can be used in design and evaluation. Note that when P
n
= P
a
(no
pressure variation) the system CU equals the CU. If pressure regulators are used at each
sprinkler, the system CU is approximately equal to 0.95CU (same for DU). If flexible
orifice nozzles are used, calculate system CU as 0.90CU (same for DU)
The P
a
for a system can often be estimated as a weighted average of P
n
& P
x
:
3
P 2P
P
x n
a
+
= (3.12)
Where, P
x
is the maximum nozzle pressure in the system
DU = f(P,P,S,d
n
,WDP,WS) (3.31)
E
a
= f(P,P,S,d
n
,WDP,WS,I.i
s
,t
i
,SWD) (3.32)
34
Fig. 3.6 Due to parabolic loss vs. flow rate relation, the average is closer to P
n
Computer Software and Standards
There is a computer program called "Catch3D" that performs uniformity calculations on
sprinkler catch-can data and can show the results graphically.
Jack Keller and John Merriam (1978) published a handbook on the evaluation of
irrigation systems, and this includes simple procedures for evaluating the performance of
sprinkler systems.
The ASAE S436 (Sep 92) is a detailed standard for determining the application
uniformity under center pivot (not a set sprinkler system, but a continuous move system).
ASAE S398.1 provides a description of various types of information that can be collected
during an evaluation of a set sprinkler system.
General Sprinkle Application Efficiency
The following material leads up to the development of a general sprinkle application
efficiency term (Eq. 6.9) as follows:
Design Efficiency:
E
pa
= DE
pa
R
e
O
e
(3.13)
Where, DE
pa
is the distribution efficiency (%); R
e
is the fraction of applied water that
reaches the soil surface; and O
e
is the fraction of water that does not leak from the system
pipes.
35
The design efficiency, E
pa
, is used to determine gross application depth (for design
purposes), given the net application depth.
In most designs, it is not possible to do a catch-can test and data analysis you have to
install the system in the field first; thus, use the "design efficiency".
The subscript "pa" represents the "percent area" of the field that is adequately irrigated
(to d
n
, or greater). For example, E
80
and DE
80
are the application and distribution
efficiencies when 80% of the field is adequately irrigated.
Simulate different lateral spacings by "overlapping" catch-can data in the direction of lateral
movement (overlapping along the lateral is automatically included in the catch-can data, unless
it's just one sprinkler).
Figure 3.7 An example to calculate data to measure irrigation uniformity for sprinkler
irrigation system using single lateral
Field Evaluation of Sprinklers
Catch-can tests are typically conducted to evaluate the uniformities of installed sprinkler
systems and manufacturer's products.
Catch-can data is often overlapped for various sprinkler and lateral spacings (S
e
& S
l
) to
evaluate uniformities for design and management purposes.
A computer program developed at USU does the overlapping: CATCH3D.
Note that catch-can tests represent a specific wind and pressure situation and must be
repeated to obtain information for other pressures or wind conditions.
36
S
e
S
l
Typical catch-can spacings are 2 or 3 m on a square grid, or 1 to 2 m spacings along one
or more "radial legs", with the sprinkler in the center.
Set up catch-cans with half spacing from sprinklers (in both axes) to facilitate overlap
calculations.
See Merriam & Keller (1978); also see ASAE S398.1 and ASAE S436
Choosing a Suitable Sprinkler Heads
The system designer doesn't "design" a sprinkler, but "selects" a sprinkler.
There are hundreds of sprinkler designs and variations from several manufacturers, and
new sprinklers appear in the market quite often.
The system designer must choose between different nozzle sizes and nozzle designs for a
given sprinkler head design.
The objective is to combine sprinkler selection with S
e
and S
l
to provide acceptable application
uniformity, acceptable operating costs, and acceptable hardware & installation costs.
Manufacturers provide recommended spacings and pressures.
There are special sprinklers designed for use in frost control.
General Spacing Recommendations
Sprinkler spacing is usually rectangular or triangular.
Triangular spacing is more common under fixed-
system sprinklers.
Sprinkler spacings based on average (moderate) wind
speeds:
1. Rectangular spacing is 40% (S
e
) by 67% (S
l
) of the
effective diameter
2. Square spacing is 50% of the effective diameter
3. Equilateral triangle spacing is 62% of the effective
diameter [lateral spacing is 0.62 cos(60/2) = 0.54,
or 54% of the effective diameter,
See Fig. 3.8 about profiles and spacings.
See the following figure about spacings and overlaps.
37
Rectangular: 40% x 67%
Square: 50% x 50%
Triangular: 62% x 54%
Figure 3.8 Overlapping patterns for different spacing of laterals
38
Windy Conditions
When winds are consistently recurring at some specific hour, the system can be shut
down during this period.
For center pivots, rotation should not be a multiple of 24 hours, even if there is no
appreciable wind (evaporation during day, much less at night).
If winds consistently occur, special straightening vanes can be used upstream of the
sprinkler nozzles to reduce turbulence; wind is responsible for breaking up the stream, so
under calm conditions the uniformity could decrease.
For periodic-move systems, laterals should be moved in same direction as prevailing
winds to achieve greater uniformity (because S
e
< S
l
).
Laterals should also move in the direction of wind to mitigate problems of salt
accumulating on plant leaves.
Wind can be a major factor on the application uniformity on soils with low infiltration
rates (i.e. low application rates and small drop sizes).
In windy areas with periodic-move sprinkler systems, the use of offset laterals (S
l
) may
significantly increase application uniformity.
Alternating the time of day of lateral operation in each place in the field may also
improve uniformity under windy conditions.
Occasionally, wind can help to increase uniformity, as the randomness of wind
turbulence and gusts helps to smooth out the precipitation profile.
Wind effects on the diameter of throw:
0-3 mph wind: reduce manufacturer's listed diameter of throw by 10% for an effective
value (i.e. the diameter where the application of water is significant)
over 3 mph wind: reduce manufacturer's listed diameter of throw by an additional 2.5% for
every 1 mph above 3 mph (5.6% for every 1 m/s over 1.34 m/s)
In equation form:
For 0-3 mph (0-1.34 m/s):
diam = 0.9diam
manuf
(3.14)
For > 3 mph (> 1.34 m/s):
diam = diam
manuf
[0.9 - 0.025 (wind
mph
- 3)] (3.15)
or,
diam = diam
manuf
[0.9 - 0.056(wind
m/s
-1.34)] (3.16)
39
Example: a manufacturer gives an 80-ft diameter of throw for a certain sprinkler and
operating pressure. For a 5 mph wind, what is the effective diameter?
80 ft - (0.10)(0.80) = 72 ft (3.17)
72 ft - (5 mph - 3 mph)(0.025)(80 ft) = 68 ft (3.18)
or,
diam = 80(0.9-0.025(5-3))=68 ft (3.19)
Pressure-Discharge Relationship
Equation 5.1:
q = K
d
\P
where, q is the sprinkler flow rate; K
d
is an empirical coefficient; and P is the nozzle pressure
The above equation is for a simple round orifice nozzle. It can be derived from Bernoullis
equation like this:
2
2 2
2 2
P
gA
q
g
V
= =

(3.20)
q P K
g
P
d
= =

2
2gA
(3.21)
Where, the elevations are the same (z
1
= z
2
) and the conversion through the nozzle is
assumed to be all pressure to all velocity.
P can be replaced by H (head), but the value of K
d
will be different
Eq. 5.1 is accurate within a certain range of pressures
See Table 5.2 for P, q, and K
d
relationships
K
d
can be separated into an orifice coefficient, K
o
, and nozzle bore area, A:

P A K q
o
= (3.22)
Where by,
/ 2 =
o
K (3.23)
Where, the value of K
o
is fairly consistent across nozzle sizes for a specific model and
manufacturer.
Application Rates
The application rate should be selected to match conditions of the soil and crop. The
average application rate of a sprinkler can be calculated as:
40
A
q
K I = (3.24)
Where,
I = application rate (mm/h, in/h)
A = wetted area of sprinkler (m
2
, ft
2
)
K = unit constant (K equal to 60 for I mm/h, q l/min and A m
2
) or equal to 96.3 for
I in/h, q gpm, and A ft
2
A single stationary sprinkler applies water on a circular area. Water depth observe maximum
near the head and decreases outward as shown in Figure 3.1. Thus to obtain a reasonable uniform
water application, the wetted circular areas of adjacent sprinklers are overlapped as shown in
Figures 3.2 and 3.3 above. The resulting accumulated depth is almost uniform.
The average application rate along a lateral can be calculated as:
LS
Q
K I
l
=
(3.25)
Where, Q
l
is the discharge of the lateral, L is the length of lateral, and S is the spacing between
adjacent lateral, I and K are as previously defined.
Allowable Application Rate
Sprinkle systems are normally designed for no surface runoff. Thus, application rate of a
sprinkle system is designed to apply less water than the infiltration rate of the soil. Complication
arises as the soil infiltration rate decreases with time of application as illustrated by a curve in
Figure 3.9.
Figure 3.9 Relationship between infiltration rate of a soil and constant application
rates.
I
n
f
i
l
t
r
a
t
i
o
n

R
a
t
e

(
L
T
-
1
)
Curve a
b
c
Time (T)
41
Curve in the above figure shows for an unlimited amount of water available at the soil
surface that the infiltration rate is initially higher than the application rate (line b). Thus, no
runoff occurs in the beginning of irrigation. As the infiltration rate decreases with time, runoff
may occur if irrigation is continued at rate b. No runoff will occur if the application rate is
everywhere less than curve corresponding to line c.
EVALUATION OF SPRINKLE SYSTEMS
Irrigation efficiency and uniformity of water application are often used to describe
effectiveness of any sprinkle system. Keller and Bliesner (1988, page 86) have identified the
following factors which affect water application efficiency of sprinkle systems.
- Variation of individual sprinkler discharge throughout the lateral lines. This variation
can be held to a minimum by proper pipe network design or by employing pressure or
flow-control devices at each sprinkler or sprinkler nozzle.
- Variation in water distribution within the sprinkler-spacing area. This variation is
caused primarily by wind. It can be partly overcome for set sprinkler systems by close
spacing of the sprinklers. In addition to the variation caused by wind, there is a
variability in the distribution pattern of individual sprinklers. The extent of this
variability depends on sprinkler design, operating pressure, and sprinkler rotation.
- Loss of water by direct evaporation from the spray. Losses increase as temperature
and wind velocities increase, and as drop size and application rate decrease.
- Evaporation from the soil surface before the water is used by the plants. This loss will
be proportionately smaller as greater depths of water are applied.
Uniformity
The performance of any irrigation system is judged by the evenness or uniformity of its
water application. Only an ideal system will apply water with 100 percent uniformity. For
sprinkler systems catch-cans normally placed in a grid pattern are used to estimate the uniformity
with volumes of water collected in the cans measured in graduated cylinders. Christiansen's
coefficient of uniformity, after used to measure the uniformity of application in sprinkle
irrigation, is calculated (Christiansen, 1942) from:
]
x
- [1.0 100 CU
MN
L
= (3.26)
Where
CU = Christiansen uniformity coefficient in percent,
x = deviation of an individual observation from the mean,
M = mean of all observations, and
N = number of observations
42
The higher the coefficient of uniformity, better the water application of the system will be.
Usually, a system is considered satisfactory if it has a coefficient of uniformity higher than 70
percent.
In the above equation, it is assumed that each catch-can represents an equal sample area.
This is the case in sprinkle systems other than center pivots. For the center pivot system if cans
are set out radially from the pivot point, this would only be true if the can spacing is varied
inversely with the distance from the pivot. Heermann and Hein (1968) suggested a method of
calculating the coefficient of uniformity for the center-pivot system for radially located cans.
They determined the sample area represented by each can and then by considering the volume of
water applied to this area rather than the depth applied at a point, a formula very similar to above
equation was used to calculate the coefficient of uniformity. If the distance between the pivot and
first can is one-half of the distance between the succeeding cans, then the relationship proposed
by Heermann and Hein reduces to the following (Ring and Heermann 1978):
s
s s
R D
D D R
L
L
=
)
1 ( 100 CU
w
(3.27)
Where,
CU
w
= weighted Christiansen coefficient of uniformity,
D
s
= depth at any point
D = weighted mean depth, given as
D = (R
s
D
s
)/ R
s
R
s
= weighting factor equal to the distance from the pivot, given as
R
s
= 0.5S + (i - 1) S where:
i = subscript referring to the ith can from the pivot
S = spacing between cans
Merriam and Keller (1978) recommended another indicator called Distribution
Uniformity (DU) to measure uniformity. The Distribution uniformity indicates the uniformity
throughout the field and is given by:
Distribution Uniformity = Average low-quarter depth of water applied x 100 (3.28)
Average depth of water applied
The average low-quarter depth is the average of the lowest one-quarter of the data set (assuming
each value represents an equal area) and it does not refer to the lowest one fourth of the field.
The CU and DU are approximately related as (Keller and Bliesner, 1988, p. 87):
CU = 100 - 0.63 (100 - DU) (3.29)
or
DU = 100 - 1.59 (100 - CU) (3.30)
43
Irrigation scheduling and water application practices can be combined to determine the water
application uniformity and efficiency of an irrigation system as shown in the equations below.
DU = f(P,P,S,d
n
,WDP,WS) (3.31)
E
a
= f(P,P,S,d
n
,WDP,WS,I.i
s
,t
i
,SWD) (3.32)
Where, P = pressure available at the sprinkler; P = variation of the pressure along the lateral; S
= spacings of the sprinklers along and between the laterals; d
n
= nozzle diameter; WDP = water
distribution pattern of the sprinkler; WS = wind speed and direction; I= intake characteristics of
the soil; i
s
= application rate of the sprinkler; t
i
= duration of the irrigation even; and SWD = soil
water deficit before the irrigation event.
Application Efficiency
Effectiveness of a given sprinkle system can be determined from how much of the
applied water is stored in the crop root zone. Many definitions of application efficiency are found
in the literature therefore the readers are cautioned to use this term carefully. Sometime less
water is applied than the required soil moisture deficit (SMD) in order to efficiently utilize
rainfall. This practice is called under or deficit irrigation. The application efficiency is usually
defined as:
Ea = Average depth added to the root zone storage x 100 (3.33)
Average depth applied to the field
Data Collection: Data for evaluation of a sprinkle system may be collected while actual
system is being normally operated. Cans of uniform shape, size and diameter are placed between
the sprinklers and laterals in a uniform grid except center-pivot system. First row of cans should
be placed at one-half of the grid spacing so that each data point represents equal area. The system
should be operated for at least one to two hours to collect sufficient water in the containers. The
test should be conducted during calm day in order to lessen the drift loss due to wind.
Evaporation during the test period should also be measured. For this purpose, use cans of the
same size like the one used for collecting catch cans.
If actual system is not in operation, data for DU and CU may be collected by operating a
few sprinklers on a lateral. Place the cans between any two sprinklers and note the data as
discussed above. The uniformity coefficients should be calculated by overlapping the data
collected on the other side of the lateral. If symmetrical conditions are assumed on both side of
the lateral, data collected even on one side of the lateral may be used to calculate the uniformity
coefficients by overlapping as illustrated in the following example.
Sample Example: A sprinkle system consisting of single lateral was operated to collect data
to calculate uniformity coefficients as shown in Fig. 3.10. Calculate and compare DU and CU for
different lateral spacings.
44
1.0 1.7 3.5 5.0 X 5.1 3.6 2.0 1.0
0.9 2.0 3.7 6.1 5.2 3.9 1.5 0.8
1.0 2.1 3.6 5.5 5.5 4.0 2.0 1.0
0.8 1.8 3.8 5.2 5.6 3.8 1.5 1.0
Figure 3.10 Data collection to measure Distribution Uniformity and Christiansen Uniformity
Coefficient. Values shown are water depth collected in cans in cm (Grid 3m x 3m).
Calculations are given in Fig. 3.11 for 9 m x 15 m spacings using the above data.
5.1 3.6 2.0 1.0 0.0
0.0 1.0 1.7 3.5 5.0
5.1 4.6 3.7 4.5 5.0
(0.34)* (0.16) (1.06) (0.26) (0.24)
5.2 3.9 1.5 0.8 0.0
0.0 0.9 2.0 3.7 6.1
4.2 4.8 3.5 4.5 6.1
(0.44) (0.04) (1.26) (0.26) (1.34)
5.5 4.0 2.1 1.0 0.0
0.0 1.0 2.0 3.6 5.5
5.5 5.0 4.1 4.6 5.5
(0.74) (0.24) (0.66) (0.16) (0.74)
5.6 3.8 1.5 1.0 0.0
0.0 0.8 1.8 3.8 5.2
5.6 4.6 3.3 4.8 5.2
(0.84) (0.16) (1.46) (0.04) (0.44)
*Figures in parentheses are deviations from the mean.
Figure 3.11 Calculation of Distribution Uniformity and Christiansen Coefficient of
Uniformity by Overlapping water depth collected in cm.
X
1.5m 3 m
3 m
Sprinkler
head
Lateral
Imaginary
lateral
45
There are 20 data points. Mean, M is equal to 4.76 cm. Mean of the lowest one quarter, M
LQ
is
M
LQ
= 3.3 + 3.5 + 3.7 + 4.1 + 4.5 = 19.1/5 = 3.82 cm
DU = 3.82/4.76 x 100 = 80.25%
Sum of deviations = 10.72
CU = (1 10.72/20 x 4.76) x 100 = 88.74%
Similarly data given in Figure 3.11 may be used to calculate the uniformity coefficients
for other lateral spacings. Table 3.1 summarizes the results for DU and CU for different
spacings. The results in the table shows that the uniformity coefficients decrease as lateral
spacing is increased as expected.
Table 3.2 Measured DU and CU values for different lateral spacings.
Spacings (m)
9 x 12 9 x 15 9 x 18
DU, Percent 89.50 80.25 67.25
CU, Percent 93.20 88.74 75.85
46
C A T C H 3 D Sprinkler Overlap Program
Dr. R.G. Allen, Dept. Ag ansd Irrigation Engr.
Utah State University, Logan, Ut 84322-4105
ph (801) 750-2798
The CATCH3D program is an interactive IBM-PC program written in Microsoft Basic. A
compiled version (CATCH3D.EXE) should be on this disk. CATCH3D is designed to simulate
the water application uniformities of rectangular sprinkler patterns (Se x Sl) by overlapping catch
can measurements from either a single sprinkler head test or single lateral line test.
Sprinkler (Se) and lateral (Sl) spacings evaluated must be integer multiples of the catch can
grid. The catch can grid must be square. In other words, if the catch can grid is 2m by 2m, the
following sprinkler spacings could be evaluated by CATCH3D: 4m x 4m, 4m x 8m, 8m x 4m,
16m x 12m, etc.
Catch can data can be stored by CATCH3D in a data file for future use. Therefore, the data
needs to be typed in only once. Sprinkler and lateral spacings to be evaluated can also be stored
in this same file, along with descriptive information concerning the catch can test: number of
rows and columns of catch-cans, location of the sprinkler, measurement units, catch-can size,
duration of test, nozzle discharge, wind speed and direction, and grid spacing.
Program calculations include estimates of the uniformity coefficient (CU), distribution
uniformity, application efficiencies of the low half and low quarter (AELH, AELQ), and catch
can efficiencies. Example data sets on this disk include CATCH3D.DAT, CONE.DAT and
DONUT.DAT. These files can be read into CATCH3D to provide example calculations and
demonstration of program operation for an actual test, and for synthetic conical and donut shaped
patterns.
The resulting overlapped application patterns can be printed on a line printer, or 3-dimensional
graphs of the patterns can be plotted on the screen (color graphics card with 640x200 resolution
required), or plotted on a Hewlett-Packard 7475A table top plotter. Depicted plots are of a
rectangular overlap section with sprinklers located at each corner of the pattern. To print a copy
of the screen generated 3-d plot onto an Epson-type printer, you must have previously had this
disk (with IBM DOS3.1, GRAPHICS.COM, and AUTOEXEC.BAT) in drive A upon turning on
the computer. Then, after the plot is completed on the computer screen, press the SHIFT and
PRT SC keys simultaneously.
To evaluate a catch-can test from a single lateral with multiple sprinklers, set the sprinkler
spacing (Se) to the actual spacing of the tested lateral, and vary the lateral spacing (Sl) to
evaluate effects of various lateral spacings.
The CATCH3D program is copyrighted, 1986, by the author and by Utah State University.
Neither the author nor Utah State University assumes any liability resulting from use of this
program.
47
Unit 4
SPRINKLE IRRIGATION PLANNING AND DESIGN
Hardware Design Process
1. Sprinkler selection
2. Design of the system layout
3. Design of the laterals
4. Design of the mainline
5. Pump and power unit selection
Farm Systems vs. Field Systems
A complete farm sprinkle system can be defined as a system planned exclusively for a given
design area or farm unit on which sprinkling will be the primary method of water application.
Planning for complete systems includes considering specified crops and crop rotations, water
quality, and the soils found in the specified design area.
A farm sprinkle irrigation system includes sprinklers and related hardware; lateral, sub-
main, and main pipelines; pumping plant and boosters; operation control equipment; and other
accessories required for the efficient application of water. Large farm systems are made up of
several field systems. A field system is designed either for use on several fields of a farm unit or
for movement between fields on several farm units. Field systems are planned for stated
conditions, generally for pre-irrigation, for bringing up seedlings, or for use on special crops in a
crop rotation.
`"A poorly designed system that is well managed can often perform better than a well
designed system that is poorly managed"
A Farm system may have many field systems.
Planning considerations should include the possibility of future expansions and extra
capacity.
Permanent buried mainlines should generally be oversized to allow for future needs. It is
much cheaper to put a larger pipe in at the beginning than to install a secondary or larger
line later.
Consider the possibility of future automation.
Consider the needs for land leveling before burying pipes.
How will the system be coordinated over many fields?
What if the cropping patterns change? (tolerance to salinity, tolerance to foliar wetting,
peak ET rate, root depth, need for crop cooling or frost protection, temporal shifting of
peak ET period).
What if energy costs change?
What if labor availability and or cost change?
48
What if the water supply is changed (e.g. from river to groundwater, or from old well to
new well)?
What if new areas will be brought into production?
II. Outline of Sprinkler Design Procedure
1. Make an inventory of resources
Visit the field site personally if at all possible, and talk with the farmer.
Get data on soil, topography, water supply, crops, farm schedules, climate, energy, etc.
Be suspicious of parameter values and check whether they are within reasonable
ranges.
2. Calculate a preliminary value for the maximum net irrigation depth, d
x.
3. Obtain values for peak ET rate, U
d
, and cumulative seasonal ET, U (Table 3.3).
4. Calculate maximum irrigation frequency, f
x
, and nominal frequency, f'.
This step is unnecessary for automated fixed systems and center pivots.
5. Calculate the required system capacity, Q
s
first, calculate gross application depth, d
for center pivots use d
n
/f = U
d
, and T 90% of 24 hrs/day = 21.6
6. Determine the "optimum" (or maximum) water application rate.
a function of soil type and ground slope (Table 5.4).
7. Consider different types of feasible sprinkle systems.
8. For periodic-move and fixed (solid-set) systems:
(a) Determine S
e
, q
a
, nozzle size, and P for optimum application rate (Tables 6.4 to 6.7)
(b) Determine number of sprinklers to operate simultaneously to meet Q
s
(N
s
= Q
s
/q
a
)
(c) Decide upon the best layout of laterals and mainline
(d) Adjust f, d, and/or Q
s
to meet layout conditions
(e) Size the lateral pipes
(f) Calculate the maximum pressure required for individual laterals
9. Calculate the mainline pipe size(s), then select from available sizes.
10. Adjust mainline pipe sizes according to the "economic pipe selection method" (Chapter 10),
and check velocity limits too.
11. Determine extreme operating pressure and discharge conditions.
12. Select the pump and power unit.
13. Draw up system plans and make a list of items with suggestions for operation and
maintenance of the hardware.
III. Summary
Note that MAD is not a precise value; actual precision is less than two significant digits;
this justifies some imprecision in other values (don't try to obtain very precise values for
some parameters when others are only rough estimates).
49
When determining the seasonal water requirements we subtract P
e
from U. However, to
be safe, the value of P
e
must be reliable and consistent from year to year, otherwise a
smaller (or zero) value should be used.
Note that lateral and sprinkler spacings are not infinitely adjustable: they come in
standard dimensions from which designers must choose. The same goes for pipe
diameters and lengths.
However, buried PVC pipes in a fixed system can be cut to any length during installation,
allowing for a great deal of flexibility.
Note that design for peak U
d
may not be appropriate if sprinklers are used only to
germinate seeds (when later irrigations are by a surface method).
Sprinkle System Design
Design of a sprinkle system starts with layout of laterals and main pipelines. Pipe layout
may be simple for small regularly shaped fields to complex for large and odd-shaped areas.
Layout mainly depends on field topography, lateral settings, wind conditions and location of
water source. Design of set move, solid set gun sprinkler and center-pivot are discussed briefly.
For more details, the reader is referred to Keller and Bliesner (1990), Pair et al. (1975), Rolland
(1982), and other sources.
Set-Move and Solid Set Systems
Main steps to design a set-move or solid set sprinklers are layout of main pipe lines and
laterals, determining the number of laterals to be operated per irrigation, number of sprinkler
heads to be operated, discharge required and pump characteristics.
System Layout
Layout of the system mainly depends on topography, field size and shape, location and
source of water supply.
Required System Capacity
The capacity of a sprinkle system depends on the area to be irrigated; the gross depth of
water required for each irrigation, time allowed completing irrigation during peak water use
period, f and actual operating time per day. The capacity can be calculated by the equation as:
fT
Ad
K Q
s
= (4.1)
Where,
Q
s
= discharge capacity, l/s (gpm)
K = conversion constant, 2.78 for metric and 453 for English units
50
A = design area, ha (acres)
d = gross irrigation depth, mm (in)
f = operating time allowed to complete one irrigation, days
T = actual operating time per day, h/day
For a given area, d, f and T have significant impact on the system capacity. Greater the irrigation
depth, larger will be the system capacity. For a given depth, d, greater the operating time (IT),
the smaller will be the system capacity and vice versa. The operating time, f should be one to two
days less than the irrigation cycle during peak water use period. Similarly T should be at least 5
to 10 percent less than 24 hours.
Number of Sprinklers
Number of sprinklers operating at one time can be calculated as:
a
s
q
Q
=
s
N (4.2)
Where,
N
s
= number of sprinklers operating
Q
s
= total system discharge, L/s (gpm)
q
a
= average sprinkler discharge, L/s (gpm)
Variation in the number of operating sprinklers should be kept to minimum but it may be
sometime unavoidable at the end of an irrigation cycle. On the last day of irrigation, only a few
laterals may be needed to complete the irrigation.
Number of Laterals Operated per Set
Length of laterals may be determined by deciding the final layout of the system. Dividing
the lateral length by the spacing between the sprinkler heads on the lateral will give the number
of sprinklers operating on the lateral.
Divide the total minimum number of sprinklers required by the number of sprinklers on
one lateral to get the minimum number of lateral required. The above computations are finally
adjusted as illustrated in the following example:
51
Sample Example 4.1: Computing system capacity requirements for a single crop in the design
area.
Given: Field of corn, A = 16 ha
Design moisture use rate, U
d
= 5 mm/day
Moisture replaced in soil at each irrigation, d
n
= 60 mm
Irrigation efficiency, E
a
= 75%
Irrigation period, f = 10 days in a 12-day interval
System operating time per day, T = 20 hr/day
Electrical conductivity of the irrigation water, EC
w
= 2.1 dS/m
Calculations: Leaching requirements can be calculated by using the following equation:
100 / E LR) - (1.0
d 0.9
d
a
n
=
in) (3.54 mm 90
100 / 75 0.20) - (1.0
60 0.9
d =

=
Using the equation to compute the system capacity:
Q
s
= 2.78 Ad
fT
gpm) (317 L/s 20.0
20 10
90 16 2.78
=


=
Sample Example 4.2:A side roll sprinkle system is to be designed for the field shown below;
laterals will operate on both sides of the main line.
Pump
(in center)
1
2
0
0

f
t
2100 ft
M
a
i
n
l
i
n
e
52
The following data is for the preliminary design:
q
a
= 3.8 gpm/sprinkler f = 5 days
S
e
= 30 feet/sprinkler d = 1.11 inches
S
1
= 40 feet/position T = (3 sets/day) x (4 hrs/set)
1. What is the preliminary system capacity (gpm)?
(There are 43,560 sq ft/acre)
2. How many sprinklers will be operating simultaneously?
(Round to the nearest whole number)
3. How many sprinklers per lateral?
4. How many lateral(s) will be needed?
5. How many sets for each of the lateral(s)?
6. What is the adjusted system capacity (gpm)?
Solution:
1. gpm 484.82
5 12
1.11
43560
1200 2100
453
T.f
Ad
K Q
sinitial
=

= =
2. 128 127.58
3.8
484.82
q
Q
N
a
s
s
~ = = =
3. Sprinkler per lateral = 35
30
1050
=
4. No. of laterals needed = 4 65 . 3
35
128
~ =
5. Settings of laterals = main the of side each on 30
40
1200
=
- If 1 lateral is operated, field will be irrigated in = 60/3 = 20 days
- 2 60/3 = 20 days
- 60/6 = 10 days
- 3 60/9 = 7 days
- 4 60/12 = 5 days
Therefore 4 laterals are needed (20 each side of the main line)
Therefore setting of each lateral = 600/40 = 15
6. Adjusted Q
s
= (No. of laterals) 4 x 35 x q
a
= 4 x 35 3.8 = 532 gpm
53
Layout of Laterals for Set Sprinklers
I. Selecting Sprinkler Discharge, Spacing, and Pressure
In Chapter 6 of the textbook there are several tables that provide guidelines for nozzle
sizes for different:
Wind conditions
Application rates
Sprinkler spacings
For selected values of wind, application rate, and spacing, the tables provide
recommended nozzle sizes for single and double-nozzle sprinklers, recommended
sprinkler pressure, and approximate uniformity (CU).
Table values are for standard (non-flexible) nozzles.
Table values are for standard sprinkler and lateral spacings.
More specific information can be obtained from manufacturer's data.
Recall that the maximum application rate is a function of soil texture, soil structure, and
topography (Table 5.4).
For a given spacing and application rate, the sprinkler discharge, q
a
, can be determined
from Eq. 5.5.
to
I e
S
S S I
pa
I e n
a
E 3600
S S d
3600
) (
q = = (4.3)
Where, q
a
is in lps; I is in mm/hr; d
n
is in mm; S
to
is the operating time for each set, in
hours; S
e
and S
l
are in m and E
pa
is application efficiency.
II. Number of Operating Sprinklers
After calculating the system capacity and the design flow rate for sprinklers, the number
of sprinklers that will operate at the same time is:
a
s
s
q
Q
N = (4.2)
Where, N
s
is the minimum number of sprinklers operating, and Q
s
and q
a
have the same
units.
54
It is recommendable to always operate the same number of sprinklers when the system is
running. This practice can help avoid the need for pressure regulation, and can avoid
uniformity problems. It can also help avoid wasting energy at the pump.
For odd-shaped fields, and sometimes for rectangular fields, it is not possible to operate
the same number of sprinklers for all sets. In this case, pressure regulation may be
necessary, or other steps can be taken (multiple pumps, variable-speed motor, variable
application rates).
III. Lateral Design Criteria
Lateral pressure varies from inlet to extreme end due to:
1. friction loss
2. elevation change
The fundamental basis upon which sprinkler laterals are designed is:
"Pressure head variation in the lateral should not exceed 20% of the
average design pressure for the sprinklers"
This is a design assumption that has been used for many years, and is based on a great
deal of experience.
The 20% for pressure variation is not an "exact" value; rather, it is based on judgment and
some cost comparisons.
A designer could change this value, but it would affect system performance (uniformity),
initial system cost, operating cost, and possibly other factors.
Computer programs could be written to search for an "optimal" percent pressure variation
according to initial and operating costs, and according to crop value -- such an "optimal"
value would vary from system to system.
IV. Sprinkler Lateral Orientation
It is usually preferable to run laterals on contours (zero slope) so that pressure variation in
the lateral pipes is due to friction loss only.
It is advantageous to run laterals downhill, if possible, because the gain in energy due to
elevation change will allow longer laterals without violating the 20% rule. But, if the
slope is too steep, pressure regulators or flow control nozzles may be desirable.
55
If the ground slope is equal to the friction loss gradient, the pressure in the lateral will be
constant.
However, the friction loss gradient is nonlinear because the flow rate is decreasing with
distance along the lateral.
Figure 4.1 An example of pressure drop on zero slope, downhill and uphill lateral positions.
Lateral
Ground
56
It is usually not recommendable to run laterals in an uphill direction. In this case:
1. Both friction loss and elevation are working to reduce pressure toward the end of
the lateral, and length is more restricted if the 20% rule is still used.
2. However, for small slopes, running laterals uphill may be required to reduce the
total length of the mainline pipe.
Note that V
2
/2g in the lateral pipe is normally converted into total head as the water flows
through the nozzle body. Therefore, the velocity head (and EL) should normally be
considered in lateral design. However, since a portion of the velocity head is lost during
deceleration of the water at the entrance into risers and as turbulence inside the sprinkler
head, and since V
2
/2g in a lateral pipe is typically small (< 1 ft of head, or 0.2 psi, or 0.3
m head, or 3 kPa), it is normally neglected during design, and the HGL is used.
Aside from limits on pressure variation, laterals should be oriented so that they move in
the direction of the prevailing winds this is because of salinity problems and application
uniformity.
Figure 4.2 gives examples of layouts on different topographies.
Topography: Topography of the field greatly affects layout of main pipeline and laterals. As a
general rule, the lateral should be laid across the slope and pressure variation along a lateral
should not exceed 20% of the average design operating pressure (Keller and Bliesner (1990;
page 126). Variation in discharge along a lateral should be minimum and normally variation in
discharge should not exceed 10%.
Layouts of main lines and laterals for different field topography are shown in Fig. 4.2.
Note that there may be a number of pipe layouts possible but the best one should be selected
based on above principles of pressure and discharge variations. To meet these requirements, the
laterals should be laid on the contours or across prominent field slopes (Figure. 4.2 a, b & c). The
lateral length should be limited to that length in which the friction loss does not exceed 20% of
the designed operating pressure. Running laterals uphill should be avoided as far as possible. If
this situation is unavoidable, then use shorter lateral on the uphill side. Pressure and flow
regulators should be used to avoid excessive variation in pressure or discharge respectively.
Longer laterals may be used if they are laid down the slope. Under this condition,
difference in elevation in both ends of the lateral results gain in pressure head. Thus longer
laterals may be used than the similar one lay on level field (Fig. 4.2 d, e, f). Two main-lines are
laid in Figure 4.2 to avoid running the laterals uphill.
57
Figure 4.2 Layout of laterals for periodic-move sprinkle system with
different topographic conditions (Source: Keller and Bliesner,
1990; page 125)
58
V. Lateral Sizing Limitations
Lateral pipes can be designed with multiple diameters to accommodate desirable pressure
distributions, but...
Hand-move laterals should have only one or two different pipe sizes to simplify handling
during set changes.
In practice, hand-move systems and wheel lines usually have only one size of lateral
pipe.
Some wheel lines, greater than 400 m in length, may have 5-inch pipe near the inlet and
then 4-inch pipe at the end.
Layout of Mainline for Set Sprinklers
I. Mainline Layout and Sizing
If possible, run the mainline up or down slope so the laterals can be on contours (lateral
pressure variation due to friction loss only).
Can also run the mainline along a ridge so the laterals run downhill on both sides (lateral
friction loss partially offset by elevation change).
Should consider possible future expansions when sizing the mainline.
"Split-Line" Lateral Operation:
Laterals operate on both sides of the mainline.
The mainline can be sized for only half capacity halfway down the mainline if laterals are
run in different directions.
Sometimes interferes with cultural practices.
It is convenient to have the water supply in the center of one side of the field, but this is
seldom a design variable (the well is already there, or the canal is already there).
May not need pumping if the water supply is at a higher elevation than the field elevation
(e.g. 50 psi = 115 ft or 35 m of head) when pumping is not required, this changes the
mainline layout and pipe sizing strategy.
In some cases it will be justifiable to include one or more booster pumps in the design --
even when the water source is a well (the well pump may not provide enough pressure for
any of the lateral settings).
II. Design Variables to Accommodate Layout
Number of sprinklers operating
Average application rate
59
Gross application depth
Average sprinkler discharge
Sprinkler spacing
Operating hours per day
Irrigation frequency
Total operating time (fT)
System capacity
Percent probability of rain during peak-use period
MAD
It may be necessary to adjust the layout if a suitable combination of the above variables
cannot be found.
Can also use flow control nozzles or pressure regulators to accommodate a given layout.
Sample Example 4.3:
Consider a periodic-move system with S
l
= 50 ft, S
e
= 40 ft, f = 8 days, T = 11.5 hrs @ 2
sets/day, d = 2.7", and q
a
= 4.78 gpm
The field size is 80 acres (% of a "quarter section"), 2,640 ft on one side and 1,320 ft on
the other, rectangular
The laterals will have to be 1,320 ft long
System capacity:
gpm 532
hrs/set) 11.5 sets/day)( (2 days) 8 (
inch) (2.7 ac) 80 ( 453
Q
s
= = (4.4)
Number of sprinklers operating:
sprinklers 111
78 . 4
532
N
s
= = =
a
s
q
Q
(4.5)
Number of laterals,
/laterals sprinklers 33
er ft/sprinkl 40
ft/lateral 1320
= (4.6)
laterals 3.36
/lateral sprinklers 33
sprinklers 111
= (4.7)
so, round up to 4 laterals
60
Thus, two laterals on each side of the mainline (symmetry)
26.4
n ft/positio 50
pair lateral per ft 1320
= (4.8)
Round this up from 26.4 to 27 positions per lateral pair.
This gives 2 x 27 = 54 total lateral positions, and 54/4 = 13.5 sets/lateral.
Use 13 sets for two laterals and 14 sets for the other two laterals.
Then, there will be 14 sets per irrigation, even though the last set will only have two laterals
operating.
Adjusted irrigation frequency:
days 7
sets/day 2
sets 14
f = = (4.9)
Note that the value of f was for an 8-day interval
Thus, we need to increase Q
s
to complete the irrigation in less time
Adjusted system capacity:
Q
s
= (4 laterals)(33 sprinklers/lateral)(4.78 gpm/sprinkler) (4.10)
= 631gpm
Another way to adjust the system capacity:
61
gpm 608 gpm) (532
days 7
days 8
Q
s
=
|
|
.
|

\
|
= (4.11)
You might say that we are "effectively" finishing in somewhat less than 7 days, because
the last set has only two laterals in operation, giving a system capacity of 608 instead of
631.
Consider this calculation: there are 2 x 13 + 2 x 14 = 54 sets, but the last 2 sets have only
2 laterals, so, (52/54) x 631 = 608 gpm, as calculated above. Which is correct?
There is (52/54)*(4 laterals) = 3.85 laterals operating on average during each irrigation of
the field. However, you cannot always base the system capacity on the average number of
laterals operating.
The system capacity should be based on the "worst case", which is when all four laterals
operate simultaneously.
This means that the required capacity is 631 gpm, not 608 gpm.
Note that many farmers will accept some increase in system capital cost to provide more
operational flexibility and safety.
In summary, we have essentially lowered f (irrigation frequency) to accommodate the
system configuration (layout), but:
same gross depth
same number of hours per set
same sprinkler flow rate
same sprinkler spacing
increased system capacity
Sample Example 4.4: A field measuring 800 m x 400 m is considered for designing a sprinkle
system. Field data indicates that irrigation depth of 5 cm should be applied weekly during peak
water use period. There will be two 11 hours settings per day. It is desirable to complete the
irrigation in six days allowing one day for maintenance of the system and other farming activities
q
a
= 20 lpm and sprinkler spacing = 15 x 20 m
2
. Assume it is not desirable to use lateral longer
than 400 m in this example.
Calculations:
L/s 33.7
22 6
50
10,000
400 800
2.78
fT
Ad
K Q
s
=

= =
Total number of sprinklers, N
s
= Q
s
/1 = 33.7 = 101
20/60
62
Number of sprinkler per lateral = 400/15 = 26.66 27
Minimum whole number of laterals required = 102/27 = 3.8 4
The number of lateral positions on each side of the main line with S
1
= 20 m is:
= 400/20 = 20 position
Total numbers of lateral positions on both sides of the main line are 20 x 2 = 40. Thus four
laterals with two settings per day will cover the field in 5 days. Thus design is conservative and
the whole field will be irrigated in five days instead of six days.
Final adjustment: With four laterals operating simultaneously, the maximum number of
sprinklers running is:
N
sp
= 4 x 27 = 108
Actual discharge of the system should be = 108 x 20/60 = 36 Lps
Mainline with 20 hydrants at 20 m 0
0
0
0
0
0
0
0
Lateral with 27 sprinklers
Number of sprinkler operating = 27 x 4 =
108
System capacity = 36 lps
400 m
800 m
Figure 4.3 Sprinkler systems layout in 800 m x 400 mm field.
Sample Example 4.5: Calculations for a Periodic-Move System
Given:
Crop is alfalfa. Top soil is 1.0 m of silt loam, and subsoil is 1.8 m of clay loam. Field area is
35 ha. MAD is 50% and EC
w
is 2.0 dS/m. Application efficiency is estimated at 80%, and the
soil intake rate is 15 mm/hr. Lateral spacing is 15 m and lateral length is 400 m. Assume it
takes 1/2 hour to change sets. Seasonal effective rainfall is 190 mm; climate is hot. Assume
one day off per week (irrigate only 6 days/week).
From tables in the textbook:
Hot climate, table 3.3 gives ................U
d
= 7.6 mm/day, and U = 914 mm/season
Top soil, table 3.1 gives ..................................................................... W
a
= 167 mm/m
Sub soil, table 3.1 gives ..................................................................... W
a
= 183 mm/m
63
Root depth, table 3.2 gives ................................................. Z = (1.2 + 1.8)/2 = 1.5 m
Salinity for 10% yield reduction, Table 3.5 gives .............................. EC
e
= 3.4 dS/m
1. Average water holding capacity in root zone:
top soil is 1.0 m deep; root zone is 1.5 m deep...
mm/m 3 . 172
5 . 1
) 183 )( 0 . 1 5 . 1 ( ) 167 ( 0 . 1
W
a
=
+
= (4.12)
2. Max net application depth (Eq. 3.1):
mm 129.2 5) (172.3)(1.
100
50
Z W
100
MAD
d
a x
= |
.
|

\
|
= = (4.13)
3. Maximum irrigation interval (Eq. 3.2):
days 17.0
mm/day 7.6
mm 129.2
U
U
f
d
x
x
= = = (4.14)
4. Nominal irrigation interval (round down, or truncate):
f

= trunk (f
x
) = 17 days (4.15)
5. Net application depth:
d
n
= f'U
d
= (17 days)(7.6 mm/day) = 129.2 mm (4.16)
6. Operating time for an irrigation:
17 days is just over two weeks, and depending on which day is off, there could be 3 off
days in this period. So, with one day off per week, we will design the system capacity to
finish in 17 - 3 = 14 days. Thus, f = 14 days. But, remember that we still have to apply 17
days worth of water in these 14 days (we irrigate 6 days/week but crop transpires 7
days/week).
7. Leaching requirement (Eq. 3.3):
13 . 0
2.0 - 5(3.4)
2.0
EC 5EC
EC
LR
w E
w
= =

= (4.17)
64
LR > 0.1; therefore, use Eq. 5.3 b...8. Gross application depth (Eq. 5.3b):
mm 167.1
0.13)(0.8) - (1
0.9(129.2)
) 100 / LR)(E - (1
0.9d
d
a
n
= = = (4.18)
9. Actual set operating time:
With 167.1 mm to apply and a soil intake rate of 15 mm/hr, this gives 11.14 hrs minimum
set time (so as not to exceed soil intake rate). Then, we can make the nominal set time
equal to 11.5 hours for convenience. With 0.5 hrs to move each set, there are a total of
12.0 hrs per set, and the farmer can change at 0600 and 1800 (for example).
At this point we could take the sprinkler spacing, S
e
, lateral spacing, S
l
, and actual
application rate to determine the flow rate required per sprinkler.
10. Sets per day:
From the above, we can see that there would be two sets per day.
11. Number of sets per irrigation:
(14 days/irrigation)(2 sets/day) = 28 sets
12. Area per lateral per irrigation:
Lateral spacing on mainline is S
l
= 15 m. Lateral length is 400 m. Then, the area per
lateral is:
(15 m/set)(28 sets)(400 m/lateral) = 16.8 ha/lateral
13. Number of laterals needed:
35 ha/16.8 ha/lateral = 2.08 laterals (4.19)
Normally we would round up to the nearest integer, but because this is so close to 2.0 we
will use two laterals in this design.
14. Number of irrigations per season:
(U - P
e
)/d
n
= (914 mm - 190 mm)/129.2 mm/irrig = 5.6 irrigations
Thus, it seems there would be approximately six irrigations in a season. But the initial R
z
is less than 1.5 m, so there may be more than six irrigations.
65
15. System flow capacity (Eq. 5.4):
with 11.5 hours operating time per set and two sets per day, the system runs 23 hrs/day...
fT
Ad
2.78 Q
s
= (4.1)
hrs/day) (23 days) (14
mm) (167.1 ha) (35
2.78 Q
s
=
= 50.5 lps (800 gpm)
This is assuming no effective precipitation during the peak ET period, so the design should be
conservative.
66
Unit 5
PIPELINE HYDRAULICS, FRICTION LOSSES AND DESIGN OF LATERALS
Pressure and Pressure Variations
Water must be applied at suitable pressure for good performance of all sprinkle
systems. In a stationary container, pressure at any depth is same everywhere and is equal to
the product of the unit weight of water 1000 kg/m
3
at 20
o
C (62.4 lb/ft
3
at 60
o
F) and the height
of water above that point. Height of water in the column is called the head. Note that
diameter of the pipe or shape of container has no effect on pressure. Pressure is usually
expressed in kilopascals, kPa (pounds per square inch, psi) and head is expressed in m (ft) of
water. One kPa (psi) of pressure is equal to 0.102 m (2.31 ft) of water head or 1 m of pressure
head is equivalent to 9.8 kPa of pressure. In a sprinkle system, pressure or head consists of
different components: static head, pressure head, velocity head, friction head and elevation
head above a datum point.
Dynamic Head: The operating head of a sprinkle system consists of different
components as discussed below:
i) Total Static Head: Static head is equal to the difference in elevation between the
highest discharge point and the reference point. The distance from the water surface to the
center line of pump is called static suction head. If water supply is from a well or sump, there
may be excessive suction due to drawdown. Sum of static head and suction is called the total
static head.
Pressure Head: The pressure head is necessary to spray water through the sprinkler
nozzle which converts the pressure head into velocity head that carries the water away from
the sprinkler.
Velocity Head: Velocity of water in a sprinkler system seldom exceeds 2.5 m (8 ft) per
second. Therefore, the velocity head may be neglected except in computing suction head for
centrifugal pumps.
Friction Head: The pressure drops when water flows through the pipes and pipe
fittings. Sudden increase or decrease in pipe diameter also cause drop in pressure. This loss in
pressure is called friction head. Hazen-Williams equation is commonly used to calculate
friction head loss in sprinkle and trickle systems which is:
4.87
1.852
1.852
f
D
(C)
(Q)
K
L
100h
J

= = (5.1)
Where
J = head loss gradient, m/100 m (ft/100 ft)
K = conversion constant, 1.217 x 10
12
for metric units (1050 for English units)
h
f
= head loss due to pipe friction, m (ft)
67
L = length of pipe, m (ft)
Q = flow rate in the pipe, L/s (gpm)
C = friction coefficient, which is a function of pipe material characteristics
D = inside diameter of the pipe, mm (in.)
Typical values of C for use in the Hazen-Williams equation are:
Pipe Material C
Plastic 150
Epoxy-coated steel 145
Cement asbestos 140
Galvanized steel 135
Aluminum (with couplers every 30 ft) 130
Steel (new) 130
Steel (15 years old) or concrete 100
Source: (Keller and Blisner, 1988, page 135)
For smooth plastic pipes, Watters and Keller (1978) proposed the following equations as
discussed by Keller and Bliesner (1988, page 138):
J = 100 h
f
/L = K Q
1.75
/D
4.75
(for D < 125 mm) (5.2)
and
J = 100 h
f
/L = K Q
1.83
/D
4.83
(for D > 125 mm) (5.3)
Where, K and K
1
are conversion constants. K = 7.88 x 10
7
for metric units (0.133 for English
units, and K
1
= 9.58 x 10
7
for metric units and 0.100 for English units). Other symbols are as
defined above.
- For pipe friction loss, use the Hazen-Williams and Darcy-Weisbach equations
- We can use the Swamee-Jain equation instead of the Moody diagram to determine the
f value i.e.
(
(

|
|
.
|

\
|
+
=
9 . 0
10
R
74 . 5
D 75 . 3
log
0.25
f
y

(5.4)
Which is valid for turbulent flow in the range: 4,000 < R
y
< 1.0(10)
8
, where R
y
is the
Reynolds number.
The ratio /D is called "relative roughness".
68
The roughness height, , varies widely according to pipe material and
condition/age.
Use the Blasius equation (Eq. 8.6) to determine the value of "f," in some cases, for
"smooth pipes" (e.g. plastic pipes) which is:
F
f
= 0.32 R
y
-0.25
(5.5)
Where, F
f
is the Darcy-Weisbach pipe friction factor. The F
f
is related to the head loss in the
pipe, h
f
by the Darcy-Weisbach equation as:
2g
V
D
L
F h
2
f f
= (5.6)
Where
V = Velocity of flow in the pipe, (m/s) (ft/s)
g = Acceleration due to gravity, 9.81 m/s
2
(32.2 ft/s
2
)
D = Inside pipe diameter, m (ft)
Hazen-Williams Equation
This is a simple, empirical pipe friction-loss equation:
100
JL
h
f
= (5.7)
Where h
f
is the friction loss (head of water); L is the pipe length; and,
4.87 - 852 . 1
D )
C
Q
( K J = (5.8)
In which Q is the flow rate; C is a roughness coefficient; and D is the pipe inside diameter.
The value of K in this equation is:
K = 1,050 for Q in gpm and D in inches
K = 16.42(10)
6
for Q in lps and D in cm
K=1.217(10)
12
for Q in lps and D in mm
Darcy-Weisbach Equation
This is a (usually) more accurate pipe friction-loss equation:
2g D
LV
F h
2
f f
= (5.9)
Where F
f
is usually determined from Eq. 4, or from the Blasius equation (see below).
69
Blasius Equation
This can be used to estimate the F
f
value for smooth (e.g. plastic) pipes of the typical
diameters found in pressurized irrigation systems:
-0.25
y f
R 0.32 F = (5.10)
Where, 2,000 < R
y
< 100,000
Sample Example: Computation of friction losses
Given: A 100 mm (4 in) aluminum lateral pipeline 396 m long with 12 m between outlets and
sprinklers discharging 30 L/min. (The inside diameter of the pipe is 99 mm).
Calculations: The number of sprinklers: N= 400/12 = 33.33 (round no. 33)
The lateral discharge is:
Q
I
= 33 x 30 = 990 L/min = 16.5 L/s
And using the equation (8)
5.06 ) 99 (
130
5 . 16
10 x 1.21 J
87 . 4
852 . 1
12
=
|
.
|

\
|
=

And by Eq. (1) with F = 0.36 from Table 8.7:
kPa 70.7 m 7.21
100
396
0.36 x 5.06 h
f
= = =
Set Sprinkler Lateral Design
I. Basic Design Criterion
1. The basic design criterion is to size lateral pipes so that pressure variation along the
length of the lateral does not exceed 20% of the nominal design pressure for the
sprinklers.
2. This criterion is a compromise between cost of the lateral pipe and application
uniformity in the direction of the lateral.
3. Note that the locations of maximum and minimum pressure along a lateral pipe can vary
according to ground slope and friction loss gradient.
4. For laterals on level grounds, change in P
e
= 0.0 (H
e
=0.0), and the allowable pressure
loss due to friction in the lateral line will be equal to 20% of P
a
(average pressure).
5. For uphill laterals (see Figure 9.2a) P
f
(pressure loss due to friction) may be equal to 20%
of P
a
minus the static pressure difference due to elevation, P
e
, which is the difference in
elevation between the inlet and closed ends of the lateral.
6. For uphill lateral, the minimum pressure, P
n
(nominal pressure) occurs at the closed end
of the pipeline.
70
7. For downhill laterals the allowable pressure P
f
is 20% of P
a
plus the static pressure gain
due to the decrease in elevation between the inlet and closed ends of the laterals.
8. For downhill laterals P
n
occurs at the point along the lateral where the pipe friction
gradient equals the slope of the lateral (ground as shown in Figure 9.2b).
II. Location of Average Pressure in the Lateral
We are interested in the location of average pressure along a lateral pipe because it is related
to the design of the lateral. Recall that friction head loss along a multiple-outlet pipe is
nonlinear. The figure below is for a lateral laid on level ground - pressure variation is due to
friction loss only.
Figure 5.1. Example of pressure drop in a multiple outlet pipe (lateral)
For equally-spaced outlets (sprinklers) and approximately thirty outlets (or more), three-
quarters of the pressure loss due to friction will occur between the inlet and the location of
average pressure.
The location of average pressure in the lateral is approximately 40% of the lateral length,
measured from the lateral inlet.
If there were only one outlet at the end of the lateral pipe, then one-half the pressure loss due
to friction would take place between the lateral inlet and the location of average pressure, as
shown below.
71
Figure 5.2. Pressure Drop for one outlet at end of the lateral
Consider the following equations:
Total friction head loss:

=
=
n
1 i
i f total f
) (h ) (h (5.11)
1 n
) (h
) (h
n
1 i
i
1 j
j f
a f
+
=

= =
(5.12)
where n is the number of sprinklers; (h
f
)
total
is the total friction head loss from 0 to L; (h
f
)
i
is the friction head loss in the lateral pipe between sprinklers i-1 and i; and (h
f
)
a
is the
friction loss from the lateral inlet to the location of h
a
.
As indicated above, (h
f
)
a
occurs over approximately the first 40% of the lateral.
Note that between sprinklers, the friction head loss gradient is linear in the lateral pipe
with one outlet at the end.
Note also that (h
f
)
0
= 0, but it is used in calculating (h
a
)
f
, so the denominator is (n+1), not
n.
Friction head loss to location of h
a
:
72
Figure 5.3 Pressure losses are non-linear in a multiple outlet lateral
- In applying these equations with sample data, the following result can be found (see
example 9.1 in the text).
0.73
) (h
) (h
total f
a f
~ (5.13)
This supports the above claim that approximately 3/4 of the friction head loss occurs
between the lateral inlet and the location of h
a
.
Also, from these calculations it can be seen that the location of h
a
is approximately
38% of the lateral length, measured from the inlet, for laterals with approximately 30
or more sprinklers.
But, this analysis assumes a constant q
a
, which is not quite correct unless flow control
nozzles and or pressure regulators are used at each sprinkler.
We could eliminate this assumption of constant q
a
, but it involves the solution of a
system of nonlinear equations.
III. Location of Minimum Pressure in Laterals Running Downhill
The location of minimum pressure in a lateral running downhill is where the slope of
the friction loss curve, J, equals the ground slope.
73
The above assertion is analogous to a pre-calculus "max-min problem", where you take
the derivative of a function and set it equal to zero (zero slope).
Here we are doing the same thing, but the slope is not necessarily zero.
Hazen-Williams Equation
87 . 4
852 . 1
12
(10) 1.217 J

|
.
|

\
|
= D
C
Q
(5.14)
for J in meters of friction head loss per 100 m (or ft/100 ft); Q in lps; and D in mm
In this equation we will let
x
S
q
Q Q
e
a
|
|
.
|

\
|
=
1
(5.15)
for multiple, equally-spaced sprinkler outlets spaced at S
e
(m) from each other, with
constant discharge of q
a
(lps). Q
l
is the flow rate at the lateral inlet (entrance).
To find the location of minimum pressure, let J = S, where S is the ground slope (in %,
because J is per 100 m), which is negative for downhill-sloping laterals.
Combining the two above equations and solving for x,
)] ) ( ( ) 10 ( 3 [
63 . 2 54 . 0 7
D S C Q
q
S
x
I
a
e
=

(5.16)
Where, x is the distance, in m, from the lateral inlet to the minimum pressure
S is in percent; S
e
and x are in m; D is in mm; and Q
l
and q
a
are in lps.
Note that the valid range of x is: 0 < x < L, and that you won't necessarily get J = S
over this range of x values:
If you get x < 0 then the minimum pressure is at the inlet
If you get x > L then the minimum pressure is at the end
This means that the above equation for x is valid for all ground slopes: S = 0, S > 0
and S < 0
74
IV. Required Lateral Inlet Pressure Head
Except for the most unusual circumstances (e.g. non-uniform downhill slope that
exactly matches the shape of the h
f
curve), the pressure will vary with distance in a
lateral pipe.
According to Keller & Bliesner's design criterion, the required inlet pressure head to a
sprinkler lateral is that which makes the average pressure in the lateral pipe equal to
the required sprinkler pressure head, ha.
We can force the average pressure to be equal to the desired sprinkler operating
pressure by defining the lateral inlet pressure head as:
Design Equation
h
l
= h
a
+ 3/4h
f
+1/2h
e
(5.17)
h
l
is the required pressure head at the lateral inlet.
Strictly speaking, we should take approximately 0.4h
e
in the above equation, but we
are taking separate averages for the friction loss and elevation gradients.
Of course, instead of head, h, in the above equation, pressure P could be used if
desired.
The value of h
e
is negative for laterals running downhill
75
For steep downhill slopes, where the minimum pressure would be at the lateral inlet,
it is best to let.
h
f
= -h
e
(5.18)
Thus, we would want to consume, or "burn up", excess pressure through friction loss
by using smaller pipes.
To achieve this equality for steep downhill slopes, it may be desirable to have more
than one pipe diameter in the lateral. A downhill slope can be considered "steep"
when (approximately).
-h
e
> 0.3h
a
(5.19)
We now have an equation to calculate lateral inlet pressure based on h
a
, h
f
, and h
e.
However, for large values of h
f
there will be correspondingly large values of h
e
. Thus,
for zero ground slope, to impose a limit on h
f
we will accept:
h
f
= 0.20 h
a
(for S = 0 only) (5.20)
This is the same as saying that we will not allow pipes that are too small, that is, pipes
that would produce a large h
f
value.
An additional head term must be added to the equation for h
l
to account for the change in
elevation from the lateral pipe to the sprinkler (riser height):
r e f a
h h h h + A + + =
2
1
4
3
h
I
(5.21)
or, in terms of pressure
r e f a
P P P P + A + + =
2
1
4
3
P
I
(5.22)
V. Friction Losses in Pipes with Multiple Outlets
Pipes with multiple outlets have decreasing flow rate with distance (in the direction of
flow), and this causes the friction loss to decrease by approximately the square of the
flow rate (for a constant pipe diameter).
Sprinkler and trickle irrigation laterals fall into this hydraulic category.
Multiply the head loss for a constant discharge pipe by a factor "F" to reduce the total
head loss for a lateral pipe with multiple, equally spaced outlets:
100
JFL
h
f
= (5.23)
Where F is from Eq. 8.9a as given below:
76
2
6
1
2
1
1
1
N
b
N b
F

+ +
+
= (5.24)
for equally spaced outlets, each with the same discharge, and going all the way to the
end of the pipe.
All of the flow is assumed to leave through the outlets, with no "excess" spilled out
the downstream end of the pipe.
N is the total number of equally spaced outlets.
The value of b is the exponent on Q in the friction loss equation.
The first sprinkler is assumed to be located a distance of S
e
from the lateral inlet
Eq. 8.9b (see below) gives F(), which is the F factor for initial outlet spacings less than or
equal to Se.
) 1 (
) - (1 - NF
) F(


=
N
(5.25)
where 0 < < 1
Note that when = 1, F() = F
Many sprinkler systems have the first sprinkler at a distance of S
e
from the lateral
inlet ( = 0.5), when laterals run in both orthogonal directions from the mainline.
VI. Lateral Pipe Sizing for a Single Pipe Size
If the minimum pressure is at the end of the lateral, which is the case for no ground
slope, uphill and slight downhill slopes, then the change in pressure head over the
length of the lateral is:
h = h
f
+h
e
(5.26)
If we allow h = 0.20 h
a
, then
0.20h
a
= h
f
+ h
e
(5.27)
0.20h
a
- h
e
= J
a
FL/100 (5.28)
and,
|
.
|

\
| A
=
FL
h h
J
e a
a
20 . 0
100 (5.29)
Where, J
a
is the allowable friction loss gradient.
Lateral pipe diameter can be selected such that J s J
a
The above portion is the part of a standard lateral design criterion and will give a system CU
of approximately 0.97CU if lateral inlet pressures are the same for each lateral position, for
set sprinkler systems.
77
If the lateral is sloping downhill and the minimum pressure does not occur at the end of the
lateral, then we will attempt to consume the elevation gain in friction loss as follows,
e
h A =
f
h (5.30)
|
.
|

\
| A
=
FL
h
J
e
a
100 (5.31)
Note that in this case h h
f
+ h
e
. Rather, h = h
max
- h
min
, where:
1. h
max
is either at the lateral inlet or at the end of the lateral, and
2. h
min
is somewhere between the lateral inlet and the end
Given a value of J
a
, the inside diameter of the lateral pipe can be calculated from the
Hazen-Williams equation:
205 . 0
852 . 1
a
) (
J
K
D
(

=
C
Q
I
(5.32)
where Q
l
is the flow rate at the lateral inlet (Nq
a
) and K is the units coefficient in the
Hazen-Williams equation.
The calculated value of D would normally be rounded up to the next available internal
pipe diameter.
VII. Lateral Design Example
VI.1. Given information:
L = 396 m (lateral length)
q
a
= 0.315 lps (nominal sprinkler discharge)
S
e
= 12 m (sprinkler spacing)
h
r
= 1.0 m (riser height)
slope = -2.53% (going downhill)
P
a
= 320 kPa (design nozzle pressure)
pipe material = aluminum
VI.2. Calculations leading to allowable pressure head loss in the lateral:
N
n
= 396/12 = 33 sprinklers F = 0.36
Q
l
= (0.315)(33) = 10.4 lps
h
e
= SL = (-0.0253)(396) = -10.0 m
(Pf)
a
= 0.20P
a
- h
e
= 0.20(320 kPa) - 9.81(-10.0 m) = 162 kPa
(hf)
a
= 162/9.81 = 16.5 m
78
VI.3. Calculations leading to required lateral pipe inside diameter:
0.3Pa = 0.3(320 kPa) = 96.0 kPa
= 96.0/9.81 = 9.79 m
Now, 0.3h
a
< -h
e
(steep downhill). Therefore, may want to use h
f
= -h
e
. Then, Ja is:
m 100 / m 7.01
) 396 )( 36 . 0 (
) 0 . 10 (
100 )
FL
h
( 100 J
e
a
=
|
|
.
|

\
|
=

=
m
(5.33)
However, if 0.3h
a
> -h
e
, J
a
would be calculated as:
m m/100 11.6
) (0.36)(396
(16.5)
100
FL
h 0.20h
100 J
e a
a
=
|
|
.
|

\
|
= |
.
|

\
|
= (5.34)
For now, let's use J
a
= 7.01 m/100 m. Then, the minimum pipe inside diameter is (C = 130 for
aluminum):
mm 77.7 )
130
10.4
(
7.01
1.21E12
D
0.205
1.852
=
(

= (5.35)
which is equal to 3.06 inches.
In the USA, 3" aluminum sprinkler pipe has an ID of 2.9" (73.7 mm), so for this design it
would be necessary to round up to a 4" nominal pipe size (ID = 3.9", or 99.1 mm).
However, it would be a good idea to also try the 3" size and see how the lateral hydraulics
turn out (this is done below; note also that for J
a
= 11.6, D = 70.0 mm).
VI.4. Check the design with the choices made thus far
The real friction loss will be:
m m/100 2.14 mm) (99.1
130
10.4
1.21E12 J
4.87 -
1.852
=
|
.
|

\
|
= (5.36)
m
JFL
h
f
06 . 3
100
) 396 )( 36 . 0 )( 14 . 2 (
100
= = = (5.37)
The required lateral inlet pressure head is:
h
l
= h
a
+ 0.75h
f
+ 0.5h
e
+ h
r
(5.38)
h
l
= 320/9.81 + 0.75(3.06) + 0.5(-10.0) +1.0 = 30.9 m
Thus, P
l
is (30.9)(9.81) = 303 kPa, which is less than the specified Pa of 320 kPa, and this is
because the lateral is running downhill
79
VI.5. Calculate the pressure and head at the end of the lateral pipe
h
end
= h
l
- h
f
- h
e
= 30.9 - 3.06 - (-10.0) = 37.8 m (5.39)
which is equal to 371 kPa. Thus, the pressure at the end of the lateral pipe is greater than
the pressure at the inlet.
To determine the pressure at the last sprinkler head, subtract the riser height to get
37.8 m - 1.0 m = 36.8 m (361 kPa)
VI.6. Calculate the location of minimum pressure in the lateral pipe
) ) ( ( ) 10 ( 3
q
S
x
63 . 2 54 . 0 7
a
e
D S C Q
I
=

(5.40)
m 39.6 ) (99.1) 2.53) (130( 3(10) 10.4
0.315
12
x
2.63 0.54 7
= =

The result is negative, indicating that minimum pressure is really at the entrance (inlet) to the
lateral pipe. The minimum sprinkler head pressure is equal to h
l
- h
r
= 30.9 - 1.0 = 29.9 m, or
293 kPa
VI.7. Calculate the percent pressure variation along the lateral pipe
The maximum pressure is at the last sprinkler (end of the lateral), and the minimum
pressure is at the first sprinkler (lateral inlet). The percent pressure variation is:
% 21 . 0
320
293 361
min max
=

= A
a
P
P P
P (5.41)
That is, 21% pressure variation at the sprinklers, along the lateral.
This is larger than the design value of 0.20, or 20% variation. But it is very close to
that design value, which is somewhat arbitrary anyway.
VI.8. Redo the calculations using a 3" lateral pipe instead of the 4" size
In this case, the location of the minimum pressure in the lateral pipe is
m 196 ) (73.7) (130(2.53) 3(10) 10.4
0.315
12
x
2.63 0.54 7
= =

(5.42)
which is the distance from the upstream end of the lateral.
There are about 196/12 = 16 sprinklers from the lateral inlet to the location of
minimum pressure, and about 17 sprinklers from x to the end of the lateral.
80
Friction loss from x to the end of the lateral is:
87 . 4
852 . 1
12
end - x
) 7 . 73 (
130
) (17)(0.315
1.217(10) J

|
.
|

\
|
= (5.43)
= 2.65 m/100 m
m h
end x f
01 . 2
100
) 196 396 )( 38 . 0 )( 65 . 2 (
) ( =

(5.44)
Friction loss from the inlet to the end is;
87 . 4
852 . 1
12
) 7 . 73 (
130
4 . 10
) 10 ( 217 . 1

|
.
|

\
|
=
end inlet
J (5.45)
= 9.05 m /100 m
m 12.9
100
6)(396) (9.05)(0.3
) (h
end inlet f
= =

(5.46)
Then, friction loss from inlet to x is:
(h
f
)
inlet-x
= 12.9 2.01 = 10.9 m (5.47)
The required lateral pipe inlet head is:
h
l
= h
a
+ 0.75h
f
+ 0.5h
e
+ h
r
(5.48)
h
l
= 320/9.81 + 0.75(12.9) + 0.5(-10.0) +1.0 = 38.3 m
giving a P
i
of (38.3)(9.81) = 376 kPa, which is higher than P
i
for the 4 pipe
The minimum pressure head (at distance x = 196 m) is:
x inlet e x inlet f I x
) ( ) (h h h

= h (5.49)
h
x
= 38.3 10.9 (-0.0253)(196) = 32.4 m
giving a P
x
of (32.4)(9.81) = 318 kPa, which is very near Pa.
The pressure head at the end of the lateral pipe is:
h
end
= h - h
f
- h
e
= 38.3 - 12.9 + 10.0 = 35.4 m (5.50)
giving P
end
of (35.4)(9.81) = 347 kPa, which is less than P
l
. So, the maximum lateral pipe
pressure is at the inlet.
81
The percent variation in pressure at the sprinklers is based on P
max
= 376 -(1.0)(9.81) = 366
kPa, and P
min
= 318 - (1.0)(9.81) = 308 kPa:
0.18
320
308 366
P
P P
%
a
min max
=

= AP (5.51)
which turns out to be slightly less than the design value of 20% VI.9.
VI. What if the lateral ran uphill at 2.53% slope?
In this case, the maximum allowable head loss gradient is:
|
.
|

\
|
=
FL
h 0.20h
100 J
e a
a
(5.52)
m m/100 2.44
) (0.36)(396
10.0 - .81) 0.20(320/9
100 =
|
|
.
|

\
|
=
Which is negative because h
e
> 0.2h
a
, meaning that it is not possible to have only a 20%
variation in pressure along the lateral, that is, unless flow control nozzles and or other design
changes are made.
VI.10. Some observations about this design example
Either the 3" or 4" aluminum pipe size could be used for this lateral design. The 4"
pipe will cost more than the 3" pipe, but the required lateral inlet pressure is less with
the 4" pipe, giving lower pumping costs, assuming pumping is necessary.
Note that it was assumed that each sprinkler discharged 0.315 lps, when in reality the
discharge depends on the pressure at each sprinkler. To take into account the
variations in sprinkler discharge would require an iterative approach to the
mathematical solution (use a computer).
Most sprinkler laterals are laid on slopes less than 2.5%, in fact, most are on fields
with less than 1% slope.
82
Unit 6
ECONOMIC PIPE SELECTION METHOD
I. Introduction
The economic pipe selection method (Chapter 8 of the textbook) is used to balance
fixed (initial) costs for pipe with annual energy costs for pumping
With larger pipe sizes the average flow velocity for a given discharge decreases,
causing a corresponding decrease in friction loss
This reduces the head on the pump, and energy can be saved
However, larger pipes cost more to purchase
Pipe Size (diameter)
Figure 6.1 Influence of pipe size on fixed energy and total costs for a given flow rate
To balance these costs and find the minimum cost we will annualize the fixed costs,
compare with annual energy (pumping) costs
We can also graph the results so that pipe diameters can be selected according to their
maximum flow rate
We will take into account interest rates and inflation rates to make the comparison
This is basically an "engineering economics" problem, specially adapted to the
selection of pipe sizes
This method involves the following principal steps:
1. Determine the equivalent annual cost for purchasing each available pipe size
2. Determine the annual energy cost of pumping
3. Balance the annual costs for adjacent pipe sizes
83
4. Construct a graph of system flow rate versus section flow rate on a log-log scale
for adjacent pipe sizes
We will use the method to calculate "cut-off" points between adjacent pipe sizes so
that we know which size is more economical for a particular flow rate
We will use HP and kW units for power, where about % of a kW equals a HP
Recall that a Watt (W) is defined as a joule/second, or a N-m per second
Multiply W by elapsed time to obtain Newton-meters ("work", or "energy")
II. Economic Pipe Selection Method Calculations
1. Select a period of time over which comparisons will be made between fixed and annual
costs. This will be called the useful life of the system, n, in years.
The "useful life" is a subjective value, subject to opinion and financial
amortization conditions
This value could alternatively be specified in months, or other time period, but the
following calculations would have to be consistent with the choice
2. For several different pipe sizes, calculate the uniform annual cost of pipe per
unit length of pipe.
A unit length of 100 (m or ft) is convenient because J is in m/100 m or ft/100 ft,
and you want a fair comparison (the actual pipe lengths from the supplier are
irrelevant for these calculations)
You must use consistent units ($/100 ft or $/100 m) throughout the calculations,
otherwise the J values will be incorrect (see Step 11 below)
So, you need to know the cost per unit length for different pipe sizes
PVC pipe is sometimes priced by weight of the plastic material (weight per unit
length depends on diameter and wall thickness)
You also need to know the annual interest rate upon which to base the
calculations; this value will take into account the time value of money, whereby
you can make a fair comparison of the cost of a loan versus the cost of financing it
"up front" yourself
In any case, we want an equivalent uniform annual cost of the pipe over the life of
the pipeline
Convert fixed costs to equivalent uniform annual costs, UAC, by using the
"capital recovery factor", CRF
UAC = P (CRF) (6.1)
1 i) (1
i) i(1
CRF
n
n
+
+
= (6.2)
84
where P is the cost per unit length of pipe; i is the annual interest rate (fraction);
and n is the number of years (useful life)
Of course, i could also be the monthly interest rate with n in months, etc.
Make a table of UAC values for different pipe sizes, per unit length of pipe
The CRF value is the same for all pipe sizes, but P will change depending on the
pipe size
Now you have the equivalent annual cost for each of the different pipe sizes
3. Determine the number of operating (pumping) hours per year, O
t
:
hrs/year
capacity) (system
depth) annual s area)(gros (irrigated
= =
t
O (6.3)
Note that the maximum possible value of O
t
is 8,760 hrs/year (for 365 days)
Note also that the "gross depth" is annual, so if there is more than one growing
season per calendar year, you need to include the sum of the gross depths for each
season (or fraction thereof)
4. Determine the pumping plant efficiency:
The total plant efficiency is the product of pump efficiency, E
pump
, and motor
efficiency, E
motor
E
p =
E
pump
E
motor (6.4)
This is equal to the ratio of "water horsepower", WHP, to "brake
horsepower", BHP (Epump = WHP/BHP)
Think of BHP as the power going into the pump through a spinning shaft, and
WHP is what you get out of the pump - since the pump is not 100% efficient in
energy conversion, WHP < BHP
WHP and BHP are archaic and confusing terms, but are still in wide use
E
motor
will usually be 92% or higher (about 98% with newer motors and larger
capacity motors)
Epump depends on the pump design and on the operating point (Q vs. TDH)
WHP is defined as:
85
QH
102
QH
WHP = = (6.5)
where Q is in lps; H is in m of head; and WHP is in kilowatts (kW)
If you use m in the above equation, UAC must be in $/100 m
If you use ft in the above equation, UAC must be in $/100 ft
Note that for fluid flow, "power" can be expressed as pgQH = yQH
Observe that 1,000/g = 1,000/9.81 102, for the above units (other conversion
values cancel each other and only the 102 remains)
The denominator changes from 102 to 3,960 for Q in gpm, H in ft, and
WHP in HP
5. Determine the present annual energy cost:
p
f
E
C
t
O
E = (6.6)
where Cf is the cost of "fuel"
For electricity, the value of C
f
is usually in dollars per kWh, and the value used in
the above equation may need to be an "average" based on potentially complex
billing schedules from the power company
For example, in addition to the energy you actually consume in an electric motor,
you may have to pay a monthly fee for the installed capacity to delivery a certain
number of kW, plus an annual fee, plus different time-of-day rates, and others
Fuels such as diesel can also be factored into these equations, but the power output
per liter of fuel must be estimated, and this depends partly on the engine and on
the maintenance of the engine
The units of E are dollars per WHP per year, or dollars per kW per year; so it is a
marginal cost that depends on the number of kW actually required
6. Determine the marginal equipment cost:
Note that Cf can include the "marginal" cost for the pump and power unit (usually
an electric motor)
In other words, if a larger (higher head to overcome friction losses) pump & motor
cost more than a smaller pump & motor, then C
f
should reflect that, so the full
cost of friction loss is considered
If you have higher friction loss, you may have to pay more for energy to pump,
but you may also have to buy a larger pump and/or power unit (motor or engine)
It sort of analogous to the Utah Power & Light monthly power charge, based
solely on the capacity to deliver a certain amount of power
C
f
($/kWh) = energy cost + marginal cost for a larger pump & motor (6.7)
OtCf
86
where "marginal" is the incremental unit cost of making a change in the size of a
component
This is not really an "energy" cost per se, but it is something that can be taken into
account when balancing the fixed costs of the pipe (it falls under the operating
costs category, increasing for decreasing pipe costs)
That is, maybe you can pay a little more for a larger pipe size and avoid the need
to buy a bigger pump, power unit and other equipment
To calculate the marginal annual cost (MAC) of a pump & motor:
) kW (kW O
) $ ($ CRF
MAC
small big t
small big

= (6.8)
where MAC has the same units as C
f
; and $
big
-$
small
is the difference in pump +
motor + equipment costs for two different capacities
The difference in fixed purchase price is annualized over the life of the system by
multiplying by the CRF, as previously calculated
The difference in pump size is expressed as ABHP, where ABHP is the difference
in brake horsepower, expressed in kW
To determine the appropriate pump size, base the smaller pump size on a low
friction system (or low pressure system)
For BHP in kW:
pump
pump s
102E
H Q
BHP = (6.9)
Round the BHP up to the next larger available pump + motor + equipment size to
determine the size of the larger pump
Then, the larger pump size is computed as the next larger available pump size as
compared to the smaller pump
Finally, compute the MAC as shown above
The total pump cost should include the total present cost for the pump, motor,
electrical switching equipment (if appropriate) and installation
C
f
is then computed by adding the cost per kWh for energy
Note that this procedure to determine MAC is approximate because the marginal
costs for a larger pump + motor + equipment will depend on the magnitude of the
required power change
Using $
big
-$
small
to determine MAC only takes into account two (possibly
adjacent) capacities; going beyond these will likely change the marginal rate
However, this simple procedure provides a means to account for this potentially
real cost
It also means you can have different values of E for each pair of adjacent pipe
sizes
87
7. Determine the equivalent annualized cost factor:
This factor takes inflation into account:
(

+
(

+ +
=
1 i) (1
i
i e
i) (1 e) (1
EAE
n
n n
(6.10)
where e is the annual inflation rate (fraction), i is the annual interest rate
(fraction), and n is in years
Notice that for e = 0, EAE = unity (this makes sense)
Notice also that Eq. 85 has a mathematical singularity for e = i (but, in practice, i
is usually greater than e)
8. Determine the equivalent annual energy cost:
E' = (EAE)(E) (6.11)
This is an adjustment on E for the expected inflation rate
No one really knows how the inflation rate might change in the future
But, how do you know when to change to a larger pipe size (based on a certain
sectional flow rate)?
Beginning with a smaller pipe size (e.g. selected based on maximum
velocity limits), you would change to a larger pipe size along a section of
pipeline if the difference in cost for the next larger pipe size is less than
the difference in energy (pumping) savings
Recall that the velocity limit is usually taken to be about 5 fps, or 1.5 m/s
Determine the difference in WHP between adjacent pipe sizes by equating the annual plus
annualized fixed costs for two adjacent pipe sizes:
E

(HP
s1
) + UAC
s1
= E

(HPs
2
)
+
UAC
s2
(6.12)
'
s1 s2
s2 s1
E
) UAC (UAC
WHP

=

(6.13)
The subscript s
1
is for the smaller of the two pipe sizes
The units of the numerator might be $/100 m per year; the units of the
denominator might be $/kW per year
This is the WHP (energy) savings needed to offset the annualized fixed cost
difference for purchasing two adjacent pipe sizes; it is the economic balance point
88
10. Determine the difference in friction loss gradient between adjacent pipe sizes:
|
|

\
|
=

s
s2 s1
s2 - s1
Q
WHP
102 J (6.14)
This is the head loss difference needed to balance fixed and annual costs
for the two adjacent pipe sizes
The coefficient 102 is for Qs in lps, and WHP in kW
You can also put Q
s
in gpm, and WHP in HP, then substitute 3,960 for
102, and you will get exactly the same value for J
As before, AJ is a head loss gradient, in head per 100 units of length (m
or ft, or any other unit)
Thus, J is a dimensionless "percentage": head, H, can be in m, and when you
define a unit length (e.g. 100 m), the H per unit meter becomes dimensionless
This is why you can calculate J using any consistent units and you will get the
same result
IV. Other Pipe Sizing Methods
Other methods used to size pipes include the following:
1. Unit head loss method: the designer specifies a limit on the
allowable head loss per unit length of pipe
2. Maximum velocity method: the designer specifies a maximum average
velocity of flow in the pipe (about 5 to 7 ft/s, or 1.5 to 2.0 m/s)
3. Percent head loss method: the designer sets the maximum pressure variation
in a section of the pipe, similar to the 20%P
a
rule for lateral pipe sizing
It is often a good idea to apply more than one pipe selection method and compare the
results
For example, don't accept a recommendation from the economic selection method if it
will give you a flow velocity of more than about 7 ft/s (2 m/s), otherwise you may
have water hammer problems during operation
However, it is usually advisable to at least apply the economic selection method
unless the energy costs are very low
In many cases, the same pipe sizes will be selected, even when applying different
methods
For a given average velocity, V, in a circular pipe, and discharge, Q, the required
inside pipe diameter is:
V
4Q
D = (6.15)
89
The following tables show maximum flow rates for specified average velocity limits and
different pipe inside diameters (you can easily make tables like this in a spreadsheet
application)
Liters per Second
Velocity Limit
D (mm) A (m2) 1.5 m/s 2 m/s
10 0.00008 0.1 0.2
20 0.00031 0.5 0.6
25 0.00049 0.7 1.0
30 0.00071 1.1 1.4
40 0.00126 1.9 2.5
50 0.00196 2.9 3.9
75 0.00442 6.6 8.8
100 0.00785 11.8 15.7
120 0.01131 17.0 22.6
150 0.01767 26.5 35.3
200 0.03142 47.1 62.8
250 0.04909 73.6 98.2
300 0.07069 106 141
400 0.12566 188 251
500 0.19635 295 393
600 0.28274 424 565
700 0.38485 577 770
800 0.50265 754 1,005
900 0.63617 954 1,272
1000 0.78540 1,178 1,571
1100 0.95033 1,425 1,901
1200 1.13097 1,696 2,262
1300 1.32732 1,991 2,655
1400 1.53938 2,309 3,079
1500 1.76715 2,651 3,534
1600 2.01062 3,016 4,021
1700 2.26980 3,405 4,540
1800 2.54469 3,817 5,089
1900 2.83529 4,253 5,671
2000 3.14159 4,712 6,283
2100 3.46361 5,195 6,927
2200 3.80133 5,702 7,603
2300 4.15476 6,232 8,310
2400 4.52389 6,786 9,048
2500 4.90874 7,363 9,817
2600 5.30929 7,964 10,619
2700 5.72555 8,588 11,451
2800 6.15752 9,236 12,315
2900 6.60520 9,908 13,210
3000 7.06858 10,603 14,137
3100 7.54768 11,322 15,095
3200 8.04248 12,064 16,085
3300 8.55299 12,829 17,106
3400 9.07920 13,619 18,158
D (inch) A (ft2) 5 fps 7 fps
0.5 0.00136 3.1 4.3
0.75 0.00307 6.9 9.6
1 0.00545 12.2 17.1
1.25 0.00852 19.1 26.8
1.5 0.01227 27.5 38.6
2 0.02182 49.0 68.5
3 0.04909 110 154
4 0.08727 196 274
5 0.13635 306 428
6 0.19635 441 617
8 0.34907 783 1,097
10 0.54542 1,224 1,714
12 0.78540 1,763 2,468
15 1.22718 2,754 3,856
18 1.76715 3,966 5,552
20 2.18166 4,896 6,855
25 3.40885 7,650 10,711
30 4.90874 11,017 15,423
40 8.72665 19,585 27,419
50 13.63538 30,602 42,843
Cubic Feet per Second
Velocity Limit
D (ft) A (ft2) 5 fps 7 fps
1 0.785 3.93 5.50
2 3.142 15.71 21.99
3 7.069 35.34 49.48
4 12.566 62.83 87.96
5 19.635 98.17 137.44
6 28.274 141.37 197.92
7 38.485 192.42 269.39
8 50.265 251.33 351.86
9 63.617 318.09 445.32
10 78.540 392.70 549.78
11 95.033 475.17 665.23
12 113.097 565.49 791.68
13 132.732 663.66 929.13
14 153.938 769.69 1,077.57
15 176.715 883.57 1,237.00
16 201.062 1,005.31 1,407.43
17 226.980 1,134.90 1,588.86
18 254.469 1,272.35 1,781.28
19 283.529 1,417.64 1,984.70
20 314.159 1,570.80 2,199.11
Gallons per Minute
90
Unit 7
CENTER-PIVOT SPRINKLER IRRIGATION SYSTEM
I. Introduction and General Comments
Center pivots are used on about half of the sprinkler-irrigated lands in the USA.
Center pivots are also found in many other countries.
Typical lateral length is 1,320 ft (400 m), or mile.
The lateral is often about 10 ft above the ground.
Typically, 120-ft pipe span per tower (usual range: 90 to 250 ft), often with one-
horsepower electric motors (geared down).
At 120 ft per tower, a 1,320-ft lateral has about 11 towers; with 1-HP motors, that
comes to about 10 HP just for moving the pivot around in a circle.
The cost for a mile center pivot is typically about $45,000 (about $354/ac
or $875/ha), plus about $25,000 for a corner system.
For a mile lateral, the cost may be about $75,000 (with out corner system).
In the state of Nebraska, there are said to be 50,000 installed center pivots, about 10 to
15% of which have corner systems.
The state of Texas has over 10,000 center pivots.
Center pivots are easily (and commonly) automated, and can have much lower labor
costs than periodic-move sprinkler systems.
Center pivot maintenance costs can be high because it is a large and fairly complex
machine, operating under "field" (sun, wind, dust, dirt) conditions.
The typical maximum complete rotation is 20 hrs or so, but some (120-acre pivots)
can go around in only about 6 hrs.
Running a pivot around at relatively high speeds makes the mechanical parts wear out
faster, and may also increase evaporation losses.
IPS 6" lateral pipe is common (about 6-5/8 inches OD); lateral pipe is generally 6 to 8
inches, but can be up to 10 inches for 2,640-ft laterals.
Long pivot laterals will usually have two different pipe sizes.
Typical lateral inflow rates are 45 - 65 lps (700 to 1,000 gpm).
At 55 lps with a 6-inch pipe, the entrance velocity is a bit high at 3 m/s.
Typical lateral operating pressures are 140 - 500 kPa (20 to 70 psi).
The end tower sets the rotation speed; micro switches and or cables keep other towers
aligned.
Without a corner system or end gun (see below), t/4 = 79% of the square area is
irrigated. An end-gun will irrigate additional 5% of the total area. Whereas a corner
pivot can irrigate additional 12% of the total area. Thus total area irrigated comes out
96% (79% + 5% + 12% = 96%) leaving rest of the area un-irrigated.
For a 1,320-ft lateral (without an end gun), the irrigated area is 126 acres.
For design purposes, usually ignore soil WHC (W
a
Z); but, refill root zone at each
irrigation (even if daily).
Center pivots can operate on very undulating topography.
Some center pivots can be moved from field to field.
Below are some sample center pivot arrangements.
91
Figure 7.1 Showing the layout of the center-pivot systems
92
Many farmers like extra capacity in the center pivot so they can shut off during windy times of the
day, and still complete the irrigations in time.
Some center-pivot irrigated fields have half the circle in one crop, and the other half in another
crop (or fallow) - it can be difficult to make the pivot to stop at exactly the right position when
some sections of the circle are irrigated differently than others.
Center pivots on rolling terrain usually have pressure regulators at each sprinkler; otherwise, there
is too much pressure variation among sprinklers, and the application uniformity is low.
Some engineers claim that center pivots can have up to about 90%application efficiency.
A stationary sprinkle system applies water at constant rate, while the water
application rate at a point varies continuously with time during application by any moving
sprinkle system. As the system approaches a point, the application rate starts from zero,
increases to a maximum, and then decreases to zero again as the moving system passes over
the location. The water application under a center-pivot lateral is even more complicated. The
application rate varies along the lateral from a low value near the pivot to higher values at the
outer end (Figure 2). This variation in the application rates is essential due to two reasons: i)
area of the circular irrigated field increases with square of the radius, ii) time of water
application at a point decreases with increasing distance from the pivot as shown in the above
figure. Thus instantaneous application rate patterns are different at different locations along
the lateral. Since the application rate is the highest near the end of a center-pivot system and
runoff may occur at this location if it exceeds the soil intake rate.
Figure 7.2 Measured water application rates and application times along a center-
pivot lateral.
Keller and Bliesner (1988) derived water-application rate profiles along different
points along the lateral of a center-pivot system as shown in Figure 8.8. The rate is maximum
near the end of the lateral and it decreases inward as explained earlier. Water application rate
for an elliptical profile is given by:
xj
aj
xj aj
aj
e e
j
I
T
I T
T
Q dR
I
4 4 60 /

= = = (7.1)
93
Where,
I
j
= average application rate required at radial distance r
j
, mm/hr (in/hr)
d = gross depth of water required per irrigation, mm (in)
T
aj
= application time at radial distance rj, min
I
xj
= peak application rate at radial distance rj, mm/hr (in/hr)
Q
e
= ratio of water effectively discharged through sprinklers to total system
discharge,
R
e
= effective portion of water discharged from sprinklers, most of which reaches
the soil-plant surface in decimal
Figure 7.3 Water application rates at different points along a center-pivot lateral
(Source: Keller and Bliesner, 1990)
94
Thus average application rate is a function of the radial distance r
j
from the pivot and it is
proportional to the length of circular travel path at r
j
divided by the width of the application
profile. It is given by:
T
d
.
wj
r 2
I
pi j
j
= (7.2)
Combining the above equations gives:
2
e e b
j
j
L
O R Q K
wj
r 2
I = = (7.3)
Where
I
j
= average application rate at radius r
j
, mm/hr (in/hr)
r
j
= radial distance from pivot to point under study, m (ft)
d
pi
= average peak gross infiltration required per day, mm (in.)
w
j
= width of stationary application patterns at r
j
, m (ft)
T = average daily operating time, hr
K = conversion constant 1146 for metric units 30.6 for English units)
L = radius of the irrigated field, m (ft)
Corner Systems
Corner systems are expensive, and they don't irrigate the whole corner.
Corner systems on center pivots have buried cables so that sprinklers in the corner
extension arm turn on and off individually (or in groups) as the arm swings out and then
back in again.
The cables are buried about 5 ft (1.5 m) below the surface to prevent damage during
tillage operations (such as chiseling or "ripping").
The end tower on the main part of the lateral has an antenna facing downward to detect
the signal from the buried cable, and to know when to begin to swing the extension arm
outward.
95
Figure 7.4 A view of center-pivot system with a corner system
Usually, groups of sprinklers are controlled by separate valves in the extension arm
because the application rate and uniformity would differ too greatly from the desired
values if all the corner system sprinklers turned on (and off) simultaneously.
There may be a control box at both the pivot point (center of field) and at the joint
where the extension arm connects to the main lateral; either box can be used to
program and control the pivot (e.g. to shut it down).
Application uniformity in the corners is not as high as in the main circle area.
End Guns
Some pivots have an end gun that turns on in the corners, in which all other sprinklers
shut off via individual solenoid-actuated valves. The pivot stops in the corner while the end
gun runs for a few minutes. Others just slow down in the corners, turning on an end gun, but
leaving the other sprinklers running (at lower discharges); in these cases, a pressure
transducer might be used at the end of the lateral to indicate when the gun turns on (resulting
in a pressure drop), thus knowing when to slow the pivot. The figure below shows an
example of the approximate corner areas than might be irrigated by an end gun.
96
Figure 7.5 (a). Area irrigated by the main system and end gun.
The end-guns turn on when the lateral is in pointing toward a corner area (which is
when they should turn on) by angle measurements at the pivot point. In other cases, an
end gun on a center pivot lateral might be set to run all of the time (i.e. all the way
around the circle), effectively increasing the diameter of the irrigated area, but not
dealing with un-irrigated corner areas.
Figure 7.5 (b). Area irrigated by the main system and end gun.
97
Safety Switches
Most center pivots have one or more safety switches.
Some have switches to shut the whole machine off if any tower gets too far out of
alignment.
Some also have safety switches to shut them off if the temperatures gets below
freezing (ice builds up and gets heavy, possibly collapsing the structure) - typically
shuts off at 32F and turns back on at 40F.
With drop-down sprayer sprinklers, ice build-up is not so problematic.
Some have safety switches connected to timers: if a tower has not moved in a
specified number of minutes, the system shuts down.
There may also be safety switches associated with the chemical injection equipment at the
lateral inlet location.
Figure 7.6 A view collapsed center pivot system
Drop Down Sprinkler
- Almost all new center pivots have drop-down, low-pressure (20 - 25 psi) sprinklers.
- Many older center pivots, which formerly had above-lateral impact sprinklers (50 - 60
psi) have been retrofitted with drop-down, low-pressure sprinklers. These features
help reduce energy (pumping) costs and reduce evaporation and wind drift.
- Gooseneck tubes are used to convert older pivots to drop-down sprayers (sprinklers) -
they connect to the top of the center pivot lateral pipe at the same locations as the
previously-used impact sprinklers. New systems, designed for drop-down sprayers,
also use the gooseneck tubes because they help reduce the amount of sand and other
sediments, which might be in the lateral pipe, from moving through (and wearing out)
the sprinklers.
98
- Center pivot laterals sometimes have manual flush valves (gate or butterfly) at the end
to periodically wash sediments out of the pipe. Some of the drop-down tubes are
polyethylene, others are PVC, and still others are galvanized steel.
- The steel drop-down tubes have the advantage of rigidity and do not sway nearly as
much as polyethylene hoses (which usually have weights at bottom) under windy
conditions.
Figure 7.7 System with sprinkler heads below the lateral
II. System Capacity
The general center pivot design equation for system capacity is based on Eq.
5.4 from the textbook:
pa
f d
s
TE k
k U R
fT k
d R
fT
Ad
K Q
1
2
1
2
'
= = = (7.4)
Where, K is 2.78 for metric units and 453 for English units k
l
is (10,000 m
2
/ha)/(2.78 TI) =
1,145 for metric units k
l
is (43,560 ft
2
/acre)/(453 %) 30.6 for English units k
f
is the peak
period evaporation factor (Table 14.1) A is area (ha or acre) d is gross daily application depth
(mm or inch) d' is defined in the equation below (mm or inch) f is frequency in days per
irrigation.
T is operating time (hrs/day) R is the effective radius (m or ft) U
d
is the peak-use ET rate of
the crop (mm/day or inch/day) Q
s
is the system capacity (lps or gpm).
99
The gross application depth, d, is equal to d
n
/E
pa
, where E
pa
is the design application
efficiency, based on uniformity and percent area (pa) adequately irrigated.
The operating time, T, is generally 20-22 hrs/day during the peak-use period.
R is the effective radius, based on the wetted area from the center pivot.
The effective radius is about 400 m for many center pivots.
R L + 0.4w, where L is the physical length of the lateral pipe, and w is the wetted
diameter of the end sprinkler.
This assumes that approximately 0.8 of the sprinkler radius beyond the lateral pipe is
effective for crop production.
Note that, for center pivots, Q
s
is proportional to U
d
, and d and f are generally not
used, which is similar to drip irrigation design.
Usually, for center pivot design, f = 1 day. So, if you use f > 1 day, you must also
increase d or d' to accommodate f number of days.
III. Gross Application Depth
If a center pivot is operated such that the water holding capacity of the soil is
essentially ignored, and water is applied frequently enough to satisfy peak-
use crop water requirements, then use d
n
/f = U
d
, and
e e pa
d f
pa
d f
s
O R DE
U k
E
U k
d = = ' (7.5)
where d' is the gross daily application depth (mm/day or inches/day); and k
f
is a peak-
use period evaporation factor, which accounts for increased soil and foliage
evaporation due to high frequency (daily) irrigation.
When LR > 0.1, the LR can be factored into the equation as:
pa
d f
e e pa
d f
E LR
U k
O R DE LR
U k
d
) 1 (
9 . 0
) 1 (
9 . 0
'

= (7.6)
Which is the same as Eq. 14.1b from the textbook, except that DE
pa
, R
e
and O
e
are all
as fractions (not percent).
Values of k
f
can be selected for the peak period from Table 14.1 of the textbook for
varying values of frequency, f.
Values for non-peak periods can be computed as described in the textbook on page
314:
0 . 1
/ ) 100 (
' / ) ' 100 (
) 1 ( k
'
f
+

=
PT PT
PT PT
k
f
(7.7)
100
Where, k
f
and PT are for the peak-use period (Table 14.1), and k'
f
and PT are the
frequency coefficient and transpiration percentage (PT) for the non-peak period.
ET
T
PT = (7.8)
PT and PT' can be thought of as the basal crop coefficient (Kc
b
), or perhaps
Kcb - 0.1 (relative to alfalfa, as per the note in Table 14.1).
It represents the transpiration of the crop relative to an alfalfa reference.
IV. Water Application along the Pivot Lateral
A major design difficulty with a center pivot is maintaining the application rate
so that it is less than the intake rate of the soil.
This is especially critical near the end of the lateral where application rates
are the highest.
As one moves along the center pivot lateral, the area irrigated by each unit length of
the lateral (each 1 ft or 1 m of length) at distance r from the pivot point can be
calculated as:
r r r a 2 ) 5 . 0 ( ) 5 . 0 (
2 2
= + = (7.9)
This is equal to the circumference at the radial distance r.
The portion of Q
s
(called q) which is applied to the unit strip at distance r is:
2 2
2 2
R
r
R
r
A
a
Q
q
s
= = =

(7.10)
2
2
R
rQ
q
s
= (7.11)
where q can be in units of lps per m, or gpm per ft
This gives the amount of water which should be discharging from a specific unit
length of lateral at a radial distance r from the pivot point.
The q value at the end of the lateral (r = R) per ft or m is:
Use q to select the nozzle size, where q
nozzle
= q S
e
101
UNIFORMITY OF A CENTER PIVOT
I. Introduction
The calculation of an application uniformity term must take into account the irrigated
area represented by each catch container.
It is more important to have better application uniformity further from the pivot point
than nearer, because the catch containers at larger distances represent larger irrigated
areas.
If the catch containers are equally spaced in the radial direction, the area represented
by each is directly proportional to the radial distance.
II. Equation for Center Pivot CU
The equation for CU proposed by Heermann and Hein is (ASAE/ANSI S436):
(7.12)
Where, CU is the coefficient of uniformity; d
i
is the depth from an individual container; r
i
is the radial distance from the pivot point; and, n is the number of containers.
First calculate the summations
) ( 1
1
i i
n n
i
r d i and

=
=
(7.13)
Then, perform the outer summation to determine the CU value.
That is, don't recalculate the inner summation values for every iteration of the outer
summation - it isn't necessary.
It is usually considered that a center pivot CU should be greater than 85%.
Note that nobody ever puts a whole grid of catch cans in a center pivot field.
If the radial distances, n, are equal, the sequence number of the can (increasing with
increasing radius) can be used instead of the actual distance for the purpose of
calculating application uniformity.
Consider the following two figures:
102
Figure 7.8 Illustration of data collection for evaluating the system
103
III. Standard Uniformity Values
- You can also calculate the "standard" CU or DU if you weight each catch value by
multiplying it by the corresponding radial distance.
- To obtain the low , rank the un-weighted catches, then start summing radii
(beginning with the radius for the lowest catch value) until the cumulative value is
approximately equal to of the total cumulative radius.
- This may or may not be equal to of the total catch values, because each catch
represents a different annular area of the field.
- Finally, divide the sum of the catches times the radii for this approximately area by
the cumulative radius.
- This gives the average catch of the low ...
- Don't rank the weighted catches (depth x radius) because you will
mostly get the values from the low r values (unless the inner catches are
relatively high for some reason), and your answer will be wrong.
- Don't calculate the average of the low like this. (because the lowest
of the catches generally represents something different than of the
irrigated area):
- Actually, the equation at the right is all right, except for the value "n/4", which is
probably the wrong number of ranked values to use in representing the low .
- You can set up a table like this in a spreadsheet application.
Ranked Center Pivot Catches
Radius, r Cumulative
r
Depth, d d*r Cumulative
d*r
smallest
largest
Totals:
Note that when you rank the depths, the radius values should stay with the same depth
values (so that the radius values will now be "unranked"; all mixed up).
To get the average weighted depth for the whole pivot area, divide the total "Cumulative
d*r" by the total "Cumulative r" (column 5 divided by column 2).
104
Find the row corresponding closest to of the total "Cumulative r" value, and
take the same ratio as before to get the weighted average of the low % area.
Look at the example data analysis below:
Radius,
r
Cum. r Depth,
d
d*r Cum.
d*r
120 120 0.52 62.6 62.6
900 1,020 0.95 851.9 915
160 1,180 1.29 205.8 1,120
340 1,520 1.31 445.6 1,566
1000 2,520 1.46 1,456.
3
3,022
1040 3,560 1.46 1,514.
6
4,537
240 3,800 1,48 355.3 4,892
800 4,600 1.50 1,203.
9
6,096
860 5,460 1.53 1,315.
0
7,411
480 5,940 1.58 757.3 8,168
1280 7,220 1.58 2,019.
4
10,188
980 8,200 1.60 1,569.
9
11,758
540 8,740 1.63 878.2 12,636
360 9,100 1.65 594,2 13,230
460 9,560 1.67 770.4 14,000
880 W.440 1.70 1,495.
1
15,496
320 10,760 1.72 551.5 16,047
1140 11,900 1.75 1,992.
2
18,039
1160 13,060 1.75 2.027
2
20,067
280 13,340 1.82 509.7 20,576
720 14,060 1.82 1,310.
7
21,887
1300 15,360 1.82 2,366.
5
24,253
200 15,560 1.84 368.9 24,622
420 15,980 1.84 774.8 25,397
440 16,420 1.89 833.0 26,230
1020 17,440 1.89 1,931. 28,161
105
Notice that the depth values (3
rd
column) are ranked from low to high.
Notice that the maximum value of cumulative r is 44,220 & maximum cumulative d*r
is 87,784. Then, the weighted average depth for the entire center pivot is equal to
87,784/44,220 = 1.985 (whatever units).
One quarter of 44,220 is equal to 11,055 which corresponds most closely the row in
the table with depth = 1.72. For the same row, divide the two cumulative columns (Col
5/Col 2) to get 16,047/10,760 = 1.491, which is approximately the average of the low
.
Finally, estimate the distribution uniformity for this data set as:
% 75
985 . 1
491 . 1
100 ~ |
.
|

\
|
= DU (7.14)
Note that in this example, the average of the low was, in fact, based on approximately the
first n/4 ranked values.
Consider the weighed catch-can data plotted below:
Figure 7.9 Accumulative water application depth along the lateral
As in any application uniformity evaluation, there is no "right" answer. The results are
useful in a comparative sense with evaluations of other center pivots and other on-farm
irrigation systems.
However, a plot of the catches can give indications of localized problems along the
center pivot radius.
106
IV. The Field Work
It may take a long time for the full catch in containers near the pivot point, and
because these represent relatively small areas compared to the total irrigated area, it is
usually acceptable to ignore the inside 10% or 20% of the radius.
The pivot quickly passes the outer cans, but takes longer to completely pass the inner
cans, so you can collect the data from the outer cans sooner.
The pivot should not be moving so fast that the application depth is less than about 15
mm.
Catch containers can be placed beyond the physical length of the lateral pipe, but if
they are so far out that the catches are very low, these can be omitted from the
uniformity calculations.
Catch containers should be spaced in the radial direction no further than about 30% of
the average wetted diameter of the sprinklers.
There is often an access road leading to the pivot point for inspection, manual
operation, maintenance, and other reasons.
If the crop is dense and fairly tall (e.g. wheat or maize) it will be difficult to perform
the evaluation unless the cans are placed on the access road.
Otherwise, you can wait until the crop is harvested, or do the test when the crop is still
small.
Some people recommend two radial rows of catch cans, or even two parallel rows, to
help smooth out the effects of the non-continuous movement of towers (they start and
stop frequently to keep the pivot lateral in alignment).
Some have used troughs instead of catch cans to help ameliorate this problem.
Note that if the field is sloping or undulating, the results from one radial row of catch
cans may be quite different from those of a row on another part of the irrigated circle
See Merriam and Keller (1978).
LINEAR MOVE SYSTEMS
I. Introduction
Mechanically, a linear move system is essentially the same as a center pivot lateral, but
it moves sideways along a rectangular field, perpendicular to the alignment of the
lateral pipe
The variation of flow rate in a linear move lateral is directly proportional to distance
along the lateral pipe, whereas with center pivots it is proportional to a function of the
square of the distance from the pivot point.
A center or end tower sets the forward speed of the machine, and the other towers just
move to keep in line with the guide tower (this is like the far end tower on a center
pivot).
Usually, each tower is independently guided by cables and micro-switches as for a
center pivot - this keeps the lateral pipe in a straight line (aligned with itself).
107
Alignment with the field is usually not mechanically "enforced", but it is "monitored"
through switches in contact with a straight cable along the center of the field, or along
one end of the field.
The center tower has two "fingers", one on each side of the cable; usually slightly
offset in the direction of travel (they aren't side by side). The fingers should be in
constant contact with the cable - if one is lifted too far a switch will be tripped,
shutting the system down (because the whole lateral is probably getting out of
alignment with the field).
If the cable is broken for any reason, this should also shut the system down because
the fingers will lose physical contact.
If the lateral gets out of alignment with the field and shuts off, it will be necessary to
back up one side and or move the other side forward until it is in the correct position.
This can involve electrical "jumps" between contacts in the control box, but some
manufacturers and some installers put manual switches in just for this purpose.
Some linear moves are fitted with spray nozzles on drop tubes or booms.
If they are spaced closely along the lateral, it may be necessary to put booms out
beyond the wheels at tower locations, either in back of the lateral or on both sides of
the lateral.
Water Supply
Water is usually supplied to the lateral via:
1. a concrete-lined trapezoidal-sectioned ditch, or
2. a flexible hose (often 150 m in length), or
3. automatic hydrant coupling devices with buried mainline
Hose-fed systems require periodic manual reconnection to hydrants on a mainline - it is
kind of like a periodic move system, and you have to ask yourself whether the Wnear
move machine is worth the cost in this case.
With the automatic hydrant coupling machines (see Fig. 15.3) there are two arms with
pipes and an elbow joint that bends as the Wnear move travels down the field. The two
arms alternate in connecting to hydrants so as neither to disrupt the irrigation nor the
forward movement of the machine. These are mechanically complex.
The advantage of hose-fed and automate coupling Wnear moves is that you don't need
to have a small, uniform slope in the direction of travel because water is supplied from a
pressurized mainline instead of an open channel.
On ditch-fed systems there can be a structure at the end of the field that a switch on the
linear move contacts, shutting down the pump and reversing the direction of movement so
that it automatically returns to the starting end of the field.
The advantages and disadvantages of the ditch-feed system are:
108
Pros
Low pressure (energy) requirement
Totally automated system
More frequent irrigations than hose-fed, since no one needs to be available to
move the hose
Cons
Trash and seeds and sediment pass through screen and may plug nozzles.
The pump must be on (move with) the lateral, causing extra weight.
Should have uniform slope along the lateral route.
Pros and Cons Compared to a Center Pivot
Pros
Easy irrigation of a rectangular field (important if land is expensive, but not
important if land is cheap and water is scarce).
Application rate is uniform over length of lateral, rather than twice the average
value at the end of the center pivot.
No end gun is required.
Cons
The lateral does not end up right back at the starting point immediately after
having traversed the irrigated area you either have to deadhead back or irrigate in
both directions.
May be more expensive than a pivot due to extra controls, pump on ditch feed, or
more friction loss in the flexible feed hose (the hose is fairly expensive).
109
110
111
112
113
114
Unit 8
TRICKLE IRRIGATION
INTRODUCTION
Trickle irrigation is a relatively new irrigation method developed in middle of the
twentieth century. In this method water is applied slowly through tiny openings called
emitters or applicators located on water delivery lines near the plants. Most emitters are
placed on the ground surface, but these can be buried as well. One or more emitters are used
to apply water to a plant depending upon water requirement of the plant and some other
factors which will be discussed later in this Chapter. This method is extremely useful in water
scarcity areas. Micro irrigation and localized irrigation are other names often used for trickle
irrigation. Trickle irrigation is further classified as drip, spray, bubbles, and hose-basin
irrigation techniques depending upon the type of emitter used to apply water as discussed in
the next section. Some writers use the words 'trickle' and 'drip' synonymously which is
technically not correct as drip is one type of trickle (Keller and Blissner, 1990).
ADVANTAGES
Trickle irrigation has become popular in water scarcity areas due to its high
efficiency. Efficiency of this system may be 90 percent or more if properly designed,
managed and operated. Some of the main advantages of this method are:
i) Water Saving: High efficiency of the system and partial wetting of the soil surface
result in much saving of irrigation water. Furthermore water savings are also achieved by
light water application to young plantings and young orchards because with conventional
methods heavy irrigations result in more wastage. Surface water runoff and deep percolation
are also avoided or reduced in trickle irrigation. Small stream flows can be more conveniently
used with this system.
ii) Control of Weed Growth: Weed growth is minimized due to limited soil surface
wetting. Field operations are also less interrupted due to partial surface wetting.
iii) Fertilizer Application: Dissolved fertilizers and other chemicals can be applied more
effectively as they are applied near the plants. This will also save additional cost of fertilizer
application. The fertilizers may be injected by venturi or injection pump (positive
displacement) which may be electric or water driven.
iv) Labour Saving: This system can be easily automated which will result in saving of
irrigation labour.
115
v) Saline Water: Irrigations are applied frequently at high moisture content that keeps
the soil solution more dilute. Due to this more salty water can be used by this method as
compared to other irrigation methods.
vi) Levelling: A field with rolling topography can be easily irrigated with trickle
irrigation, thus saving cost of land levelling.
vii) Crop Yield: The plants are not subjected to water stress due to frequent water
application. This may enhance plant growth and crop yield.
DISADVANTAGES
High Initial Cost: Like sprinkler system, initial cost of trickle system is also high.
Therefore use of trickle irrigation is suitable for high value cash crops, vegetables and
orchards.
Clogging: Filtered water is essential to avoid clogging. Emitters are subjected to clogging
by mineral particles and organic matter, present in irrigation water. Such particles must be
removed by filtration before water enters the pipe networks. Water may be drained from the
pipes or it should be chemically treated to avoid clogging by algae and rust.
Salt Accumulation: Salt accumulated on the soil surface and around the wetted perimeter
may become injurious if use of saline water is prolonged. Light rain shower may leach down
these salts into active root zone under such condition, irrigation should be applied as planned
to dilute and leach any salts out of the root zone.
TYPES OF TRICKLE IRRIGATION
Drip Irrigation: In this method, water is applied slowly like drops by single or multi outlet
emitters. The emitters may be spaced apart in case of point emitters whereas they are more
closely spaced for line source emitters. Due to small opening of the emitters, special care
should be taken to avoid clogging.
Bubbler Irrigation: In bubbler irrigation water is applied through small diameter tube (less
than 10 mm) which is connected to the lateral line to basin of individual tree. Application rate
may be varied by varying diameter or length of the bubbler tube. Higher discharge of this
system may cause surface runoff and deep percolation.
Spray Irrigation: Spray irrigation resemble micro-sprinklers in which small sprinklers are
used to spray as a mist on the soil surface. This system may be designed to wet all or part of
the soil surface. Discharge rate is usually less than 2 lpm. In this system evaporation and drift
losses are more than the other type of trickle systems.
Subsurface Irrigation: In subsurface irrigation the emitters of a drip system are placed
below the land surface. The subsurface irrigation is different than sub-irrigation in which
water table is raised by blocking outlet of tile drainage system or by some other arrangement.
116
COMPONENTS OF A TRICKLE IRRIGATION SYSTEM
Basic components of a typical trickle irrigation system consists of pump, control head,
main and sub-main lines, lateral lines, emitters, fittings, valves, pressure and flow measuring
devices. Main components of a typical trickle system are shown in Fig. 8.1. The main line is
connected to the water supply source and it supply water to the manifolds (sub-mains) and
laterals. The manifold is the part of pipe network between the main line and laterals. It is
normally burried but can be laid on surface as well. Lateral lines are 10 to 25 mm diameter
polyethylene or PVC hose or tubing and they supply water to the emitters.
Control Head: The components of control head comprise filters, chemical injection, tank
and injector flow measuring and pressure regulating devices. Good filtration is essential to
avoid clogging of emitters which causes poor water distribution. Screen and graded gravel-
sand filters and settling basins are commonly used for filtering and keeping contamination
out of the system. Positive displacement pump with piston-type injector or a venturi device
are commonly used to inject liquid fertilizers and other chemicals (insecticides acids,
algacides, etc) in irrigation in the trickle irrigation system.
Figure 8.1 Primary components of a trickle irrigation system.
Main and Sub-main pipelines: Filtered water is supplied through the main pipeline to the
sub-mains (manifold), which serves a group of laterals. For trickle irrigation, the operating
pressure is lower as compared to sprinkler; therefore relatively cheaper pipe can be used.
Normally these pipes are of polyvinyl-chloride (PVC) and sometime asbestos-cement pipe is
used for the main line. The sub-main is of polyethylene or PVC with 20 to 80 mm diameter.
117
There should be values on the mainline and sub-main to regulate the flow and for periodic
flushing.
Lateral Lines: The lateral lines are usually connected to the sub-main but in some cases
directly to the main line. The lateral lines are made of black polyethylene (PE) plastic ranging
in diameter from 8 to 25 mm. For good uniformity, pressure variation in a lateral line should
not exceed 10 percent.
Valves: Different types of valves are used for proper operation and management of a trickle
irrigation system. Manifold valves control water supply from the main pipeline to the
manifolds. Lateral or header valves are located on the manifold risers that supply water to
each lateral or header. Flush-out valves are attached at the ends of main lines, manifolds, and
laterals to flush the pipe network periodically.
Emitters: Emitters are water applicator devices used in different types of trickle systems.
They are designed to dissipate pressure and to discharge a small uniform flow at a constant
rate. Ideal emitters should have a constant flow over some variation of pressure and also
some means of flushing to reduce clogging. Types of emitters include long-path emitters,
orifice emitters, vortex emitter and flushing emitters. The later emitters are further of
continuous-flushing type if they permit continuous passage of large solid particles while
operating or compensating emitters which have constant discharge over wide range of
operating pressure. In addition to the single outlet emitters, the multi-outlet emitters supply
water to two or more points through auxiliary tubing.
Point-Source and Line-Source Applications: When emitters are widely spaced on the
lateral tubing, wetting pattern of an individual emitter is of elliptical shape as shown in Fig.
8.2.
Figure 8.2 Wetting profiles for equal volumes (12 gal) of water applied at three different application
rates on on a dry soil (Source: Keller and Bliesner, 1990, p.458)
60 45 30 15 15 30 45 60
Width, cm
Legend
o 1gph for 12 h
2 gph for 6 h
4 gph for 13 h
Emitter
1
5
3
0
4
5
6
0
118
This is called point-source application and emitters are spaced 1 m or more. In line-source
application, the emitters are closely spaced at intervals of about 0.6 m on small diameter hose
(less than 25 mm). The line-source tubing may be single chamber tubing or double-chamber
tubing. In double chamber tubing larger holes at longer intervals are made in the inner wall
while smaller diameter orifices are punched at 0.15 to 0.6 m in the outer wall. Porous-wall
hose has small holes uniformly spaced through which water oozes under pressure.
There is a lateral for each row of crops or plants. But for some horticultural crops, one
line may serve two crop rows. On the other hand two lateral lines may be need for each plant
row if emitter flow is small and water need of plants is high. Special layout of emission
points such as zigzag, pigtail or multi-outlet emitters are used for widely-spaced trees as
shown in Fig. 8.3.
Small water application devices used to spray or mist water are called sprayers. They
discharge a uniform spray of water to cover an area of 1 to 10 m
2
.
System Layout: The pipe network layout of a trickle system should be in such a way as to
deliver water to the emitters at the designed pressure. The system should be designed such
that a part of it can be operated at a time if so required. On level field with mild slope, the
manifold is placed in middle of the field in order to place the laterals on both side of the
manifold as shown in Fig. 8.4. If the slope is high, move the manifold uphill in order to offset
the excessive pressure drop. Gain in pressure due to elevation difference will allow to use
longer lateral on the down hill side of the manifold as shown in Fig. 8.5.
Emission Uniformity: Distribution of water in a trickle irrigation system is measured by
emission uniformity, EU which can be calculated from field data as:
EU = 100 q
n
/q
a
(8.1)
Where
q
n
= average discharge of the lowest one-fourth of the data set (emitter discharges),
L/h (gph)
q
a
= average discharge rate of all the emitters measured in the field, L/h (gph)
A system with EU greater than 90% is excellent, between 90 to 80% good, 80 to 70% fair
while less than 70% is considered poor (Merriam and Keller, 1978, page 137).
Wetted area
119
Figure. 8.3 Various Emission Point Layouts for Wide-spread Tree Plants.
(Source: Keller and Bliesner, 1990, page 436).
S
p
S
p
Tree
S
e
Wetted area
Lateral width
Emitters
W
Shaded area
S
r
S
1
=

S
r
A. Single lateral for each row of trees (3 emitters/tree, P
w
= 30%)
Area per
Tree
Wetted
Area
S
r
B. Double laterals for each
Tree row (P
w
60%)
C. Zigzag lateral for each
Tree row.
Emitter
Estimation
Points
D. Pigtail with 4
emitters
Per tree. (P
w
40%)
F. Multiexit 6 outlet emitter
with distributuin tubing
(P
w
60%)
S
p
120
Figure 8.4 A typical layout for Trickle Irrigation System.
Figure 8.5 A subunit with manifold placed uphill from middle of the field.
I
II
Subunit
Subunit
Subunit
IV
Subunit
III
Manifold
Control Valve
Water supply &
Control valves
Lateral
Emitter
Pressure regulator
Header
Laterals
Connection with
mainline
S
l
o
p
e
Slope
121
Wetted Area: Trickle irrigation does not wet the whole soil surface. Only the soil surface
near the emitter is wetted which depends on rate and volume of the emitters, slope and soil
characteristics. The wetted area on the soil surface is usually small and it is maximum at 15 to
30 cm depth beneath the emitters. The wetted area, P
w
is defined as the average wetted area at
15 to 30 cm depth divided by the total surface area of the field. The wetted area of a trickle
irrigation system is usually between 30 to 70 percent. Apart from other factors, the wetted
area also depends on rainfall, type of crop and plants grown. Higher rainfall will give better
crop production for same wetted area and vice versa.
Estimating Wetted Area: The wetted area is the average horizontal area wetted in top 15 to
30 cm as explained earlier. The wetted area can be computed from the following equation for
straight single line lateral (Keller and Bliesner, 1990):
100
100 /
P
w

d r p
e P
P S S
w S N
(8.2)
Where, P
w
= percentage of land wetted along a horizontal strip 15 to 30 cm below the land
surface, N
p
= number of emitters per tree, w is wetted width in m, S
p
= is the distance
between plants in the row, m, and S
r
= is the distance between plant rows, m.
The above equation is valid if Se Se where Se is the optimal emitter spacing,
which is normally 80% of the wetted diameter. If Se > Se
-
, then replace S
e
in the above
equation by S.
For double-lateral system and other layouts shown in Fig. 8.3, the lateral should be laid S
e

apart to maximize the wetted area which can be estimated as:


100
100 /
2 / ) ( N
P
'
e
w


d r p
e e
P S S
w S S
(8.3)
Example 8.1: Tree spacing in a garden is 3 m with 5 m row to row distance and 3 emitters
per tree. Discharge of emitter is 4 litres/hr. Calculate wetted area for a) A single row of
emitters & b) Zigzag layout as shown in Fig. 8.3C. Measurement shows wetted width is 1.5
m and S
e
= 1.2 m.
a) Using equation for single lateral, N
p
= 3
36% 100
5 3
1.5 1.2 3
P
w

122
b) For the zigzag layout, N
p
= 4
62.4% 100
5 3
2 / ) 5 . 1 2 . 1 ( 1.2 4
P
w

WATER REQUIREMENTS
Crop water requirements for trickle irrigation are different than the convential crop
water needs due to partial wetting of the soil surface. Under trickle irrigation evaporation
from soil surface is less due to its partial wetting. Water use under trickle irrigation mostly
depends on transpiration due to minimum soil evaporation. Transpiration under trickle
irrigation can be estimated as (Keller and Bliesner, 1990):
T
d
= U
d
[0.1 (P
d
)
0.5
] (8.4)
Where T
d
= average daily transpiration rate during the peak use month for a particular crop
under trickle irrigation, mm/day, U
d
= Conventially estimated average daily water
requirements during the peak use month for a mature crop, mm/day, and P
d
= percentage of
soil surface area shaded by the crop at midday in percent. This can be estimated by inspection
at noon and it may vary from some low value to a maximum of 80% of mature orchard.
Seasonal transpiration can be estimated by replacing U
d
with the total estimated
seasonal consumptive use as:
T
s
= U [0.1(P
d
)
0.5
] (8.5)
Where T
s
= seasonal transpiration under trickle irrigation, mm and U = conventially
estimated seasonal consumptive use for the mature crop, mm.
Seasonal Water Requirement
Natural rainfall fulfils some of crop water needs therefore net irrigation requirement
should be decreased accordingly. Thus net seasonal irrigation depth, D
n
is given by:
D
n
= (U R
e
M
i
) [0.1 (P
d
)
0.5
] (8.6)
Where R
e
= effective precipitation during the growing season, mm and M
s
= initial soil
moisture stored in the root zone, mm if any
Depth per Irrigation: Trickle irrigation usually wets the soil surface partially. Therefore,
irrigation depth is calculated as:
Z
P
w
a x
W
100 100
MAD
d (8.7)
123
Where d
x
= maximum net depth of water to be applied per irrigation, mm, MAD =
management allowed deficit in percent, W
a
= available water holding capacity of the soil,
mm and Z = plant root depth, mm.
Net water depth can be calculated as:
d
n
= T
d
f (8.8)
Where d
n
= net water depth applied per irrigation to meet crop water needs, mm, f = actual
irrigation interval (frequency), days, and T
d
= average daily transpiration during peak-use
period, mm.
Leaching Requirement: Salinity may develop if saline water is used for longer time in arid
and semi-arid regions where rainfall is insufficient to wash salts below the root zone. Under
such condition, additional water is applied to leach salts out of the active root zone. However,
it is not essential to apply extra water for leaching with each of irrigation. The following
equation may be used to estimate leaching requirement.
dw
iw
N n
N
n n
n
EC
EC
L D
L
L d
L

) ( ) (
LR
t
(8.9)
Where
LR
t
= leaching requirement ratio under trickle irrigation,
d
n
= net depth of application per irrigation to meet consumptive use requirements,
mm
D
n
= net annual or seasonal irrigation depth to meet consumptive use requirements,
mm
L
n
= net leaching requirement for each irrigation, mm
L
N
= net annual leaching requirement, mm
EC
w
= electrical conductivity of the irrigation water, dS/m (mmhos/cm), and
ECd
w
= electrical conductivity of the drainage (deep percolation) water, dS/m.
Gross Irrigation Requirements: Gross irrigation depth and volume required can be
estimated from net water requirement, water required for leaching and irrigation efficiency of
the system. Irrigation efficiency in a trickle system mainly depends on application uniformity
(emission uniformity) if losses due to leaks, runoff and deep percolation are not much. The
application efficiency of low-quarter, E
q
is the average of the lowest one-fourth volume
applied to the over all average of the measured data set. If losses are negligible then E
q
is
equal to the measured EU.
Gross Depths and Volumes: Total or gross amount of water per irrigation, d may be
calculated knowing irrigation efficiency or emission uniformity. If leaching requirement is
small (LR
t
= 10%) then it may be neglected and irrigation depth is given by:
124
100 /
d
d
n
EU
T
r
(8.10)
or
100 /
T
d
d
EU
T
r

On the other hand if LR
t
> 10% then
) 0 . 1 (
d 100
d
n
t
LR EU
(8.11)
t
LR EU ) 0 . 1 (
T 100
d
d


Where,
d = gross depth of application per irrigation, mm
d' = maximum gross daily irrigation requirement, mm
T
r
= peak-use period transmission ratio
EU = emission uniformity, %
LR
t
= leaching requirement under trickle irrigation.
Gross volume of water required per day per plant is given by:
G = K d/f S
p
S
r
Or
G = K d S
p
S
r
(8.12)
Where,
G = gross volume of water required per plant or per unit length of lateral per day, L/day
K = conversion constant which is 1.0 for metric units
Similarly gross seasonal volume of water can be estimated for TR > 1.0/(1.0- LR
t
) as:
100 /
D
V
n
s
EU
T
K
A
r
(8.13)
for T
R
1.0/(1.0 LR
t
)
) 0 . 1 (
D 100
V
n
s
t
LR EU K
A

(8.14)
System Discharge: The total system discharge or capacity depends on the maximum number
of emitters operating at a given time. It can be calculated as:
125
r p s
a
s
S S N
q
p
s
N
N
A
K Q (8.15)
For uniformity spaced laterals that supply uniformly spaced emitters, the Q
s
is given by:
l e s
a
s
S S N
q
N
A
K Q
s
(8.16)
Where,
Q
s
= Total system capacity, L/s
K = Conversion constant, 2.778 for metric units
A = Field area, ha, and
N
s
= Number of operating stations
For line-source tubing where the discharge per unit length is given, replace q
a
/S
e
in the last
equation with discharge per meter of the lateral.
Operating Time: Operating time per season can be calculated as:
s
s
Q
V
K O
t
(8.17)
Where, O
t
= average operating time per season, hr, and K = conversion constant = 2.778 for
metric units.
Net Application Rate: The net application rate is the lowest application rate at which water
is applied to plants. It can be estimated as:
r p
a
S S
q
K
p
n
N
100
EU
I (8.18)
Where,
I
n
= net application rate, mm/h
K = conversion constant, 1.0 for metric units
Criteria for Emitter Selection
The emitters are the most important and sensitive component of trickle irrigation.
Selection of the emitters mainly depends on the percentage of wetted area, reliability of the
emitter against clogging and malfunctioning, type of soil to be wetted, emitter discharge and
quality of irrigation water. Ideal emitters should be durable, cheap and give reliable
performance over wider range of pressure variation.
126
Emitter Discharge: Discharge or flow from an emitter or a sprayer can be expressed by an
empirical equation of the form:
Q = K
d
H
x
(8.19)
Where,
q = emitter flow rate, L/h (gph)
K
d
= discharge coefficient
H = working pressure head at the emitter, m (ft)
x = emitter discharge exponent
The value of x characterizes discharge versus pressure relationship of an emitter. Higher the
value of x, more discharge will be affected by variations in pressure and vice versa (Fig. 8.6).
Figure 8.6 Discharge variation due to pressure changes for emitters having different
discharge exponents (Source: Keller and Bliesner, 1990, page 430)
The exponent of long-path emitters is between 0.7 to 0.8 while tortuous path emitters have
values between 0.5-0.7. The orifice and nozzle type emitters and sprayers have the exponent
of 0.5. For fully compensating, x = 0.0 (Keller and Bliesner, 1990, p. 429).
The exponent x can be calculated knowing discharge at two different operating pressures as:
) / log(
) / (q log
x
2 1
2 1
H H
q
(8.20)
Pressure head variation, %
Rated discharge
30
20
10
0
-10
-20
D
i
s
c
h
a
r
g
e

v
a
r
i
a
t
i
o
n
,

%
-30 -20 -10 0 10 20
30
x = 1.0
x = 0.75
x = 0.5
x = 0.4
x = 0.0
Laminar, x = 1.0
Long path, x = 0.75
Orifice, x = 0.5
Vortex, x = 0.4
Fully compensating, x = 0.0
127
Where q
l
and q
2
are measured emitter discharges at pressure head H
1
and H
2
respectively. K
d
can be calculated knowing the discharge exponent and using equation.
Sensitivity to Clogging: This is one of the major factors to be considered while selecting
emitters. Clogging mostly depends on size of the emitter openings and velocity of water
through the emitter. For low discharge, the size of the emitter should be between 0.25 to 2.5
mm but small diameter emitters are more sensitive to clogging and vice versa. Therefore to
avoid clogging, general recommendation is to remove all particles larger than one-tenth of the
diameter of emitter. Relatively larger opening and self-flushing emitters are less subjective to
clogging.
Coefficient of Variation: Discharge of emitters varies due to non-ideal operating and
manufacturing factors. The coefficient of variation is determined by measuring flow rate of
several identical emitters and it is computed as:
2 / 1
2 / 1
3 2 1
v
) 1 (
) .......
C

n q
nq q q q q
a
a n
(8.21)
Where
C
v
= manufacturer's coefficient of variation
q
l
, q
2
,q
3
..q
n
discharge of individual emitters, l/h
q
a
= average emitter discharge (q
l
+ q
2
+q
3
+ q
n
)/n, L/h, and
n = number of emitters tested.
The coefficient of variation for good point-source emitters should be less than 0.05. The
performance of the system would be unacceptable if C
v
is higher than 0.15. For line-source
system, the above limits are doubled.
Temperature: Discharge of an emitter may be sensitive to water temperature due to
dependence of viscosity on temperature. As water flows through the pipes, its temperature
increases toward the ends of the laterals. Increase in discharge due to rise in temperature
partially offsets decrease in discharge due to pressure drop in long laterals. For different
emitter devices the change in discharge is about 1 percent for each 2 to 4C. For more details,
the reader is referred to (Keller and Bliesner, 1990, page 495).
Evaluation and Design Considerations
It is not possible to present here detailed design of trickle irrigation. For this the
reader is referred to an excellent and relatively latest text book by Keller and Bliesner (1990).
128
Evaluation: Evaluation of trickle irrigation system includes measuring different parameters
including discharge of emitters along selected laterals, emission uniformity, pressure
variation and efficiency factors. Discharge of emitters can be measured by collection, water
in a measuring flask or graduated cylinder. Pressure can be easily measured by pressure
gauge. Recommended locations for measurements are at the inlet end, at one-third, at two-
third and near the far end of laterals and manifold. Detailed form given by Merrian and Keller
(1978) may be helpful in data collection.
Example 8.2: Data was collected on a lateral of a trickle irrigation system at different
pressures as shown below.
Calculate and compare different factors of the system. Emitters were orifice type. Assume
discharge correction factor (DCF) is 1.0 for simplicity (see Merriam and Keller, 1978, P. 137-
139).
Average volume collected per minute at
Pressure
(m)
Inlet end 1/3 down 2/3 down Far end
5 (3.52) 48.0 2.88 51.5 3.09 53.0 3.18 57.5 3.45
15 (10.56) 121 5.26 131 7.86 119 7.14 116 6.96
25 (17.60) 146 8.76 145 8.70 147 8.82 150 9.0
Sample calculations for 5 psi low one quarter = 2.88 lph
Average discharge (2.88 + 3.09 + 3.18 + 3.45) = 3.15 lph
Emission Uniformity = Minimum rate of discharge per plant
Average rate of discharge per plant
= 2.88 x 100 = 91.4%
3.15
PELQ = 0.9 EU = 0.9 x 91.4 = 82.3%
Exponents, x
1
= log (q
1
/q
2
)
log (H
1
/H
2
)
= log (2.88/5.26)
log (3.52/10.56)
= 0.55
129
By rearranging the above equation
K
d
= q/H
*
or
K
d1
= 2.88 = 1.44
3.52
0.55
System PELQ
Efficiency Reduction Factor, ERF = 1/DCF = 1
System PELQ = ERF x Test PELQ
= 1 x 82.3
= 82.3%
Control of Clogging
Clogging of emitters is very difficult to control. It is difficult to detect in the field and
expensive to clean or replace the clogged emitters. Clogging may be caused by suspended
particles of sand, silt, clay and debris. Small particles deposit in low-velocity areas in the
different parts of the trickle system. With passage of time these particles gather mass due to
coagulation and occur may often clog the system. Clogging if irrigation water contains high
amount of soluble salts such as calcium, magnesium, carbonates, bicarbonates and sulphides.
Sometime growing of algae and bacteria may also clog the system.
Thus appropriate measures are essential to prevent clogging. Suspended particles can
be removed by using settling ponds and filters. Periodic flushing of pipes, emission devices
and filters is also helpful to check clogging. Chemical treatments by algaecides and bacteria-
cides are used to control biologically caused clogging.
Filtration
Removal of suspended particle is essential for satisfactory performance of trickle
system. The suspended material that can plug the emitters may be inorganic or organic. The
organic material includes algae, bacteria, larvae, fish, snails, and different parts of plants. The
major inorganic solids are sand, silt and clay particles. Settling basins, sand and screen filters,
cartridge and centrifugal separators are commonly used to remove suspended material.
Settling Basins: Settling basins, ponds or reservoirs can remove large volume of suspended
particles. Longer detention time is needed to remove smaller particles and vice versa.
Removal of clay particles is accelerated by using flocculating agent e.g. alum or poly-
electrolytes. The settling basins may need to be cleaned several times in a year depending
upon sediments in irrigation water. In addition to this, algae growth and wind blown
130
contaminants can be severe problems in settling basins. Due to these setbacks, setting basins
should be used only in extreme conditions.
Sand Filters: Sand or sand media filters are commonly used to remove heavy load of organic
and inorganic materials. These filters consist of cylindrical tanks having layers of sand and
fine gravel. Performance of these filters depends on water quality, flow rate, allowable
pressure drop, and sizes of sand used. The maximum recommended pressure drop across a
sand filter is about 70 Kpa (10 psi). Back flushing should be frequent enough to avoid
excessive pressure drop. Cleaning of the filters may be either manually or automatically.
Automatic systems should have at least two tanks. Out of this one is operating while the other
is being back-flushed as shown in Fig. 8.7.
(a) (b)
Figure 8.7 A sand media filter: a) Filtering process, b) Back process
(Source: James, 1988, p.290)
Centrifugal Separators: Centrifugal filters are also called sand or cyclonic separators. They
use centrifugal force to remove and eject high density particles from water. However,
centrifugal separators are not effective to remove organic contaminants. These filters should
be placed upstream of sand, screen and cartridge filters that remove finer particles and
organic materials.
Screen Filters: Long cylindrical filters made of stainless steel or nylon are the most common
type of filters used with trickle systems. Like sand filters, the screen filters should be
frequently washed for proper functioning of the filters.
Cartridge Filters: These filters can remove organic materials and extremely fine particles
that cannot be removed by other filters. Cartridge may be disposable or washable that is
usually made of nylon, cotton and fibreglass.
131
REFERENCES
Allen, R.G. 2004. Manual of CATCH3D Sprinkler Overlapping Program. Utah State
University, USA.
Callies, R.E. 1978. Corner pivot, an effective corner watering system. ASAE Paper No. 78-
2006.
Christiansen, J.E. 1942. Irrigation by sprinkling. University of California, Berkely, Bull. 670.
Ghinassi G. 2008. Manual for performance evaluation of sprinkler and drip irrigation
systems. International Commission on Irrigation and Drainage (ICID Publication No. 94),
New Delhi, India
Heermann, D.F. and P.R. Hein. 1968. Performance characteristics of self propelled center
pivot sprinkler irrigation system. Trans. ASAE 11: (1) 11-15.
James, Larry, G. 1988. Principles of farm irrigation system design. John Wiley & Sons, New
York, U.S.A.
Keller Jack and Bliesner, Ron D. 1990. Sprinkle and trickle irrigation. Van Nostrand
Reinhold, New York, U.S.A.
Korven, H.C. and W.E. Randall. 1975. Irrigation on the practices. Agriculture Canada,
Publication 1488, Ottawa, Ontario, Canada.
Merkley, G. P. and Allen, R. G. 2007. Sprinkler and trickle irrigation, Lecture Notes.
Biological and Engineering Department, Utah State University, USA.
Merriam, J.L, and J. Keller. 1978. Farm irrigation system evaluation, 3
rd
ed., Logan, Utah:
Agricultural and Irrigation Engineering Department, Utah State University, USA.
Phocaides, A. 2000. Technical handbook on pressurized irrigation techniques. Food and
Agriculture Organization of the United Nations.
Ring, L., and D.F. Heermann. 1978. Determining center pivot sprinkler uniformities. ASAE
Paper No. 78-2001.
Sne, Moshe. 2006. Guidelines for planning and design of micro irrigation in arid and semi-
arid regions. International Commission on Irrigation and Drainage, India.
Splinter, W.E. 1976. Center pivot irrigation. Scientific American. Vol. 234, No. 6, pp.90-99.
Watters, G.Z., and Keller, J. 1978. Trickle irrigation tubing hydraulics. ASCE Tech. Paper
No. 78-2015.
United States Department of Agriculture. 1960. SCS National Engineering handbook, Section
15 Irrigation, Chapter 11 Sprinkler Irrigation, USDA, SCS, Washington, D.C. 83 p.

You might also like