Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 118

CHAPTER 1

INTRODUCTION

TERMS:

Irrigation – is the application of water to the soil to supplement low rainfall to provide
moisture timely and appropriate in quantity and distribution for plant growth.
Irrigation is the controlled application of water to croplands. Its primary objective is to create
an optimal soil moisture regime for maximizing crop production and quality while at the same
time minimizing the environmental degradation inherent in irrigation of agricultural lands.

-
- Irrigation is the application of water to the soil to supplement natural precipitation and
provide an environment that is optimum for crop production. Well Irrigated crops
produce more food.

Engineering - is the science, skill, and profession of acquiring and applying scientific,
economic, social, and practical knowledge, in order to design and also build
structures, machines, devices, systems, materials and processes.

Irrigation Engineering - Irrigation is the application of water to the soil to supplement natural
precipitation and provide an environment that is optimum for crop production through
the application of engineering principles and processes or systems.

Irrigation Water – is the portion of the total water required by a crop to produce an optimum
yield that is not supplied by rainfall.

Irrigation Water Management – integration of storage, diversion, conveyance, regulation,


measurement, distribution and application of the right amount of water at the proper
time and removal of excess water from farms to promote increased production in
conjunction with improved cultural practices.

Evapotranspiration – represent the combination of water evaporated from the plant and the
soil surface plus the amount of water which passes through the soil into the roots,
through the stem of the plant, and to the leaves where it passes thorugh small pores
called stomates.

Transpiration – the translocation of moisture from the root zone through the plants.

Basal Evapotranspiration – ETo from available soil moisture at time of planting plus
precipitation.

Fundamental Principles
- Rapid population growth – recognition of the urgent importance of irrigation as the
basis of the agricultural economy.
- Irrigation is basically an attempt by man to locally alter the hydrologic cycle and to
promote agricultural productivity.
- Irrigation engineering integrates numerous disciplines and combines elements of
soil science, hydraulics, hydrology and economic analysis.

Purposes of Irrigation
 Great impact on the development modern civilization. Decreased risk of food
instability brought about by irrigation. Stability of food resources allowed progress
(from nomadic, food gathering societies to groups with permanent or semi-permanent
dwellings. Countries of large proportion of irrigated land tend to have increased
stability in food production.
1
 Increased productivity in agriculture through irrigation also leads to increased
opportunities in businesses which supply the agricultural sector (sales in irrigation
equipment and indirectly in sales of seed, fertilizer, pesticides, herbicides, and
agricultural machinery.

Expected Benefits of Irrigation in Terms of Crop Yield

- To quantify the economic benefit of irrigation, it is necessary to quantify the


expected increase in yield as a function of increasing amount of water delivered.

- Crop water production function (Straightforward method)

A. Applied Water VS Yield

Max yield
Decreased
aeration and
Y2 impedance of
necessary gas
transfer causes
the decrease in
Y1 crop yield

AW1 AW2

Figure 1.1. Yield as a function of applied water for SJ2 variety cotton

Assumption: Nutrient levels will be assumed adequate to produce maximum growth


at any level of irrigation and the crop will be assumed free of diseases, weeds, and pests. The
crop response to irrigation will therefore represent a maximum at any given water level.

2. Consumptive Use or ETo VS Yield

2
Figure 1.2. Yield as a function of ETo for SJ2 cotton.

Figure 1.3. Idealized water production function for yield versus applied water

- Lowest level of ETo is assumed due to available soil moisture at time of planting
plus precipitation throughout the growing season.

- Applied water includes initial soil moisture plus growing season precipitation.

- As the level of applied water increases, less applied water is converted to ETo.

Example 1
Assume that the applied water for Figure 1.3 includes the ASM at time of planting
plus precipitation and irrigation. The yield versus applied water relationship may be described
by the following quadratic equation.

Y = -4941 + 35.83(AW) – 0.0195(AW)2


Where:
Y = yield in kg/ha
AW = applied water, mm

and the yield VS ETo:

Y = -3783 + 25.65(ET)

a) Compute the ratio of the change in yield to the change in applied water when
applied water is increased from 400 to 500 mm and from 700 to 800mm.
b) Define the efficiency as the amount of evapotranspiration produced for a given
level of applied water expressed as a percent. Compute the efficiency at an
applied water level of 450 mm and at a level of 750 mm.

Issues and Problems in Irrigation Development

1. Poor performance of existing irrigation systems.


2. Neglect of Irrigation R and D and Inadequate Technological Base for Planning and
Design of Irrigation Systems.
3. High cost of irrigation development and operation and maintenance of Irrigation
systems.
4. Sustainability
CHAPTER 2
BASIC SOIL-WATER-PLANT RELATIONSHIPS

IMPORTANT PROPERTIES OF SOIL WATER SYSTEMS

- Soil consists of solid materials, moisture, and gases in various proportions.

3
- Soil solids composed mainly of inorganic or mineral materials and relatively small
proportions of organic matter.

Soil Texture – defined on the basis of the size distribution of its particles.
- The texture of the soil gives qualitative estimates of the ease with which water can
move through it.

Light-textured soils – has more large diameter particles (i.e. sand). Offers less resistance to
the movement of water, hence, have good internal drainage. Such soils are, however,
associated with high seepage and percolation loss rates. It is also an index of water retaining
capabilities, porosities, and erosivities of soils.

Heavy-textured soils – has more small diameter particles (i.e. clay)

Pore Spaces – air and water in soil water systems occupy the pore spaces or voids in
between soil particles. Pore spaces vary in size and shape. The smaller pores are usually
filled with water except when the soil is extremely dry. Pore spaces play key roles in water
management. Among other things, they influence the movement of air, water, nutrients,
pesticides and microorganisms in the soil.

Va

Vw

Vs

Figure 2-3. A hypothetical representation of the proportions by volume of air, water,


and solids in soils.

Where:
Va = Volume of air
Vw = Volume of water
Vs = Volume of solids

Vv = Volume of voids = Va + Vw
VT = Total volume = Va + Vw + Vs

Porosity
Porosity (p) of the soil is defined as the ratio of the volume occupied by voids to the
total volume of the soil system.

Vv V a +V w
p= =
VT VT (2-1)

Moisture Content
The moisture content of a soil may be expressed on a volume or dry weight basis. On
a volume basis, the percent moisture content (Pv) may be expressed as:
4
Vw
Pv = x 100
VT (2-2)

On a dry weight basis, the moisture content of the soil (Pw) is defined as the ratio of the
weight of its water to the weight of its solids. That is:

Ww ρV w
Pw = x 100= x 100
Ws ρsV s (2-3)

Where:
Ww = Weight of water in the soil
Ws = Weight of solids or oven-dried weight soil
 = Density of water
s = Density of soil solids

When expressed on a volume basis, the moisture content of a soil allows for the
direct calculation of the depth of water (dw) on a given depth of soil (D). In equation form:
Pv
d w= x D
100 (2-4)

dw = Pw. D. As

Qt = Ad

Equation 2-4 is quite useful. During irrigation for example, the irrigation water

requirement is calculated in terms of the depth of water to be applied.

Specific Gravity

The specific gravity of a substance is defined as the ratio of the unit weight of that
substance to that of water. In soils, the specific gravity may be expressed in terms of either
the real (true) or apparent specific gravity.
Real specific gravity (Rs) is the ratio of the dry weight of the soil particles to the weight
of water having the same volume as these particles. In equation form:
Ws
Rs =
ρVs (2-5)

It should be noted that in determining the weight of an equivalent volume of water, the volume
of voids is excluded and only the volume of solids (Vs) was used. The real specific gravity of
most soils ranges from 2.4 to 2.8.

5
Ws
As=
ρV T
Pv
As=
Pw
A s =Rs (1−P)
1-P = As/Rs P = 1-(As/Rs)

Problems:

1. A soil sample weighing 700 grams when brought from the field in a cylinder of 50
grams, weighed 595 grams when dried. The length of the column of soil is 12 cm and
the diameter of the cylinder is 6 cm. Compute a) Pw, b) Pv and c) As.

Wt = 700g
Ws = 595-50g = 545g
Vt = (3.1416/4) 6^2 (12) = 339.29 cm3

Solution:
a) Pw = (Ww/Ws) Ww =( Wt – 50) – 545 = 650-545 = 105g
= (105/545)*100 = 19.27%

b) Pv = (Vw/Vt)*100 = ((Ww/w)/339.29)*100 = ((105/1g/cc)/333.29)*100 = 30.95%

c) As = (Pv/Pw) = (30.95/19.27) = 1.61

Or As = Ws/(w*Vt) = 545/(1*339.29) = 1.61

2. If the total volume of soil is 250 cubic centimeter, what is the diameter of the sampler if
the height of soil column secured is 10 cm.

Vt = A*h A = /4 d^2


Vt = /4 d^2 * h
d^2 = Vt*4/(*h) = 250*4/(*10)
d = 5.64cm

Problems:

1. Given a saturated soil sample with mass equal to 283 g and an oven dry mass of 202
g, find the water content on a mass and volume basis, the soil bulk density, and the
porosity. Assume density of water equals 1.00g/cm 3 and density of soil particles
equals 2.65 g/cm3.

Given:

6
Wt = 283g
Ws = 202 g
s = 2.65 g/cc
a) Pw = (Ww/Ws)* 100 = ((283-202)/(202))*100 = 40.1% @ Ww = 81g
b) Pv = (Vw/Vt)*100
Vt = Vw + Vs
Vt = (81/1) + (202/2.65) = 157.23 cc

Pv = (81/157.23)*100 = 51.52%

b = Ws/Vt = (202/157.23) = 1.28g/cc

P = (Vv/Vt)*100 = (81/157.23)*100 = 51.52%

As =( Ws/wVt) = 202/(1*157.23) = 1.28

Rs = (Ws/bVs) = 202/(2.65*202) = 2.65

P = 1 – (As/Rs)*100 = (1 – (1.28/2.65))*100 = 51.52%

2. How much water (m3) must be added to a field of area 3 ha to increase the volumetric
water content of the top 40 cm from 16 to 28 percent? Assume all water added to the
field stays in the top 40 cm.

3. A soil sample with a wet weight of 300 g has a 28.0% water content on a mass basis.
Its saturated water content is 36.1% on a mass basis. Assume density of water
equals 1.00g/cm3 and density of soil particles equals 2.65 g/cm 3.
a. Find the mass of water and soil porosity of the sample at 28.0% water
content.
b. If the sample at 28.0% water content is representative of the top 45 cm of a
0.3 ha plot, how much water, in m3 and as an average areal depth in cm,
must drain from the top 45 cm to increase air filled porosity to 25.0%

4. A soil sample taken in the field has a wet weight of 373 g. The sample is dried at 105-
110oC to a constant weight of 295 g. The apparent specific of the soil is 1.4 and the
density of the soil particles is 2.58 g/cm3. Find the water content on a mass and
volume basis and soil porosity.

FORCES ACTING ON SOIL WATER


Water molecules are dipole in nature.
Electrical Charges –
Cohesive Forces – due to hydrogen bonding (water to water)
Osmotic Forces – caused by salt or ion concentration differences or gradients.
Imbibitional or Adhesive Forces – strong attraction of liquid and solid
(Bunderwell’s).

7
Soil Moisture Tension
Soil Moisture Tension

Field Capacity
(0.33 atm)

Saturation Point

Hydroscopic Water Permanent Wilting


(10,000 atm) Point (15 atm)
Water Film
Thickness
Clay Clay
Form of Soil
UW Available Water Excess Water
Water

Predominant
Adhesive Cohesive Gravitational Forces

The clay to water forces diminish rapidly with the distance from the clay particle. As the water
film increases in thickness, the weight of water increases and the gravitational forces
predominate.

Terms:

Field Capacity (FC) – The point where the gravitational forces equal to cohesive forces. It is
a measure of the water holding capacity of the soil.
- the moisture content of the soil when free drainage have ceased.

Saturation Point (SP) – The moisture level at which all voids are filled with water.

Permanent Wilting Point (PWP) – It represent the point of moisture level where a plant
permanently wilts.

Forms of Soil Water

Gravitational or excess water – The soil moisture in between the field capacity and the
saturation point which is subjected to free drainage.

Available Water – The soil moisture formed between the field capacity and the permanent
wilting point.

Readily Available Moisture – 75% of the available water.

Unavailable Water – The soil moisture storage below the permanent wilting point.

Note:
The saturation point and the field capacity are soil physical constants and as such
vary only with soil physical properties such as soil structure, texture and stratification.

8
The permanent wilting point, on the other hand, is arbitrarily defined as it is extremely
variable. Among other things, it varies with the plant species, plant variety, soil physical
properties affecting soil moisture movement, the stage of growth of the plant and the plant’s
aerial environment.

Saturation Point

FC

Drainage
Soil
Bucket Available Moisture (AM)

PWP

Hydroscopic

When to Irrigate
If FC – PWP = AW (2.0)
at ½ AW --- conservative
at ¼ AW --- sensitive

note: Maximum Allowable Depletion (MAD) = ¾ of AW

Depth of Irrigation
dw = Pv As Drz (2.1)

Where:
dw = Depth of irrigation
Pv = Moisture content volume basis
As = Apparent specific gravity
Dr = Depth of root zone
z

Irrigation Interval (Ii)


Ii= (dw/c.u.) (2.2)
@ c.u. = ETo x Kc

The Concept of Surface Tension


Water molecules on the water surface VS water molecules below the water surface.
- water surfaces is considered under tension
- a greater force is needed to pull a molecule from the surface than to move a
molecule within a liquid.

9
Surface tension or surface tension coefficient (T) – is expressed in force per unit
length (in contrast with tensile stress which is force per unit area). It may be viewed as energy
stored in a film of water for a unit increase in the films’ surface area.

Illustration: Wired being pulled being from water (Kirkham and Powers, 1972)

F = 2 TL (2.3)

Where:
F = upward force exerted to balance the surface tension forces
T = surface tension coefficient
L = length of the wire
The multiplier 2 is used since surface tension forces act on both sides of the
wire.

In terms of Work (F x d):

W = F x d = 2TL d

W
T=
or 2Ld (2.4)

Surface tension is very important property of water. It helps explain some energy
concepts in soil moisture. Curvatures or increases in surface areas in water pores may be
viewed as stored energy. A sphere of water dropping a certain distance from a dining table to
the floor will splatter into smaller droplets having more surface areas or stored energy to
compensate for the decrease in its potential energy. A beach stroller usually notices that his

10
feet sink deeper when the sand is saturated or extremely dry (as the particles easily rill on top
of one another) than when the sand is not too wet nor dry (more water film surface areas).

Surface tension also helps explain the capillary rise of water. If a tube of capillary
dimension is inserted on a free water surface, water will rise on the tube in response to the
water-to-solid (adhesive) forces (AF).

Terms:

Matric Potential – is that portion of the water potential than can be attributed to the
attraction of the soil matrix to water, and includes the capillary pressure. The matric
potential is more negative as the water content decreases.

Solute Potential – is that portion of the water potential that can be attributed to the
presence of solutes in the water; it is also called osmotic potential. It is zero for pure
water.

Gravitational Potential – is the energy associated with the vertical location of the water
or its elevation with respect to an arbitrary reference level.

Pressure Potential – is that portion of the water potential that results from an over-all
pressure that is different from the reference pressure (usually atmospheric pressure).

Derivation of the Height with Which Water Will Rise in a Capillary Tube:

AF

11
w
T G

At point A, F = AF = (2r) T

Resolving into resultant vertical forces:

(2r) T cos w = weight of water

(2r) T cos w = (r2h)  g

Where:
g = gravitational constant
r = radius of the capillary tube
h = height of the water column
or
2Tcos w
h=
ρrg (2.5)

In soils underlain with shallow water table the capillary rise of water can sometimes
be sufficient to support a stand of field crops (i.e. central Luzon, Israel).

Sample Problem

The field capacity and permanent wilting point of the soil has been determined to be 28% and
13%, respectively. It is desired that a corn field area of 5000 m 2 be irrigated up to field
capacity when 75% of the available moisture has been depleted. What is the moisture content
of the soil when irrigation should start, and the depth of water to be applied (D = 0.90m). If an
irrigation pump is to deliver a discharge of 10 li/s, how many hours are required to irrigate the
whole area? (use the formula Qt = Ad; where Q = pump discharge, t = time of irrigation, A =
area to be irrigated, and d = depth of irrigation).

1. For MC when irrigation should start:


a) 16.75% b) 11.25% c) 15.00% d) 12.00%

2. For depth of water to be applied:


a) 0.1012 m b) 1.012 m c) 0.21 m d) 0.141 m

12
3. For irrigation time:
a) 0.586 day b) 1 day c) 0.25 day d) 2 days

PRACTICE PROBLEMS: SOIL MOISTURE CONTENT/SOIL-WATER-PLANT


RELATIONSHIPS

1. Determine the soil water content n percent by weight for a 120g sample in a moist
condition which when oven dried, weighed 90g.

2. The combined weight of a sample and a tin is 50.0 grams. The tin weighs 20.0 grams.
After oven drying, the soil sample and tin have a weight of 45 grams. Determine the
soil water content (by weight) of the soil.

3. Determine the porosity of a soil profile if it has a bulk density of 1.8g/cc and the
density of the soil particles is 2.7g/cc.

4. Based on the field test, the bulk density of a soil profile was found to be 99.8 lb/ft 3 and
the specific gravity of the soil particles is 2.8. Determine the porosity and soil water
content (by weight) of the profile.

5. The porosity of the soil is 0.4 (cc/cc). Determine the volumetric soil water content as a
fraction of the degree of saturation (the percentage of the void volume by moisture) is
60%. Also, determine the depth of water contained in the top 50 cm of a soil with a
degree of saturation of 60% and porosity of 0.4.

6. Determine the soil water content at field capacity and plant available soil water
content in inches for a 30-inch soil column. The field capacity by volume is 35% and
the wilting point is 10%.

7. The field capacity of a soil is 35% and the wilting point is 15% by weight. The bulk
density of the soil is 78 lb/ft 3. Determine the available water content in percent by
weight and by volume. Also determine the plant available soil water content in the top
2ft of soil. Express the answer as depth of water in inches.

8. A moist soil sample (diameter 1.2 inches, length 2.0 inches) weighs 65.42 grams.
After drying in an oven the sample weighs 56.14 grams. The soil particles have
specific gravity of 2.65. Determine the soil moisture of the sample in (a) grams per
gram of soil; (b) grams per gram of moist soil; and (c) cubic centimeters per cubic
centimeters of total soil volume. Find the (d) porosity of the soil sample and the (e)
percentage of saturation.

9. A sharp-edged cylinder 7.5 cm in diameter is carefully driven in to the soil so


negligible compaction occurs. A 12-cm column of soil is obtained. The fresh and dry
masses of the soil sample are 910.6 and 736.3 grams, respectively. Assume the
particle density of the soil is 2.6 grams per cubic centimeters. Determine the
following:

a. apparent specific gravity


b. percent porosity
c. percent moisture content on a dry weight basis
d. percent moisture content on a volume basis
e. equivalent depth of water present within the soil column

10. Three soil samples were taken from the three points within a hectare farm were
placed in an oven for moisture content determination. The results show that the

13
average soil moisture content on dry weight basis is 18.5%. The bulk density of the
soil is found to be 1.35 grams/cm3. Determine:

a. equivalent depth of water present in the farm within a depth of 50 cm.


b. volume of water needed to bring the present moisture content of 18.5% to
30.5% by dry weight for every meter depth of soil.

11. (BP ’82) A farmer collected a soil sample two days after irrigation. The cylindrical
sample has the following dimensions: diameter 12.7 cm, height 20.32 cm. The
sample weighed 3000 grams before drying and 2340 grams after drying. Estimate the
field capacity in cm per meter depth of soil.

12. A barrel of soil was collected in the field and determined to have a wet mass of 230
kilograms. If the dry moisture content of the soil was 0.18, find the (a) weight of the
water and (b) dry weight of solids in the volume.

13. Calculate the bulk volume of 45.5 kg wet soil containing of 8.2 kg of associated water
if its bulk density is 1.25 g/cm3.

14. Determine the moisture content of the clay loam if the amount of water used by the
plant for its growth is 45%. Clay loam soil had field capacity and permanent wilting
point of 32% and 28% dry weight, respectively.

15. The root zone of a silt loam is at its permanent wilting point at 13% dry weight. What
depth of water must be applied in order to bring its water content up to its field
capacity of 22% dry weight? Assume that the root zone depth is 0.16m and soil bulk
density is 1.280 kg/cm3.

16. A soil has initial volumetric moisture content of 12%. What volume of water will be
needed to wet a hectare of this soil to its field capacity of 19% dry weight? Bulk
density is 1.320 kg/m3 and the depth of the root zone is 0.5m.

14
CHAPTER 3
MOISTURE MEASUREMENT IN IRRIGATED SOILS

Measurement of Soil Moisture


The moisture content of the soil may be determined using any of the several known
methods ranging from direct weighing through the more indirect ones such as neutron
scattering, tensiometry and electrical resistance blocks. The choice of any method to use will
depend, among other things, on the desired accuracy, aerial coverage, and the expected
range of variations in moisture content.

The Feel and Appearance Method


The feel and appearance method is employed in the field where no
instruments are available. Samples of soil are obtained at the root zone level of the
crop; and by looking and looking and feeling between the fingers, the moisture
content may be determined approximately (Table 1). The soil plasticity indices may
be used as a guide to determine the approximate soil moisture content.

Direct Weighing or Gravimetric Method


This method involves the weighing of moist soil sample (50 to 100 grams),
placing it in an oven at 105-110oC until all moisture is driven off as indicated by no
further loss in sample weight with time, and then weighing the oven-dry sample. The
difference in weight before and after oven-drying represents the soil moisture and is
expressed in terms of dry weight of the soil multiplied by 100 to give the percent
moisture content.

Advantages:
1. It is simple, direct and as such does not require expensive or sophisticated
gadgetry;
2. It can be very accurate as evidenced by the fact that all other methods of
measuring soil moisture are calibrated against it.

Disadvantages:
1. It requires considerable physical efforts;
2. It is time consuming (usually 2-3 days);

15
3. It is destructive form of sampling whereby it is not possible to make repeated
sampling at the same point.

Electric Resistance Block (Bouyoucous Method)


Since natural soil water is a very good conductor of electricity the electrical
resistance or conductance of the soil may be used as an indirect measure of soil
moisture content. In this method, two electrodes are embedded in a small porous
block made of gypsum (plaster of Paris), nylon or fiberglass. The unit is then set at
the desired soil depth with the leads from the electrodes jotting out from the soil
surface. Since the block is porous, it becomes and integral part of the soil and
maintains a soil moisture content in equilibrium with the soil. The fluctuation of
moisture in the block affects the electrical resistance measured by a portable
Wheatstone bridge. It reflects the moisture content of the soil.

A resistance block must be first calibrated before actual soil moisture


measurements are made. Its readings are usually calibrated against soil moisture
content or tension. Calibration for moisture content is done using the gravimetric
method while for the soil moisture tension is done with the help of a pressure
chamber-pressure plate apparatus.

Advantages:
1. Relatively cheaper and easier to fabricate;
2. They operate satisfactorily in the available soil moisture range;
3. It is possible to make repeated observations at any given point;
4. Rapid observations are possible, with one observation taking no longer than a
minute.

Disadvantages:
1. Can be influenced by salt concentrations in the soil;
2. Gypsum tend to dissolve in soil water and may not be suited for long term use;
3. With time, resistance blocks exhibit a high degree of hysteresis, that is, there is a
remarkable drift in their calibration curves;
4. It is extremely difficult to bring them to their original saturated condition readings
even if the soil is flooded for a considerable length of time.
5. They exhibit a very long time lag in response to changes in soil moisture content.

Table 1. General relationship between soil moisture and the feel and appearance of the
soil.
         
FEEL OR APPEARANCE
Moisture
Between
Wilting point Heavy and Very
Coarse Soil Light Soil Medium Soil
and Field Heavy Soils
Capacity

0 (Wilting point) Dry, loose, single Dry, loose, flows Dry, sometimes Hard, baked, cracked,
grained, flows through fingers crusted but sometimes has loose
through fingers breaks down crumbs on surface
easily to powder
condition

50% or less Appears dry, will Appears dry, will Somewhat Somewhat pliable
not form a ball not form a ball crumbly, holds balls under pressure
16
together under
pressure

50% - 75% Same as above Tends to ball Forms ball, Forms ball, ribbons
under pressure, somewhat plastic, out between thumb
but seldom holds will sometimes and forefinger
stick with
pressure

75% to FC Tends to stick Forms weak ball, Forms ball, very Easily ribbons out
together, breaks easily, will pliable, sticks between fingers, has
sometimes forms not stick readily if high in slick feeling
very weak ball clay
under pressure

FC Wet outline of ball Same as coarse Same as coarse Same as coarse soil
soil soil

Above FC Appearance of free Free water Can squeeze out Puddles and free
water from soil is released with water water forms on
balled in hand kneading surface

Tensiometer Method
A tensiometer is a simple device for measuring soil moisture tension in the
range of 0 to 700 millibars. With the use of a soil moisture versus soil tension
calibration curve, it is a good instrument for soil moisture determination.

A tensiometer consists of a porous cup (filled with water) attached to a


mercury manometer. Water is free to move in and out of the porous cup. If the soil
moisture is not saturated, water moves from the cup into the soil causing mercury to
rise in the manometer tube. The height of mercury represents the soil moisture
tension.

A Tensiometer:

17
Neutron Dispersion Method
Soil moisture content can be determined by the neutron scattering technique. When
high energy neutrons are emitted from a source of radioactive material, in the soil they
collide with atomic nuclei. When the particle with which the neutron collides is
approximately the same size and mass as the neutron, approximately half of the original
energy of the neutron will be transmitted to the particle, the neutron will have only half of
its initial velocity. Such slow neutrons are caught by a detector and counted by a counting
unit. The atomic nucleus most effective in slowing neutrons is that of hydrogen. In the soil
hydrogen occurs in large amounts as a component of water. As the water content
increases more neutrons are slowed near the source, and their count is higher. The
volumetric moisture, w is:

1 NC
θw =
F No
where F is a parameter which is usually experimentally determined, N o is the count in a
standard medium, and Nc is the count in the soil. This method is most accurate to study
moisture changes in the same field.
(The neutron dispersion method was developed with the introduction of radioactive
materials. Hydrogen nuclei reduce the energy and high velocity of fast neutrons being
emitted by the radioactive substances. Through collision and scattering, some neutrons
traveling away from the radioactive source are returned as slowed neutrons. Therefore,
the more hydrogen atoms there are in the soil, the more slowed neutrons will occur.
Since hydrogen nuclei occur in water, the number of slowed neutrons is measured and
correlated with the amount of water present in the soil).

18
Figure. Neutron scattering assembly for soil moisture measurement.

The Psychrometer Method


The principle in the use of psychrometer is the measurement of vapor
pressure. The degree of saturation of air by water vapor in equilibrium is related to
the water potential or the free energy of the water. It has been shown that the
difference in water potential is equal to the natural logarithm of the relative humidity
multiplied by the absolute temperature and the gas constant. The difference in water
potential results from changes in both solute constant and the affinity of water to the
soil matrix which in effect measures the total water potential. A change in
psychrometer reading may represent a change in either matric potential or solute
potential or both. It is possible that changes in one of these component potential will
be marked by the changes in the opposite direction of the other. The use of properly
calibrated psychrometer can provide accurate results from 0 to well over 40 bar
suctions.

Gamma Ray Attentuation


This technique depends on the fact that  rays lose part of their energy by
absorption upon striking another substance. Gamma rays that are emitted from
radioactive isotopes, e.g. 137cesium, can be focused into narrow beams to study
moisture in layers as thin as 1 cm, in contrast to neutron scattering which gives
moisture in fairly large volumes, such as sphere of about 15 cm. The disadvantage is
that the readings depend on the soil density, and this may change with moisture
content as a result of swelling. The method is very useful in laboratory work,
especially for measuring relative moisture changes.

Reflectometry Method

19
Pressure Plate and Membrane Apparatus

Table 2. Physical properties of selected soils.


       
APPARENT TOTAL
SATURATION
SPECIFIC POROSITY
SOIL TYPE CAPACITIES
GRAVITY (Percent by
(Percent)
(As) Volume)

Sandy loam 1.26 41.50 52.30


1.66 21.90 36.40

Loams 1.16 47.30 54.90


1.44 31.70 45.70

Silt loams 1.05 58.10 61.10


1.46 30.30 44.30

Gravelly silt loams 1.02 59.60 60.90


1.42 32.20 45.80

Clay and clay loams 0.94 68.10 64.10


1.54 26.90 41.40
       

Table 3. Typical moisture values for various soil types.


         
PERCENT DRY WEIGHT OF SOIL
APPARENT
FIELD WILTING AVAILABLE
SOIL TYPE SPECIFIC
CAPACITY POINT WATER
GRAVITY

Sand 5.00 2.00 3.00 1.52

Sandy loam 12.00 5.00 7.00 1.45

Loam 19.00 10.00 9.00 1.36

Silt loam 22.00 13.00 9.00 1.28

Clay loam 24.00 15.00 9.00 1.28

20
Clay 36.00 20.00 16.00 1.20

Peat 140.00 75.00 65.00 0.32


         

Calculation of soil moisture content by mass and volume


A soil sample weighed 230 g in a moisture box. The mass of the moisture box was 78 g. After
drying at 105 degrees C to a constant mass, the soil and box weighed 204 g. The soil sample
filled a 1000 cc container as it was taken from the field. Find the moisture percentage in the
soil by mass and by volume.
1. Water (%) by mass:
Wet mass of soil = (wet mass of soil + box)-(mass of box)

= 230 - 78 = 152 g

Dry mass of soil = (dry mass of soil + box)-(mass of box)

= 204 - 78 = 126 g

Water (%) by mass = (wet mass - dry mass / dry mass) x 100

= ( 26 / 126 ) x 100 = 21 %
2. Water (%) by volume:
= (vol. of water / bulk vol. of soil) x 100

Vol. of water = mass of water / density of water

= 26 g / 1 g per cc = 26 cc
Hence: Water (%) by volume =( 26 cc / 1000 cc) x 100 = 2.6 %

So, the answers are:

1. Water by mass = 21 %

2. Water by volume = 2.6 %

                                                                                                                                            

Calculation of forms of soil water on mass basis


 
The following data represent a soil sample:
Soil mass at field capacity =  85 g Find :
Soil mass at wilting point =  71 g Water (%) at field capacity  
Air-dry mass                      =  64 g Water (%) at wilting point  
Oven-dry mass                  =  58g Available water (%)  

21
                                         85  -  58
Water at field capacity = ----------------- x 100 = 47 %
                                           58
 
                                       71  -  58
Water at wilting point = ----------------- x 100 = 22 %
                                       58
Available water  =  Field capacity  -  Wilting point
=       47  -   22   =   25%

Calculation of forms of soil water on percentage basis

The soil contained the following moisture contents:


Moisture at field capacity =  25 % Find :
Moisture at wilting point =  16 % Available water (%)
Air-dry moisture =  12 % Capillary water (%)
Oven-dry moisture =  11 % Hygroscopic water (%)
Available water  (%) =  Field capacity  -  Wilting point
=       25  -  16   =   9%
Hygroscopic water  (%) =  air dry water   -  oven dry water
=       12  -  11   =   1%
Capillary water  (%) =  Field capacity  -  hygroscopic water
=       25  -  1   =   24%

---------------------------                                                                                                                     
  
Calculation of depth of water, water holding capacity, porosity

A cube of soil measures 10 cm x 10 cm x 10 cm (D = 10 cm, A = 100 sq cm) and has a total


wet mass of 1760 g of which 290 g is water. Assume the density of water is 1 g per cc and

22
particle density is 2.65 g per cc. Find the depth of water, water-holding capacity and aeration
porosity of the soil.

                                water volume


 1. Depth of water  = --------------------
                               surface area             
                                                      
water volume = mass of water  / density of water
=  290  g / 1  g  per  cc = 290 cc
Dw = 290/100 = 2.9 cm
c = solids volume  / surface area
solids volume = solids mass  / particle density
=  (1760  -  290)  g / 2.65  g  per  cc 
= 555 cc
Hence, c= 555 / 100 sq cm = 5.55 cm
Now we will calculate a as shown in the diagram:
a =D - (b + c)
=10 cm - (2.9 cm + 5.55  cm) = 1.55
Therefore:

a in the diagram is thus1.55 cm b in the diagram is thus 2.9 cm c in the diagram is thus 5.55
cm

2. Water-holding capacity (saturation water content)


= mass of water when saturated / mass of dry soil
= mass of water when saturated / mass of dry soil
= 1 g per cc x ( 1.55 + 2.9 ) / 2.65 g per cc x 5.55 = .3
Water-holding capacity (%)= .3 x 100 = 30 %
3. Aeration porosity = air filled pore vol. / bulk vol. of soil = aA / DA = a / D = 1.55 cm / 10 cm
= .15 Aeration porosity = .15 x 100 = 15 %

23
So the answers are: 1. Depth of water = 2.9 cm 2. Water-holding capacity = 30 % 3.
Aeration porosity = 15 %

CHAPTER 4
CROP WATER REQUIREMENTS/CONSUMPTIVE USE OF WATER

EMPIRICAL EQUATIONS USED IN CALCULATING REFERENCE CROP


EVAPOTRANSPIRATION (ETo)

Introduction

Reference crop evapotranspiration (ETo) or sometimes referred to as consumptive


use (CU) takes into account the effect of climate on crop water requirement. It is defined as
the rate of evapotranspiration from an extensive surface of 8 to 15 cm tall, shading the ground
and not short of water. It is affected by climatological elements, irrigation practices, length of
growing season, stage of crop growth and other related factors. ETo is expressed in mm per
day.

Estimation of ETo

Most estimates of ETo are based on empirical equations involving one or more
meteorological parameters. These are usually verified using other techniques including use
of lysimeter and water budget study of field plots. Some of the empirical equations are the
following:
24
1. Blaney-Criddle equation:

This equation uses measured temperature data as well as general levels of humidity,
sunshine duration and wind velocity. The relationship recommended, representing mean
value over a given month is expressed as:

ETo = c[p(0.46T + 8.13)]

where ETo = reference crop ET in mm/day for the month considered


c = adjustment factor which depends on minimum RH, sunshine hrs,
and daytime wind estimates
p = mean daily percentage of total annual daytime hrs for a given month
and latitude (Table 1)
T = mean daily temperature in oC for the month considered

Table 1. Mean daily percentage (p) of annual daytime hours for 15 oN and 20oN latitudes

MONTH LATITUDE MONTH LATITUDE


15oN 20oN 15oN 20oN
Jan .26 .25 Jul .29 .30
Feb .26 .26 Aug .28 .29
Mar .27 .27 Sept .28 .28
Apr .28 .28 Oct .27 .26
May .29 .29 Nov .26 .25
June .29 .30 Dec .25 .25

Example1. Calculate the ETo for October at 15oN using the data below:

Tmax or maximum air temperature = 28oC


Tmin or minimum air temperature = 16oC
c or the adjustment factor = 1.1

The procedure in calculating ETo will be:

a) Determine the mean air temperature or T


T = (Tmax + T min)/2
= (28 oC + 16oC)/2
T = 22oC

b) From Table 1, p on October at 15oN = 0.27

c) Therefore, substituting the vales of c, T, and p in the formula,

ETo = 1.1 [0.27{(0.46)(22)}+ 8.13]


= 1.1 [0.27 (10.12 + 8.13)]
ETo = 5.4 mm/day

The Blaney-Criddle method is suggested for areas where air temperature is the only
climatic data available. Calculation of daily ETo should be made for periods no shorter
than one month. Since for a given location climatic conditions and consequently ETo may
vary greatly from year to year, ETo should preferably be calculated for each calendar

25
month for each year of record rather than by using mean temperatures based on several
years of records.

2. Radiation equation:

This equation is suggested for areas where available climatic data include measured
air temperature and sunshine, cloudiness or radiation, but not measured wind and
humidity. The relationship recommended is expressed as:

ETo = c (WRs)

where:
ETo = reference crop ET in equivalent mm/day for the periods considered
c = adjustment factor which depends on mean humidity and daytime
wind conditions
W = weighting factor which depends on temperature and altitude (Table
2)
Rs = solar radiation in equivalent evaporation in mm/day

Rs = (0.25 + 0.50 n/N)Ra

n = actual measured bright sunshine hours


N = maximum possible sunshine hours (Table 3)
Ra = extra terrestrial radiation expressed in equivalent evaporation in
mm/day (Table 4)

Table 2. Values of weighting factor (W) for the effect of radiation on ETo at different
temperature and altitudes.

ALT TEMPERATURE (oC)


(m) 12 14 16 18 20 22 24 26 28 30 32 34 36 38
0 - 499 .58 .61 .64 .66 .68 .71 .73 .75 .77 .78 .80 .82 .83 .84
500-999 .60 .62 .65 .67 .70 .72 .74 .76 .78 .79 .81 .82 .84 .85
1000 .61 .64 .66 .69 .71 .73 .75 .77 .79 .80 .82 .83 .85 .86

Table 3 Mean daily duration of maximum possible sunshine hours (N) for different months,
and for 15oN and 20oN latitudes.

MONTH LATITUDE MONTH LATITUDE

15oN 20oN 15oN 20oN

Jan 11.3 11.0 Jul 12.9 13.2


Feb 11.6 11.5 Aug 12.6 12.8
Mar 12.0 12.0 Sep 12.2 12.3
Apr 12.5 12.6 Oct 11.8 11.7
May 12.8 13.1 Nov 11.4 11.2
Jun 13.0 13.3 Dec 11.2 10.9

Table 4. Extra terrestrial radiation (Ra) expressed in equivalent evaporation in mm per day.

26
ALT JAN FEB MAR APR MAY JUN JUL AUG SEPT OCT NOV DEC
m
14 12.4 13.6 14.9 15.7 15.8 15.7 15.7 15.7 15.1 14.1 12.8 12.0
16 12.0 13.3 14.7 15.6 16.0 15.9 15.9 15.7 15.0 13.9 12.4 11.6
18 11.6 13.0 14.6 15.6 16.1 16.1 16.1 15.8 14.9 13.6 12.0 11.1
20 11.2 12.7 14.4 15.6 16.3 16.4 16.3 15.8 14.8 13.3 11.6 10.7
22 10.7 12.3 14.2 15.5 16.3 16.4 16.4 15.8 14.6 13.0 11.1 10.2

Example 2. Calculate the ETo for October at 15 oN and at altitude of 18 m using the
data below:

Tmax or maximum air temperature = 28oC


Tmin or minimum air temperature = 16oC
c or the adjustment factor = 1.1
n or actual duration of bright sunshine = 8 hrs

The procedure in calculating ETo will be:

a) Obtain the value of W from Table 2. At mean air temperature of 22 oC and at


altitude of 18m (0-499 m), W = 0.71

b) Obtain the value of N from Table 3. For the month of October and at 15oN
latitude, N = 11.8 hrs

c) Obtain the value of Ra from Table 4. At latitude 18oN, Ra = 13.6 mm/day

d) Solve for Rs

Rs = (0.25 + 0.5 n/N) Ra


= (0.25 + 0.5 8hrs/11.8 hrs) 13.6 mm/day
Rs = 8.0 mm/day

e) Calculate ETo

ETo = 1.1 [(0.71)(8.0 mm/day)]


= 1.1 (5.68 mm/day)
ETo = 6.25 mm/day

Calculations should preferably be made for each month or period for each year of
record rather than using mean radiation and mean temperature data based on several
years of record. A value of ETo can then be obtained to ensure that water requirements
will be met with a reasonable degree of certainty.

3. Penman equation:

This equation uses mean climatic data on temperature, humidity, wind and sunshine
duration or radiation. An adjustment factor for day and night time weather conditions is
included. The form of Penman equation is:

ETo = c[(WRn) + (1-W) f(u) (ea-ed)

where: ETo = reference crop ET in mm/day


c = adjustment factor to compensate for the effect of day and night
weather conditions (Table 5)
W = weighting factor which depends on temperature and altitude (Table 2)
27
Rn = net radiation in equivalent evaporation in mm/day
f(u) = wind related function

ea = saturation vapor pressure at mean air temperature in mbar


(Table 6)
ed = mean actual vapor pressure of the air in mbar

f(u) = 0.27 (1 + U/100)


U = 24-hr wind run in km/day at 2 m height

Rn = Rns - Rnl
Rns = net solar radiation
Rnl = net longwave radiation

Rns = (1-α ) Rs
α = reflection coefficient which is equal to 0.25 for most crops
Rs = (0.25 + 0.50 n/N)Ra

Rnl = f(T) f(ed) f(n/N)


f(T) = effect of temperature on Rnl (Table 7)
f(ed) = effect of vapor pressure on Rnl (Table 8)
f(n/N) = effect of the ratio of actual and maximum bright sunshine on
Rnl (Table 9)

The procedures presented above may seem rather complicated. This is due to the
fact that the formula contains components which need to be derived from measured
related data when no direct measurements of needed variables are available.
Nevertheless, compared with the other methods, the Penman equation presented, is
likely to provide the most satisfactory values of ETo.

Example 3. Using the data in Examples 1 and 2, calculate ETo using the Penman
equation. Hereunder are additional data:

U = 200 km/day
Uday = 3 m/sec
Uday/Unight = 4.0
ed = 20 mbar
RH = 75%
Min and Max Temp = 28 and 16 oC
Latitude = 15oN
Month = October

The procedure in calculating ETo will be as follows:

a) Derive the value of c from Table 5. By interpolation, c = 1.21 with Uday = 3 m per
sec and Rs = 8 mm per day (from Example 2),

b) W = 0.71 (based from Example 2)

c) With α = 0.25, and Rs = 8 mm/day, compute for Rns as follows:

Rns = (1- α) Rs
= (1-0.25) 8 mm/day
Rns = 6.0 mm/day

d) Calculate Rnl. From Table 7, f(T) = 15 (at T = 22oC); from Table 8 f(ed) =
28
0.14 (at ed = 20 mbar). Estimate f(n/N).

f(n/N) = 0.1 + 0.9 (n/N)


= 0.1 + 0.9 (8/11.8)
f(n/N )= 0.712

Therefore, Rnl = f(T) f(ed) f(n/N)


= (15)(0.14)(0.712)
Rnl = 1.495 mm per day

e) Calculate Rn

Rn = Rns - Rnl
= 6.0 - 1.495
Rn = 4.505 mm per day

f) Calculate f(u)

f(u) = 0.27 (1 + U/100)


= 0.27 (1 + 200/100)
f(u) = 0.81

g) Based from Table 6, ea = 26.4 mbar

h) Finally calculate ETo

ETo = c[(WRn) + (1-W) f(u) (ea-ed)]


= 1.21[(0.71)(4.505) + (1-0.71) (0.81) (26.4 -20)]
= 1.21 [(3.19855) + (1.50336)]
= 1.21 (4.70191)
ETo = 5.689 mm per day

Table 5. Adjustment factor (c) in presented Penman equation.

RHmax = 30% RHmax = 60% RHmax = 90%


Rs (mm/day) 3 6 9 12 3 6 9 12 3 6 9 12
Uday (m/sec) Uday/Unight = 4.0
0 .86 .90 1.00 1.00 .96 .98 1.05 1.05 1.02 1.06 1.10 1.10 1
3 .79 .84 .92 .97 .92 1.00 1.11 1.19 .99 1.10 1.26 1.33
6 .68 .77 .87 .93 .85 .96 1.11 1.19 .88 1.01 1.16 1.27
9 .55 .65 .78 .90 .76 .88 1.02 1.14
Uday/Unight = 3.0
0 .86 .90 1.00 1.00 .96 .98 1.05 1.05 1.02 1.06 1.10 1.10
3 .76 .81 .88 .94 .87 .96 1.06 .94 1.04 1.18 1.28
6 .68 .68 .81 .88 1.12 .77 .88 1.02 .86 1.01 1.15 1.22
9 .55 .56 .72 .82 1.10 .78 .92 1.06 1.18
.67 .79 .88 1.05

Uday/Unight = 2.0

0 .86 .90 1.00 1.00 .96 .98 1.05 1.05 1.02 1.06 1.10 1.10
3 .69 .76 .85 .92 .83 .91 .99 1.05 .89 .98 1.10 1.14
6 .53 .61 .74 .84 .70 .80 .94 1.02 .79 .92 1.05 1.12
9 .37 .48 .65 .76 .59 .70 .84 .95 .71 .81 .96 1.06

29
Uday/Unight = 1.0

0 .86 .90 1.00 1.00 .96 .98 1.05 1.05 1.02 1.06 1.10 1.10
3 .64 .71 .82 .89 .78 .86 .94 .99 .85 .92 1.01 1.05
6 .43 .53 .68 .79 .62 .70 .84 .93 .72 .82 .95 1.00
9 .27 .41 .59 .70 .50 .60 .75 .87 .62 .72 .87 .96

Table 6. Saturation vapor pressure (ea) in mbar as a function of mean air temperature in oC.
TEMP 16 18 20 22 24 26 28 30 32 34 36
(oC)
ea 18.2 20.6 23.4 26.4 29.8 33.6 37.8 42.4 47.6 53.2 59.4
(mbar
)

Table 7. Effect of temperature or f(T) on net longwave radiation (Rnl).

TEMP 16 18 20 22 24 26 28 30 32 34 36
(oC)
f(T)= σTk4 13.8 14.2 14.6 15.0 15.4 15.9 16.3 16.7 17.2 17.7 18.1

Table 8. Effect of vapor pressure f(ed) on net longwave radiation (Rnl).

ed mbar 10 14 18 22 24 26 28 30 32 34 36
f (ed) = 0.34 - 0.044 √ed .20 .18 .15 .13 .12 .12 .11 .10 .09 .08 .08

Table9. Effect of the ratio of actual and maximum bright sunshine hours f(n/N) on net
longwave radiation (Rnl).

n/N .05 .10 .15 .20 .25 .30 .35 .40 .45 .50 .55 .60
F(n/N) = 0.1 + 0.9 n/N .15 .19 .24 .28 .33 .37 .42 .46 .51 .55 .60 .64

4. Evaporation pan method:

Evaporation pan provides measurement of the integrated effect of radiation, wind,


temperature and humidity on evaporation from a specific open water surface. With proper
siting, the use of pans to predict crop water requirements for the period of 10 days or
longer is warranted. To relate pan evaporation (Epan) to ETo, empirically derived pan
coefficients (Kp) are given (Table 2) which account climate and pan environment. ETo
can be obtained as:

ETo = (Kp) (Epan)

where ETo = reference crop ET in mm/day


Kp = pan coefficient (Table 10)

30
Epan = pan evaporation in mm/day and represents the mean daily value of the
period considered
Example 4. Calculate ETo when the daily pan evaporation is 5.0 mm per day. The
pan is placed in a green cropped area at a windward side distance of 100 m. The
relative humidity is 75% and the wind is 200 km per day.

a) Determine from Table 10 the pan coefficient, Kp. Based from the given condition,
Kp = 0.8

b) Calculate ETo:
ETo = (Kp)(Epan)
= (0.8) (5.0)
ETo = 4.0 mm per day

It is likewise important to take note of the following when using the Epan
method:

 the storage of heat within the pan can be appreciable and may cause almost
equal evaporation during night and day; most crops transpire only during
daytime.
 the difference in water losses from pans and from crops can be caused by
differences in turbulence, temperature and humidity of the air immediately
above the surfaces
 heat transfer through the sides of the pan can occur (which may be severe
for sunken pans)
 the color of the pan and the use of screens will affect water losses; an
increase in Epan of up to 10 percent may occur when they are painted black;
screens mounted over pans will reduce Epan by up to 10 percent
 the siting of the pan and the environment influence the measured evaporation
amounts
 the level at which water is maintained in the pan is very important; resulting
errors maybe up to 15 percent when water levels in Class A pans fall 10 cm
below the accepted standard of between 5 and 7.5 cm below the rim
 turbidity of the water in the pan does not affect Epan data by more than 5
percent

Table 10. Pan coefficient (Kp) for class A pan for different groundcover and levels of mean
relative humidity and 24-hour wind.

Class A Case A: Pan Placed in Short Green Case B: Pan Placed in Dry Fallow
Pan Cropped Area Area
RH Mean, Low Medium High Low Medium High
% <40 40-70 >70 <40 40-70 >70
Wind Windward
km/day Windward side distance
side distance of dry fallow,
of green m
crop, m
Light 1 .55 .65 .75 1 .70 .80 .85
<175 10 .65 .75 .85 10 .60 .70 .80
100 .70 .80 .85 100 .55 .65 .75
1000 .75 .85 .85 1000 .50 .60 .70

Moderate 1 .50 .60 .65 1 .65 .75 .80

31
175-425 10 .60 .70 .75 10 .55 .65 .70
100 .65 .75 .80 100 .50 .60 .65
1000 .70 .80 .80 1000 .45 .55 .60

Strong 1 .45 .50 .60 1 .60 .65 .70


425-700 10 .55 .60 .65 10 .50 .55 .65
100 .60 .65 .70 100 .45 .50 .60
1000 .65 .70 .75 1000 .40 .45 .55

Very 1 .40 .45 .50 1 .50 .60 .65


Strong 10 .45 .55 .60 10 .45 .50 .55
>700 100 .50 .60 .65 100 .40 .45 .50
1000 .55 .60 .65 1000 .35 .40 .45

Summary

The effect of climate on crop water requirements is given by the reference crop
evapotranspiration (ETo) which is defined as "the rate of evapotranspiration from an extensive
surface of 8 to 15 cm tall, green grass cover of uniform height, actively growing, completely
shading the ground and not short of water." Four methods were presented to calculate ETo
using the mean daily climatic data for 30- or 10-day periods, which are the: Blaney-Criddle,
pan evaporation, radiaton, and Penman methods. ETo is expressed in mm per day and
represents the mean value over that period. Primarily, the choice of method must be based
on the type of climatic data available and on the accuracy required in determining water
needs.

Key Terms

The following terms are listed in the order they appear in the text. Define each.
Doing so will aid you in reviewing the material covered in this topic.
evapotranspiration wind run
reference crop evapotranspiration vapor pressure
radiation reflection coefficient
solar radiation evaporation
shortwave radiation pan coefficient
longwave radiation turbulence
extra terrestrial radiation turbidity

Questions for Review

1. What is reference crop evapotranspiration?

2. Enumerate the methods used in calculating reference crop evapotranspiration. Briefly


describe each method.

3. Fill up the table below with appropriate codes as follows: ** = measured data; 0 =
estimated data; X = not needed:

METHOD TEMPER- HUMIDITY WIND SUNSHINE EVAPO- ENVIRON-


ATURE DURATION RATION MENT
Blaney-
Criddle

32
Radiation

Penman

Pan
Evaporatio
n

4. When can the Blaney-Criddle method be used in estimating reference crop


evapotranspiration?

5. Why is it preferable that calculations of ETo be made for each month or period for each
year of record rather than using mean radiation and mean temperature data based on
several years of record when using the radiation method?

6. Enumerate at least five principles to consider when using the pan evaporation method of
calculating ETo.

7. Briefly explain why the Penman method is considered to provide the best estimates of ETo
values.

8. Calculate ETo using the four methods (Blaney-Criddle, Radiation, Penman, and Epan
method) based from the data given below:

Month: April
Latitude: 18oN
Tmax: 30oC
Tmin: 18oC
RH: 90 %
Epan: 5.0 mm per day
Windward side distance of green crop: 100 m
Epan location: short green cropped area
α or (reflection coefficient): 0.25
n or actual bright sunshine duration: 8 hrs
Uday/Unight: 2.0
U or Wind: 200 km per day

References
Brouwer, C. and M. Heibloem. 1986. Irrigation water needs. FAO Water Management
Training Manual No. 3. 55 pp.

De Vera, M. R., and S.T. Mancebo. Teacher's guide and student's laboratory manual on
irrigation and drainage. TPAE, EDPITAF, and DECS Project. 189 pp.

Doorenbos, J. 1976. Agro-meteorological field stations. FAO Irrigation and drainage paper.
94 pp.

Doorenbos, J. and W.O. Pruitt. 1984. Crop water requirements. FAO Irrigation and
Drainage Paper No. 24. 143 pp

Rosenberg, N.J., B.L. Blad, and S.B. Verma. 1983. Microclimate: The biological
environment. John Wiley & Sons, Inc. 495 pp.

33
Snyder, R. and C.G. Acosta. 1990. Water balance program for irrigation planning.
Unpublished paper. 20 pp.

PRACTICAL APPLICATIONS OF THE SOIL-PLANT-WATER RELATIONS


Table 5 shows the different soil textures and their corresponding physical properties.
The data shown in the table is very important in guiding the farmer or the irrigator in
determining how much water should be applied to his farm. It must be noted that for upland
irrigation, the amount of water that should be delivered to the farm should not exceed the
moisture-holding capacity or storage capacity of the soil within the depth of root zone of the
crop grown. This capacity is defined under the column “Total Available Moisture” or the range
between FC and PWP. Otherwise, if the amount of water applied exceeds the storage
capacity of the soil, deep percolation losses will occur which will result in leaching of fertilizer
and other soil nutrients. Hence, if the irrigator knows the moisture content of the soil
right before irrigation water application, he would know how much water to be supplied
in order to raise the moisture content to FC.

Another important aspect which the irrigator/farmer should know is the range within
the total available moisture which a particular crop is allowed to utilize. That is, how much of
the soil moisture be allowed to deplete FC without significant reduction in crop yield. This
range is commonly termed as Readily Available Moisture (RAM). Vegetable crops, like
cabbage, can utilize only 22-30% of the total available moisture and other crops, like corn,
sorghum, mungo and peanut, 60-70%. How long or how many days the crop will utilize this
RAM range will depend on its evapotranspiration. Furthermore, the duration of soil moisture
utilization by a crop depends on the soil moisture storage capacity. Thus, the frequency of
irrigation water application to a particular crop grown on a sandy loam soil is more if the same
crop is grown on a clay loam soil.

Table 4. Average effective root zone of common crops grown in deep,


well drained soil under average conditions.
        
CROP DEPTH cm)   CROP DEPTH cm)

Bush bean 45 Grape 200

34
Cabbage 60 Lettuce 30
Carrots 90 Melon 155
Cauliflower 60 Mustard 105
Celery 60 Onion 30
Corn 90 Pepper 90
Cotton 120 Radish 45
Cucumber 90 Sorghum 90
Eggplant 90 Squash 90
Garlic 60 Sweet potato (BNAS-51) 30
Sweet potato (Samar Big yellow) 20
     

Example:

1. Corn is grown on a sandy loam soil:


From Table 1, depth of root zone of corn (Drz) = 90 cm
Average evapotranspiration of corn = 4.74 mm/day
Assume RAM for corn = 60%
From Table 5, FC = 14% and PWP = 6%

2. To determine moisture content of the soil (Pw) every time irrigation water is
applied:

Total Available Mositure (TAM) = FC – PWP


= 14% - 6% = 8%

Range (R) within TAM which corn is allowed to utilize,

R = RAM x TAM = 0.6 x 0.08 = 4.8%

So, Pw = FC – R = 14% - 4.8% = 9.2%

This means that corn is allowed to utilize water from FC to 9.2% moisture
before irrigation water is applied. Beyond 9.2% the plant will suffer stress.

3. Number of days required to deplete soil moisture from FC to 9.2%:

FC−9 .2%
RAM d = x A s x D rz
RAM expressed in terms of depth, 100
From Table 5, As = 1.5
14−9. 2
RAM d = x 1 .5 x 90cm
100
= 6.48 cm or 64.8 mm

RAM d
Number of days = ET of Corn = (64.8 mm)/(4.74 mm/day) = 13.67 or 14 days

35
Irrigation Scheduling
The field irrigation schedule depends upon the availability of water, properties of soil,
weather conditions, and the crop to be irrigated. At field level, the irrigation schedule is based
on the field water balance and are usually expressed in depth of irrigation application (mm)
and the irrigation-application interval (days). The depth and the frequency of irrigation
application are dependent upon the available moisture in the rootzone, the rate of water use,
and the allowable percentage of depletion of this water.

Table 5. General physical properties of soils for irrigation and drainage


                 
TOTAL AVAILABLE MOISTURE
Infiltration Apparent
Total Pore By Weight % By Volume By Depth
and Specific
Soil Texture Space % FC (%) PWP (%) Pw = FC - % of Water
Permeability Gravity
(N) PWP Pv = PwAs cm/m (d)
(cm/hr) If (As)

Sandy 5 38 1.65 9 4 5 8 8
(2.5 - 25) (32 - 42) (1.55 - 1.8) (6 - 12) (2 -6) (4 - 6) (6 - 10) (7 - 10)

Sandy loam 2.5 43 1.5 14 6 8 12 12


(1.3 - 7.6) (40 - 47) (1.4 - 1.6) (10 -18) (4 - 8) (6 - 10) (9 -15) (9 -15)

Loam 1.3 47 1.4 22 10 12 17 17


(0.8 - 2.0) (43 - 49) (1.35 - 1.5) (18 - 26) (8 -12) (10 - 14) (14 - 20) (14 - 19)

Clay loam 0.8 49 1.35 27 13 14 19 19


(0.25 - 1.5) (47 - 51) (1.3 - 1.4) (23 - 31) (11 - 15) (12 - 16) (16 - 22) (17 - 22)

Silty clay 0.25 51 1.3 31 15 16 21 21


(0.03 - 0.5) (49 - 53) (1.3 - 1.4) (27 - 35) (13 - 17) (14 - 18) (18 - 23) (18 - 23)

Clay 0.05 53 1.25 35 17 18 23 23


(0.01 - 0.10) (51 - 55) (1.2 - 1.3) (31 - 39) (15 - 19) (16 - 20) (20 - 25) (20 - 25)
                 

The depth of irrigation application (d) including application losses is:

( p Sa )D
d (mm )=
Ea
The frequency of irrigation (fi) application expressed in irrigation interval of the
individual field is:

( p Sa )D
f i (days )=
ET
Where:
p = fraction of available soil-water permitting unrestricted
evapotranspiration
Sa = total available soil water, mm/m of soil depth
36
D = depth of root zone, m
Ea application efficiency, fraction
ET Evapotranspiration of crop, mm/day

The depth and frequency of irrigation application vary significantly over the growing
season due to crop growth stages, the consumptive use or the water-use rate, and the
season in which the particular crop is grown. In general, the irrigation schedule must be
based upon the evaluation of these factors locally.

In view of the wide range of local meteorological conditions and limited Philippine
data, it seems appropriate for a guide to provide general upper limit values of water-use rate
for general irrigation scheduling, rather than attempt to provide values for the different
regions. Table 6 was constructed using water-use rate of 5.1 mm/day for dry season and, 7.6
mm/day for an area surrounded by un-irrigated land during dry season crops.

These values represent the average water use rate for a 10-day period of maximum
water use in each season, and are valid only for systems in areas of general information.

Annual crops that are produced for their fresh vegetable product, such as cabbage,
pechay, lettuce, etc., show the effects of moisture stress at an earlier time than perennial
crops or annual crops whose product is seed or root. Therefore, the maximum allowable
depletion for cabbage, pechay, lettuce, etc., is 50% of the available moisture in the rootzone.
For other crops, 75% depletion is recommended.

The time (days) shown in Table 6 is calculated by dividing either 50% or 75% of AM
in the rootzone by the appropriate water-use rate.

Example (Application of the Guide of Irrigation Scheduling)

A farmer in Tarlac is growing garlic and onions. His soil is Luisita sandy loam. He
plans to grow both a wet season crop for which only occasional irrigation is necessary, and a
dry season crop which will need complete irrigation. He wants to irrigate with furrows. The
area has limited irrigation during the dry season.

A. How to determine the frequency of irrigation and the water to be applied per irrigation

during the rainy season.

1. From Table 4, determine the depth of rootzone. The depth of rootzone for onion and
garlic is 61.0 cm. For a light-textured soil such as Luisita sandy loam, use the deeper
rootzone unless field or other information indicates a shallow soil depth.

2. From Table 7, determine the available soil moisture per meter depth. The available
moisture for sandy loam is 120 cm/m.

3. Calculate the available moisture in the rootzone. The available moisture is 12 x


61/100 = 7.32 cm.

4. From Table 6, the daily-use rate for rainy season is 0.51 cm/day with a corresponding
depletion percentage of 75%. From the same table, the frequency of irrigation is
approximately 11 days and the water to be replaced is 5.6 cm.

5. The irrigation-application efficiency for furrow irrigation is estimated to be 60%.


Accordingly, the application should be 9.3 cm or (5.6 / 0.60 = 9.3cm).

37
B. How to determine the frequency of irrigation and the water to be applied per irrigation
during the dry season.

1. From Table 6, the design daily-use rate for the dry season is 0.76 cm/day (generally
un-irrigated area) with a corresponding depletion percentage of 75%.

2. From Table 6, the frequency of irrigation is approximately 7 days; and the water to be
replaced per irrigation is 5.3 cm.

3. If the irrigation-application efficiency is estimated to be 60%, then the application


should be 8.8 cm (5.3 /0.60 = 8.8 cm).

Table 6. Irrigation frequency and amount of three consumptive-use rates and two
depletion percentages.
         
Irrigation Irrigation
Available Net Water Net Water
Frequency at Frequency at
Moisture in Replaced each Replaced each
50% Depletion 75% Depletion
Root Zone (cm) Irrigation (cm) Irrigation (cm)
(days)a (days)b

Daily Water-Use Rates for Rainy Season - 0.51 cm/day


2.5 2.0 1.0 3.0 1.5
3.8 3.0 1.5 5.0 2.5
5.1 5.0 2.5 7.0 3.6
6.4 6.0 3.0 9.0 4.6
7.6 7.0 3.6 11.0 5.6
8.9 8.0 4.1 12.0 6.1
10.2 10.0 5.1 15.0 7.6
11.4 11.0 5.6 16.0 8.1
12.7 12.0 6.1 18.0 9.1
         
Daily Water-Use Rates for Dry Season in Generally irrigated Area - 0.64 cm/day
2.5 2.0 1.3 3.0 1.9
3.8 3.0 1.9 4.0 2.5
5.1 4.0 2.5 6.0 3.8
6.4 5.0 3.2 7.0 4.4
7.6 6.0 3.8 9.0 5.7
8.9 7.0 4.4 10.0 6.7
10.2 8.0 5.1 12.0 7.7
11.4 9.0 5.7 13.0 8.6
12.7 10.0 6.4 15.0 9.5
         
Daily-Use Rates for Dry Season in Un-irrigated Area - 0.76 cm/day
2.5 1.0 0.8 2.0 1.5
3.8 2.0 1.5 3.0 2.3
5.1 3.0 2.3 5.0 3.8
6.4 4.0 3.0 6.0 4.6
7.6 5.0 3.8 7.0 5.3
8.9 5.0 4.0 8.0 6.1
38
10.2 6.0 4.6 10.0 7.6
11.4 7.0 5.3 11.0 8.4
12.7 8.0 6.1 12.0 9.1
         
a
recommended for leafy vegetables
b
recommended for other crops

C. How to determine the water flow for either season.

1. As illustrated in the foregoing example, rainy-season irrigation must be completed


within 11 days; and the application should be 9.3 cm. Consequently, the continuous
flow rate at the field must be at least 0.85 cm/day (9.3 / 11 = 0.85). The flow rate of 10
mm/ha/day may be used or an equivalent of 1.13 li/sec.

2. On the other hand, dry-season irrigation must be completed within 7 days and the
application should be 8.8 cm. The corresponding flow rate at the field must be at least
1.3 cm/day. The flow rate of 13 mm/ha/day is equivalent to 1.42 li/sec.

3. If it is designed to irrigate only 12 hours per day, these equivalent flow rates must be
doubled, and so on.

4. For bigger areas, these flow rates must be multiplied by the number of hectares.

Example A.

Given:
Plant : Onion Root zone depth: 0.61 m
Soil : Sandy Loam Irrigation Method: Furrow (60% efficiency)
ET : 5.1 mm/day (wet season)
7.6 mm/day (dry season)

a. Depth of irrigation application


b. Frequency of Irrigation Application (Wet season).
c. Frequency of Irrigation Application (Dry season).

Furrow-irrigation efficiency : 60%

Table 8. Water, soil, and climatic requirements of some upland crops


                  
CHARACTERISTIC AND WATER NEEDS   SOIL REQMT   CR
Average
Seasonal
Growing Depth of Consumptiv Effective
Water
Crop Period Root e Use Soil Texture RAM (%) Rainfall
Requirement
(day) Zone (mm/day) (%)
(cm)
(cm)

39
Corn 90 - 120 80 - 100 4-7 60 Silt loam to loam 60 - 70 90 - 100
Rice 100 - 130 15 - 35 4-6 100 clay loam 10 - 20 60 - 100
Sugarcane 270 - 365 100 - 120 4-8 150 - 250 fine sandy loam 20 - 60 70 - 100
Tobacco 90 - 120 90 5-8 40 - 60 silt loam to clay loam 20 - 68 90 - 100
Cotton 150 - 180 120 -125 5-8 70 - 130 sandy loam to clay loam 60 - 70 60 - 100
Cassava 180 - 200 50 - 75 4-6 100 - 150 sandy loam to clay loam 50 - 60 60 - 100
Sweet potato 60 - 120 100 - 150 3-6 46 sandy loam to clay loam 60 - 70 *
Irish potato 100 - 150 40 - 60 * 40 - 50 loam to sandy loam * 40 - 100
Fruit trees 240 - 365 100 - 500 7 90 - 120 loam to sandy loam 40 - 60 80 - 100
Vegetables 30 - 60 40 - 50 4-6 30 - 50 sandy loam to loam 20 - 30 *
                  

CHAPTER 5
CONVEYANCE AND MEASUREMENT OF IRRIGATION WATER

Improving water-use efficiency

In general, the term efficiency is used to quantify the relative output obtainable from a given
input. Referring to the use of water in irrigation, efficiency may be defined in various ways,
depending on the nature of the inputs and outputs to be considered. An economic criterion of
efficiency, for example, might be the financial return obtained from irrigation in relation to the
investment made in the water supply. The problem here is that costs and prices fluctuate from
year to year and vary widely from place to place. Another problem is that some of the costs of
irrigation and some of the benefits cannot easily be quantified in tangible economic or
financial terms, especially where a market economy is not yet fully developed. Often, only the
short-term costs and immediate benefits are seen, whereas the long-term advantages or
disadvantages are not fully realized at the outset. How is it possible to assess the economic
value, for instance, of saving the population of a region from the potential effects of a drought,
if the probability or severity of future drought events is not known? To some degree, therefore,
it is necessary to operate in a state of uncertainty.

In more restricted technical terms, what irrigation engineers often call conveyance efficiency
is defined as the net amount of water delivered to a farm, as a fraction of the amount taken
from some source. The difference between the two amounts represents the seepage and
evaporative losses incurred en route from source to field. Not generally considered in the term
conveyance efficiency is the possible loss of water quality through pollution - such as that
caused by wading animals or by human use of the canal water for washing and waste
disposal.

The term on-farm application efficiency or field application efficiency generally refers to
the fraction of the water volume applied to a farm or a field that is "consumed" by a crop,
relative to the amount applied. Crop consumption consists of the amount of water actually
absorbed by the crop, most of which is generally transpired to the atmosphere (only a small
fraction, often less than 1 percent, being retained in the vegetative biomass). There is much
evidence that, in a given climate, the growth of many crops is directly related to the amount of
water they transpire. The explanation is that both carbon dioxide (CO 2) for photosynthesis and

40
transpiration occur concurrently through the same stomatal openings in the leaves, so the two
processes should be roughly proportional.

In actual practice, however, the volume of water reported to be consumed in the field consists
of evapotranspiration rather than of transpiration alone. Evapotranspiration includes, in
addition to the amount of water transpired by the plants, the amount evaporated directly from
the soil surface without being taken up by the plants. In addition, evapotranspiration often
includes the amount of water intercepted by the foliage (e.g. under overhead sprinkler
irrigation) and evaporated without ever entering either the soil or the plant. The reason why
the term evapotranspiration is taken to be consumptive use is that, in practice, direct
evaporation is difficult to measure separately from transpiration, so the two terms are lumped
together merely for the sake of convenience.
Clearly, however, much of the water evaporated without entering the plant is consumed non-
productively. Therefore, any method of irrigation that minimizes evaporation (but not
transpiration) is likely to increase the efficiency of water utilization by the crop. Some of the
irrigation methods described in this publication are capable of doing just that: they introduce
water directly into the root zone without sprin-kling the foliage or wetting the entire soil
surface. Such partial-area irrigation methods offer the additional benefit of keeping the greater
part of the soil surface (between the rows of crop plants) dry. This discourages the growth of
weeds, that would otherwise not only compete with crop plants for nutrients and moisture in
the root zone and for light above ground, but also hinder field operations and the control of
pests.

Even with total evapotranspiration considered as consumptive use, field application efficiency
in most traditional irrigation schemes is still very low: typically less than 50 percent and often
as low as 30 percent. Excessive application of water generally entails losses due to surface
runoff from the field as well as to deep percolation below the root zone within the field. Both
runoff and deep percolation losses are difficult to control under flood or furrow irrigation,
where a large volume of water is applied all at once. They can, however, be minimized where
a controlled volume of water is applied at a slow rate over an extended period of time directly
to the root zone.

Even with the best irrigation practices, however, field application efficiency values cannot
attain 100 percent. Nor should that be the aim, since a certain fraction of the water applied
must be allowed to seep downwards and leach the salts that would otherwise accumulate in
the root zone.1 However, with careful management, field water application efficiency values
approaching 90 percent are possible, and values of 80 percent are practicable, by some of
the methods described in this chapter.

A word of warning is in order at this point. No irrigation method or technology in itself


guarantees the attainment of high efficiency. How the system is operated is all important. With
poor management, even the most sophisticated system can result in water loss and
inefficiency. Only knowledgeable, experienced and caring management can ensure that
appropriate irrigation systems achieve their full potential benefits (Figure 8).

41
FIGURE 8. Typical crop root distributions

Quite different from strictly technical criteria of efficiency is the physiological index, known as
crop water-use efficiency. The relevant measure here is the response of the crop to irrigation,
not in percentage terms but as total biomass produced (above-ground dry matter) per unit
mass of water taken up by the crop. Since, as mentioned above, well over 90 percent of the
water taken up by plants in the field is normally transpired, crop water-use efficiency is in
effect the reciprocal of what has long been known as the transpiration ratio. The latter is
defined as the ratio of the amount of water transpired to the amount of dry matter produced
(tonnes per tonne). That ratio can be of the order of 1 000 or more in a dry climate of high
evaporative demand.

An alternative way to characterize crop water-use efficiency is in terms of the marketable crop
produced per unit volume of water. This expression is identical to the above-ground biomass
in the case of crops grown and harvested for forage, but it is quite different where the
marketable product is only the fruit, seed or fibre. Generally, but not always, the yield of such
products is proportional to total growth, hence also to transpiration.
To maximize crop water-use efficiency, by either of the above criteria, it is necessary both to
conserve water and to promote maximal growth. The former requires minimizing losses
through runoff, seepage, evaporation and transpiration by weeds. The latter task includes
planting high-yielding crops well adapted to the local soil and climate. It also includes
optimizing growing conditions by proper timing and performance of planting and harvesting,
tillage, fertilization and pest control. In short, raising water-use efficiency requires good
farming practices from start to finish.

42
Summary of ways to improve water-use efficiency
Conservation of water
Reduce conveyance losses by lining channels or, preferably, by using closed conduits.
Reduce direct evaporation during irrigation by avoiding midday sprinkling. Minimize foliar
interception by under-canopy, rather than by overhead sprinkling.
Reduce runoff and percolation losses due to overirrigation.
Reduce evaporation from bare soil by mulch-ing and by keeping the inter-row strips dry.
Reduce transpiration by weeds, keeping the inter-row strips dry and applying weed control
measures where needed.
Enhancement of crop growth
Select most suitable and marketable crops for the region.
Use optimal timing for planting and harvesting.
Use optimal tillage (avoid excessive cultivation).
Use appropriate insect, parasite and disease control.
Apply manures and green manures where possible and fertilize effectively (preferably by
injecting the necessary nutrients into the irrigation water).
Practise soil conservation for long-term sustainability.
Avoid progressive salinization by mon-itoring water-table elevation and early signs of salt
accumulation, and by appropriate drainage.
Irrigate at high frequency and in the exact amounts needed to prevent water deficits, taking
account of weather conditions and crop growth stage.

Finally, all the above indexes of efficiency may be combined in a single concept, the overall
agronomic efficiency of water use, Fag:

where P is crop production (total dry matter or the marketable product, as the case may be)
and U is the volume of water applied.
As only a fraction of the applied water is actually absorbed and utilized by the crop, it is
necessary to consider the various components of the denominator U:

U = R + D + Ep + Es + Tw + Tc (2)

where R is the volume of water lost by runoff from the field, D the volume drained below the
root zone (deep percolation), Ep the volume lost by evaporation during the conveyance and
application to the field,2 Es the volume evaporated from the soil surface (mainly between the
rows of crop plants), T w the volume transpired by weeds, and T c the volume transpired by the
crop. All these volumes pertain to the same unit area.

Accordingly:

Under flood irrigation as commonly practised in river diversion schemes, excessive water
application often results in considerable runoff, evaporation from open water surfaces and

43
transpiration by weeds. In the experience of the author, these losses commonly amount to 20
percent or even 30 percent of the water applied. In addition, the loss of water due to
percolation below the root zone may be of the order of 30 percent or even 40 percent of the
water applied. Consequently, the fraction actually taken up by the crop is often below 50
percent and may even be as low as 30 percent.

FIGURE 9
The water balance of a field

If runoff and direct evaporation of free water are prevented, and if evaporation from the soil
surface is minimized (as under partial-area irrigation that avoids wetting the areas between
rows) and weeds are effectively controlled; and if, furthermore, water is applied in measured
quantities commensurate with crop requirements so as to avoid excessive percolation, all the
losses can be reduced to less than 20 percent of the water applied. Irrigation efficiency can
then attain or even exceed 80 percent.

Finally, and no less important, the numerator of the equation (namely, the yield attainable)
can be greatly enhanced by judicious selection of crops and varieties, optimal fertilization and
tillage and proper timing of planting and harvesting. All in all, the agronomic efficiency of
water use in irrigated farming can be significantly increased relative to the low efficiency
characteristic of traditional practice.

1
Irrigation water, even if it is of high quality, invariably contains some salts, and these are
mostly left behind as crop roots absorb water from the soil.
2
Evaporation may take place from exposed bodies of water in the case of surface irrigation,
or from wind-drift and intercepted water in the case of sprinkler irrigation.

Irrigation efficiencies

Not all water taken from a source (river, well) reaches the root zone of the plants. Part of the
water is lost during transport through the canals and in the fields. The remaining part is stored
in the root zone and eventually used by the plants. In other words, only part of the water is
used efficiently, the rest of the water is lost for the crops on the fields that were to be irrigated.

Figure 24 shows the irrigation water losses in canals; these are due to:

1. Evaporation from the water surface


2. Deep percolation to soil layers underneath the canals
44
3. Seepage through the bunds of the canals
4. Overtopping the bunds
5. Bund breaks
6. Runoff in the drain
7. Rat holes in the canal bunds

Figure 24. Irrigation water losses in canals

Figure 25 shows the irrigation water losses in the field; these are due to:

1. Surface runoff, whereby water ends up in the drain


2. Deep percolation to soil layers below the root zone

Figure 25. Irrigation water losses in the field

45
To express which percentage of irrigation water is used efficiently and which percentage is
lost, the term irrigation efficiency is used.

The scheme irrigation efficiency (e in %) is that part of the water pumped or diverted
through the scheme inlet which is used effectively by the plants. The scheme irrigation
efficiency can be sub-divided into:

- the conveyance efficiency (ec) which represents the efficiency of water transport in canals,
and

- the field application efficiency (ea) which represents the efficiency of water application in
the field.

The conveyance efficiency (ec) mainly depends on the length of the canals, the soil type or
permeability of the canal banks and the condition of the canals.

In large irrigation schemes more water is lost than in small schemes, due to a longer canal
system. From canals in sandy soils more water is lost than from canals in heavy clay soils.
When canals are lined with bricks, plastic or concrete, only very little water is lost. If canals
are badly maintained, bund breaks are not repaired properly and rats dig holes, a lot of water
is lost.

Table 7 provides some indicative values of the conveyance efficiency (ec), considering the
length of the canals and the soil type in which the canals are dug. The level of maintenance is
not taken into consideration: bad maintenance may lower the values of Table 7 by as much
as 50%.

Table 7. INDICATIVE VALUES OF THE CONVEYANCE EFFICIENCY (ec) FOR


ADEQUATELY MAINTAINED CANALS

46
Earthen canals Lined canals
Soil type Sand Loam Clay
Canal length
Long (> 2000m) 60% 70% 80% 95%
Medium (200-2000m) 70% 75% 85% 95%
Short (< 200m) 80% 85% 90% 95%

The field application efficiency (ea) mainly depends on the irrigation method and the level
of farmer discipline. Some indicative values of the average field application efficiency (ea) are
given in Table 8. Lack of discipline may lower the values found in Table 8.

Table 8. INDICATIVE VALUES OF THE FIELD APPLICATION EFFICIENCY (ea)

Irrigation methods Field application efficiency


Surface irrigation (border, furrow, basin) 60%
Sprinkler irrigation 75%
Drip irrigation 90%

Once the conveyance and field application efficiency have been determined, the scheme
irrigation efficiency (e) can be calculated, using the following formula:

with

e = scheme irrigation efficiency (%)


ec = conveyance efficiency (%)
ea = field application efficiency (%)

A scheme irrigation efficiency of 50-60% is good; 40% is reasonable, while a scheme


Irrigation efficiency of 20-30% is poor.

It should be kept in mind that the values mentioned above are only indicative values.

EXAMPLE

QUESTION:

Determine the project irrigation efficiency for a scheme with a long canal system. The canals
are constructed in heavy clay and the irrigation method is furrow irrigation. Maintenance of the
canals is adequate.

ANSWER:

Estimate the conveyance efficiency, using Table 7: ec = 80%.

Determine the field application efficiency, using Table 8: ea = 60%.

47
Calculate the scheme irrigation efficiency, using the formula:

Thus, the scheme irrigation efficiency e = 80 x 60/100 = 48% or approximately 50%. This is
considered a fairly good scheme Irrigation efficiency, for a surface Irrigation system.

Irrigation Efficiency Formulas


Conveyance efficiency

Ec =
( )
Wf
Wr
100

Application efficiency (Ec)

Ec =
( )
Ws
Wf
100

Wf = Ws + Rf + Df

( )
W f −( R f + D f )
Ec = 100
Wf
Water-use efficiency

Eu =
( )
Wu
Wd
100

Water Storage Efficiency

Es =
( )
Ws
Wn
100

Water Distribution Efficiency

( dy ) 100
Es = 1−
Consumptive use efficiency

Ecu =
( )W cu
W dep
100

Open Channels
- Trapezoidal, rectangular, triangular channels.
- Design Problems

Pls see powerpoint

48
CHAPTER 6
IRRIGATION METHODS
49
METHODS OF IRRIGATION WATER APPLICATION

Selecting the methods most suitable for applying water is important. Misuse of
irrigation water can cause soil erosion, water-logging, and a build-up of soil salinity.

The system of passing water from field to field, as commonly used in lowland rice
cultivation, is not possible for upland crops. Water for upland-crop irrigation must be supplied
from farm ditches, and a much denser network of water distribution systems is necessary.
Irrigation is done at relatively longer intervals with a sizeable stream of water.

Each method of irrigation is suited to a certain set of limiting conditions which governs
its use. A thorough understanding of soil, topography, degree or extent of land preparation,
crops grown, water supply, and other factors which can affect irrigation will be helpful in
selecting the proper method.

Irrigation water can be applied to the uplands in any of the following general ways:
(1) by overhead irrigation, wherein the soil is moistened in much the same way as
rain;
(2) by furrows, which wet only a part of the ground surface;
(3) by flooding, which wet all the land surface;
(4) by drip irrigation, wherein the water is directed to the base of the plant; and
(5) by sub-irrigation, wherein the surface is rarely wet since water is supplied from
the soil underneath.

OVERHEAD IRRIGATION

Watering Can. The simplest piece of overhead irrigation equipment is the watering
can shown in Figs 1 and 2, commonly used in small-scale upland farming. Since the water is
carried by hand, this method is limited to small plots with easily accessible source of water.
The size of the plot depends largely upon its distance from the source, and the time that it
takes to fill the can at the source. One man with a plot adjacent to an easily accessible canal,
river, or shallow well could manage a plot of about 500 m 2. If he has to fetch his water from a
point 100 m away, he could only manage about half this area.

Hose pipe. Wherever, there is a piped water-distribution system, it is possible to


connect a hose pipe to a tap or outlet, provided there is sufficient pressure in the water as it
emerges from the hose pipe (Figs 3 and 4).

This method is being practiced in the tobacco, cotton, garlic, and onion plantations of
the Ilocos Region. The water is pumped from shallow wells with the use of portable pumps
connected to a 2-inch pipe, 50 m to 100 m long.

50
Figure 1. The watering can, the simplest overhead watering equipment

Figure 2. Watering can as a form of irrigation

3a 3b
Figures 3a. Tobacco on ridges watered with hose and Figure 3b. A modified hose
irrigation in the form of spray

Sprinkler Irrigation. This method is the application of water to the surface of the soil
in the form of spray, simulating that of rain. The rain is produced by the flow of water under

51
pressure through small orifices or nozzles. The pressure is usually provided by pumping. With
careful selection of nozzle sizes, operating pressure, and sprinkler spacing, the amount of
irrigation water required to refill the crop-root zone can be applied nearly uniformly at a rate
to suit for the infiltration rate of the soil, thereby, obtaining efficient irrigation.

Sprinkler irrigation can be used for almost all crops (except rice, tobacco, and jute)
and on most soils. It is best suited to medium and large farms of 10 ha and above. This
method is usually not suitable for very fine textured soils (heavy clay soils), where the
infiltration rates are less than about 4 mm/hr. The method is particularly suited to sandy soils
that have high infiltration rate.

In many large scale sugarcane lands all over the country, sprinkler irrigation is used
(Fig 5).

Figure 5. An overhead sprinkler system irrigating sugarcane

Some of its advantages are the following:

1. It generally offers the only method of obtaining adequate distribution of water on


certain rolling lands where leveling for surface irrigation is not feasible.
2. It is suitable where depth of the soil is limited by a hardpan or other restricting layers.
3. It is suitable on porous soil, such as sands, where the water penetrates so rapidly that
irrigation by other methods gives excessive losses by deep percolation.
4. It is adapted to light application of water for shallow rooted crops, germination of
seeds, and during the seedling period.
5. It eliminates the need for farm ditches, and more area is available for crop production.
6. It may be designed for a smaller flow of water and, therefore, desirable over some
other methods. It is an economical method of irrigation where annual requirement is
low.

Some disadvantages of sprinkler irrigation are as follows:

1. It entails high initial cost of equipment.


2. Operating costs are usually higher than irrigation by surface methods.
52
3. Moving the portable pipes across muddy fields is hard and dirty work.
4. Under certain climatic conditions, spread of diseases may be encouraged.
5. Larger evaporation losses occur because sprinklers wet the entire soil surface, as
well as the leaves of the plants.
6. Winds disturb the flow pattern resulting in unequal distribution of irrigation water.
7. Mechanical difficulties are to be expected.
8. Sprinkling water containing an appreciable amount of salt may result in burn or death
of the plant leaves.

The most common type of portable sprinkler is the rotating head sprinkler, consisting
of a head with one or two nozzles, which is rotated slowly by the action of the water passing
through and which waters a roughly circular piece of land around the sprinkler (Fig 6).
Rotating sprinklers operate under a wide range of pressures and discharges. For every
operating pressure, there is optimum nozzle diameter to give the best water dispersion.
Sprinklers are classified broadly into three groups according to their operating pressures, and
these are given together with other characteristics in Table 1.

Figure 6. A rotating head sprinkler irrigating green beans

Table 1. Typical characteristics of rotating sprinklers


according to their operating pressures
       

LOW MEDIUM HIGH


CHARACTERISTICS
PRESSURE PRESSURE PRESSURE

Operating pressure (atm) 1-2 2-5 5 - 10

Nozzle diameter (mm) 1.5 - 6 6 - 20 20 - 40

Discharge (li/sec) 0.06 - 1 0.25 - 10 10 - 50

Diameter of Coverage (m) 6 - 35 25 - 80 80 - 140

Sprinkler Spacing (m) 9 - 18 18 - 54 54 - 100


       

53
Low pressure systems are used for irrigating orchards and tree crops below the leaf
canopy, for soils with high infiltration rates, and for small areas. Medium-pressure systems
cover larger areas and are generally used for irrigating sugarcane and tree crops above the
tree canopy.

A sprinkler irrigation is also designed to provide a calculated depth of water at a fixed


rate of application. The infiltration rate is determined from the infiltration characteristics of the
soil. Water infiltration rates for overhead irrigation are given in Table 2. These rates are
slightly lower than those for surface irrigation as there should be no ponding of water at the
soil surface.

Table 2. Intake rates for overhead irrigation


   
Soil Texture Intake Rate (mm/hr)

Clay 1-5
Clay loam 6-8
Silt loam 7 - 10
Sandy loam 8 - 12
Sand 10 - 25
   

A simple sprinkler layout is shown in Fig. 7. The application efficiency of the sprinkler
system is usually 70% to 80%. Shown in Table 3 is an example of specifications and
calculations for a small sprinkler irrigation system.

54
Figure 7. Layout plan for sprinkler irrigation system. The laterals are moved to successive
positions
up on one side of the main and then down on the other.

Table 3. Example of specifications and calculations for a small sprinkler irrigation system.

         
DETAILS UNITS POSSIBLE ARRANGEMENTS
      (1) (2)
Sprinkler Specifications
Operating pressure atm 2.00 3.00
Nozzle diameter mm 4.00 4.50
Discharge per sprinkler (q) li/sec 0.25 0.37
Spacing (s) m 9x9 12 x 12
Application rate (r) mm/hr 10.40 9.50

Field Details
Number of sprinklers on lateral (n)
n = ( 120 / s) 13.00 10.00
Number of lateral positions (p)
p = (160 / s) 18.00 13.00
Application time,
to apply 75 mm gross
t = 75 / r hr 7.20 7.90
Add time for moving equipment hr 0.70 0.70
Total time for one position hr 7.90 8.60
Total time for two positions per day hr 15.80 17.20
Number of days required for complete
irrigation, (p / 2) days 9.00 6.50
Pump supply required (Q = q . n) li/sec 3.25 3.70
         

55
SYSTEM COMPONENTS

56
SPRINKLER IRRIGATION

Introduction

Since the end of World War II, the development of Sprinkler Irrigation has been very
extensive. One of the factors that helped in the successful development of Sprinkler Irrigation
was the introduction of aluminum pipes. Other factors which were important in the
development of Sprinkler Irrigation were the improvements in the construction and water
distribution of the sprinklers, improvements in the construction of couplings, and the
development of various fitting for use in sprinkling irrigation systems.

Every sprinkler system is composed of a pipe network and of sprinklers. The pipes
convey the water and supply it to all the operating sprinkler at the correct pressure head. At
the nozzle of the sprinklers the pressure head of the water is converted to a velocity head.
The water flows out of the nozzle in the form of a jet.

The Main Elements of a Sprinkler System are:

1. The Source of Water ~ sprinkler system requires sources of water free of debris that
will clog the sprinkler. The common sources are wells, irrigation canals, rivers, and
lakes. Water can obtained from the source by gravitation, but sometimes pumping is
required.

2. Main Pipe Line ~ the main pipelines convey the water from the water source to the
irrigated fields. The pipelines are usually steel or asbestos cement. More recently
polyvinylchloride (PVC) has also been used.

3. Submains ~ These pipe lines branch off from the main pipe lines into the irrigated
fields and supply water to the sprinkler laterals through risers and valves spaced at
regular intervals. The submains are usually of steel, asbestos cement, or PVC. They
follow the boundaries of the irrigated plot, but sometimes they are run along the
center lines of the fields.

4. Sprinkler Laterals ~ these conveys the water, through suitable couplings are risers,
spaced at regular intervals, to the sprinklers. The sprinkler laterals are generally
aluminum and they can be transported from one irrigation position to another at he
end of an irrigation period.

5. Sprinklers ~ the most common type of sprinklers in use today is the slow revolving
sprinklers. Other types in use are the sprinklers “guns”, fixed head (garden)
sprinklers, and whirling sprinklers.

Types of Sprinkler Systems:

 Classification According to Portability


1. Stationary (Permanent) Sprinkler System
In a stationary (permanent) system all elements, starting with the water
source and including the sprinklers, are fixed in place. Such a system is
57
generally very expensive. Another disadvantage of such a system is that the
pipes and risers may interfere with field operations. Its main advantage is that
the expenses involve in operating the laterals and sprinklers are smaller.
Installations are high in initial cost but low in costs for operation. They are use in
dense orchards like banana plantation.

2. Semi-portable Sprinkler System


This system is composed of fixed mains and sub-mains, made of
Steel, asbestos cement or PVC and of light portable aluminum or plastic
sprinkler laterals. The semi-portable sprinkler system is the most widely used at
present. It is used for irrigating field and industrial crops, fruit plantations, and
vegetable gardens.

3. Portable Sprinkler System


In a portable sprinkler system all parts, including mains, sub-main
and sprinkler lateral are portable and movable. Portability of the main and
sub-main lines is important in regions where only supplemental irrigation is
practiced. Supplemental irrigation, meaning a small number of application
during the growing season, is used for winter crops in dry regions of irregular
rainfall pattern.

 Classification According to Over-Under Foliage


Basically, sprinkler irrigation is a method in which water is applied above the
foliage. The risers should, therefore, behind enough to clear the plants, so that the
water will be distributed freely and without interference. Part of it is intercepted by the
foliage and reaches the ground after some delay, during which evaporation may
occur. In fruit orchards both under-tree and over-tree sprinkling is practice. The
system selected in any case depends on local conditions.

 Under-tree Sprinkling
The basic requirement for under-tree sprinkling is to have spacing of
trees
Which permit of the sprinkler laterals. An under-tree sprinkler irrigation system, as a
rule, is a semi-portable system with stationary main and sub-main lines and portable
aluminum or plastic laterals. The laterals are place mid-way between the lines of
trees, so that sprinklers operate at the center of squares formed by each four
adjacent trees.
In some region it is suspected that over-tree sprinkling causes rotting
of oranges, bananas and vines. It is evident that in all cases of possible damage to
trees or fruits, a system of under-tree sprinkling should be designed. The low riser in
itself is not very helpful if the angle of exit of the jet is high, and cannot prevent water
from hitting the foliage.

 Over-tree Sprinkling
This method of sprinkling is used mostly in orchards where the close
spacing of trees makes it difficult, or impossible, to transport laterals. In this
case, the sprinkler system is stationary. The sprinkler laterals are usually
steel or plastic pipes. This type of sprinkler is expensive, but irrigation
operations are very simple, consisting only of the opening and closing of
valves. Even this operation can be automated. Because of the initial high
cost, medium and high pressure sprinkler are generally used over-tree
sprinkler system.
Sometimes, although not very frequently, semi-portable over-tree sprinkler system are

used in orchards, having wide tree spaces. In this cases, both laterals and tall riser are

58
portable. With a water source that supplies medium and high pressures, thus enabling

larger lateral spacing to be used, the system described above may be relatively cheap,

even though labor cost are increase.

 Special System

 Automatic Hydraulic Control


In Israel automation generally means the introduction of control
facilities in conjunction with stationary or solid set systems of the
conventional type. The simplest way of automating a system is through the
use of automatic metering valves, each controlling one plot. The water
application is set with a dial on the meter, and the valve will shut
automatically at the end of the application, which id uninfluenced by pressure
variations or durations of irrigation. The valve is controlled hydraulically.
Further advances in automation have been made to minimize labor
and save water when large, and often-distant areas are irrigated. As an
example, a number of “Follow-Thru” systems, can be controlled electronically
by a single control mechanism, and a number of the latter can further be
controlled by a master panel at the center of a farm. Ultimate automation is
achieved by the use of sensors, such as tensiometers, which are present to
activate the irrigation system at required moisture leveling the soil profile.

 Self-Propulsion
In the U.S. and Europe automation is generally achieved by
mechanical means. One example is the self-propelled lateral system, with
wheels, which is activated by a power unit. Some propelled machines
available commercially are designed to move continuously across the field at
desired speeds, normally 0.15 and 1.20 m/min.
Recently, the Center-Pinot method of irrigation has become very
widely used. A pipeline of aluminum, up to 400m in length, on which the
sprinklers are installed, has mobile supports or towers on wheels at regular
internals. This pipeline in anchored to a center-pinot structure at one end
and rotates around it with a selected speed. The wetted area is obviously
circular, but recent developments make it possible to irrigate a more or less
square shape. As more Experience has been gained, this method of
irrigation has become quite dependable, requiring minimum labor for
operations, maintenance, and repairs, In some cases the method has allows
the successful irrigation of land previously considered impossible for growing
crops, and often very high yields have been achieved.

Sprinkler Types and their Characteristics


The various Sprinklers maybe divided into two main groups:

Group A Group B
1. Revolving Sprinklers 1. Perforated Pipe
2. Whirling Sprinklers 2. Oscillating rain Pipe
3. Fixed-head (Garden) Sprinklers
4. Sprinkler guns
General Description of Group A

The Sprinklers are connected at the top of risers equally spaced along the sprinkler
laterals. The function of the sprinkler lateral is to convey the water under the desire pressure
head and divide it efficiently among the risers. Sprinklers of this group distribute the water

59
over a circular rectangular area under light wind conditions. Water distribution may be greatly
distorted under heavy wind conditions. Each sprinkler has one or more nozzle through which
the water is ejected. The water jet coming out of the nozzle breaks up into drops, which
spread over the sprinkled area.
The water distribution pattern of most sprinklers of group A (within the optimal range
of pressure heads for each sprinkler and under light or no wind conditions) is characteristically
triangular. For these patterns the maximum depth of water falls near the sprinkler, while no
water falls at the perimeter of the wetted area. The depth of water decreases gradually from
maximum to minimum. Normally, distribution patterns of existing sprinkler do not achieved
the ideal triangular form, especially when sprinklers are operated under windy conditions.
Sprinkler operation under windy conditions normally necessitates adjustment in
overlapping and spacing. Usually, efficient sprinkling under windy conditions may be
achieved by spacing the sprinklers closing together. It is more economical to decreased
spacings between the sprinklers on the lateral. To this end, sprinkler laterals should be
designed to operate at right angles to the prevailing wind direction. Operating the system with
sprinklers of higher-pressure heads and larger diameter losses also helps. Sprinkler risers
should be at right angles to the soil surface, in order to maintain a uniform distribution
patterns. Also, tall risers, as in corn or orchards, should be provided with ample support.
An efficient sprinkler system is designed so that sufficient pressure heads are
available at all parts of the irrigated fields, no matter how remote or high they may be. In this
way the proper application rate are guaranteed and the required amounts of water may be
supplied within the desired number of hours.

Detailed Description of Group A Sprinklers:

Revolving Sprinklers ~ these sprinklers are operated by hammer which acts horizontally
round a vertical pivot and is regulated by sprinkler. Two types of hammer are commonly
used: regular hammer and the wedge hammer. Rotation of the sprinkler with the regular
hammer is effected when the jet, leaving the nozzle, strikes the driving heads of the
hammer. The striking jet produces a horizontal force component, perpendicular to the
driving heads. Thus the hammer is forced to rotate in this direction, as far as the spring
would allow it. Most revolving sprinkler today operate with the wedge type hammer.
Operation with the wedge hammer generally improves the distribution patterns of the
sprinklers, especially when pressure heads are low.

Revolving sprinklers manufactured with either one or two nozzles. With two nozzles,
one is the range nozzle-wetting mainly the outer rings of the wetted area-and the other is the
spray nozzle, which opposite the driving head, or the wedge. The range nozzle is normally
the larger. Some of the better-known two-nozzle sprinklers are Naan 323/90-91, 333 and
344, and Lego S33A, S33/22 and B7. The sprinklers operate at either low or high pressures,
and have a wide range of spacings and application rates. Some revolving sprinklers are
manufactured with only one nozzle. The one nozzle combines the effects of both range and
spray nozzles. Generally, these sprinklers operate at low-pressure heads, and are uses for
either under tree or low-application irrigation. Sprinklers with low application rates are used to
irrigate either heavy soils or areas with strong varying daily winds. Where heavy soils are
encountered, low application rates keep the soil from sealing up and prevent reduction of
infiltration rates. In area of strong daily winds these low application rates enable night
operation, thus reducing hazardous wind effects.
The part-circle sprinkler is another example of the one-nozzle sprinklers. This
sprinkler is used to irrigate the boundaries of irrigated fields. Revolving sprinklers may also
be distinguished according to their operating pressure heads. A sprinkler belongs to one of
three classes of pressure heads: low, medium, or high. Low-pressure heads sprinklers
normally operate between 10 and 25 m (15 to 37.5 lb/in. 2); (examples: Naan 323/90-91, Lego
S 33 A). Medium pressure –head sprinklers operate at 25 to 35 m, (37.5 to 50 lb/in. 2;
example: Naan 333, Lego B7). High pressure-head sprinklers operate above 35 m (50 lb/in. 2:
example: Naan 344). The above definitions are, of course, arbitrary. Generally, sprinkler
60
systems are design for low or medium pressure-head revolving sprinklers. Occasionally,
when the pressure at the water source is high, or when a pump id required in any case,
systems are designed for high pressure revolving sprinklers, or even for sprinkler guns.
These types of system are usually the cheapest and most efficient. The advantage of
revolving sprinklers id that they can be adapted to all soils and crops.
The range of pressure heads of a revolving sprinkler and the choice of nozzles
sometime enable the designer to make simple and economical changes, if needed, in existing
systems. One example is that reduction of pressure supplied by the water source, after the
system has been operating for some time. The reduction in pressure head brings about the
reduction of sprinkler discharges and application rates. This in turn forces the irrigation to
increase the number of hours needed for supplying the required amounts of water. Such a
solution is not always possible. The most efficient approach is to replace the nozzle with
larger ones. This approach is valid if the larger nozzles at the lower pressure head the
original rate or water, with no change required in the existing spacings.
Another example is the reduction in infiltration rate of the soil. This may result in
ponds and erosion. To prevent possible damage, lower application rates-associated with a
greater number of hours-should be used. In this case the solution involves replacing the
original nozzles with smaller ones, thus reducing the application rate of the sprinklers. The
new should be able to sprinkle efficiently without necessitating any changes in the existing
spacings of the system.

Whirling Sprinklers
These sprinklers are simply and sturdily constructed. They have either two or three
long arms at the ends of which are the nozzles. A reaction force, resulting from the ejecting
jet effects rotation. A whirling sprinkler wets a circular area when operating without wind. It
whirls at about 60 or more r/min. it normally operates at a limited large of pressure heads, the
optical pressure head being, generally, 10 m (15 lb/in. 2). It discharges generally up to 1 m 3/hr
or 4.4 gal/min. The main disadvantages of these sprinklers are (1) the range of pressure
heads is very narrow, thus limiting the choice of spacings, (2) the high application rates limit
the use to light soils, and crops the completely cover the ground, and (3) the close spacings
necessitate a larger number of pipes and sprinklers and investment for large areas may run
high.

Fixed-Head (Garden) Sprinklers


These sprinklers operate without any moving parts. They are simple and cheap.
They sprinkler a fine spray and should preferably operate without wind. They are mainly used
for irrigating lawns and gardens with permanent systems. Fixed-head sprinklers are
manufactured to wet full circles, half-circles, or strips. They operate at low-pressure heads
and close spacings. Generally, these sprinklers operate at pressure heads of 10 m (15 lb/in. 2)
and discharge up to 0.75 m 3/hr (3.3 gal/min), with spacings up to 5 m in each direction.
Fixed-head sprinklers have high application rates that vary normally between 20 and 50
mm/hr (0.8 to 2.0 in./hr). These application rates are too high for most soil and crops, and are
mainly adapted to lawns.

Sprinkler Guns
These sprinklers operate at high-pressure heads, varying in general between 40 and
75 m (60 to 110 lb/in.2). They discharge from about 20 to 100 m 3/hr (90 to 440 gal/min).
Spacing varies generally, according to the type of guns and pressure head, between 30 x 30
m and 70 x 70 m. Because of these wide spacings, the area irrigated by one time and labor
required for moving aluminum pipes. They normally have from one to three nozzles.
Sprinklers guns may be used to irrigate pastures, forage crops, orchards, cocoa, etc. Moving
the gun from one position to another is not easy, and frequently more than one man is
needed for the job.

FURROW IRRIGATION

61
Irrigation by the furrow method is accomplished by running water through small
channels (furrows) while it moves down or across the slope of the field (Fig. 8). The water
sips into the bottom and sides of the furrows to provide the desired wetting. Careful land
grading for uniform slopes is essential with this method.

Figure 8. Furrow-irrigated tobacco plants

Furrow irrigation is generally used under the following conditions:


1. For row crops (vegetables, sugarcane, corn, cotton, sorghum, and other field
grains) tree crops, vineyards and especially tall crops. Furrows are particularly
suitable for irrigating crops which are subject to injury if water covers crown or
stems of the plants (Figures 9 and 10).
2. On regular topography and level to moderate slopes.
3. Where water supply has low pressure and small discharge rate.
4. On medium and heavy soils.

This system of irrigation is also used extensively by farmers to irrigate plants planted
on beds, ridges, or nearly level land. The beds or ridges are made by either plowing the land
to form beds or making deep furrows in the tilled land. In either case, furrows elevate the
ridges. Water supplied to the crop from the furrow with the use of a gated pipe (Fig. 11) or cut
head dikes (Fig. 12).

62
Figure 9. Furrow-irrigated corn before flowering

Figure 10. Furrow irrigated cotton plants

Figure 11. Furrow-irrigated corn on ridges with gated pipe

Figure 12. Furrow-irrigation with head dikes cut to let water flow into furrows

63
Furrows are most commonly run directly down the slope, but sometimes they can run
almost parallel to contours to control erosion. They may also run across the slope to keep the
farm fields rectangular and the row length somewhat uniform. When this is done, care must
be taken to prevent water from over-topping the furrows and breaking them.

Spacing of the furrows is ordinarily determined by the spacing of the row crop,
although soil characteristics must be considered. The lateral movement of water from furrows
with uniform profiles depends primarily on the texture of the soil, the wetting pattern, being
broader in clays than in sands. To obtain complete wetting of sandy soils to depths of 1.2 m to
1.8 m, the furrows should not be placed more than 50 cm apart. In uniform clay soils,
complete wetting to the same depths might be obtained by furrow spacings of 120 cm or
more.

The length of the furrows must be selected with care. Percolation losses increase as
the furrows become longer. They should, therefore, be as long as reasonable efficiency in
applying water will permit. Recommended lengths of furrows for different soils are given in
Table 4.

Advantages:
1. Low investment in equipment
2. Low pumping costs
3. Low operating costs
4. Full use of irregular shape fields
5. Complete irrigation of margins of fields without wetting the adjoining fields
6. Suitability for crops which are sensitive to flooding irrigation because of leaf
diseases.
7. Suitability for crops which are sensitive to sprinkler irrigation because of leaf
diseases.
8. Does not interfere with crop spraying
9. Not affected by wind conditions; can be used at all hours of the day and night
10. Suitability for irrigation with sewage water (for tree crops and industrial crops)
11. Low evaporation losses

Disadvantages:
1. Can cause soil salinity problems in strip areas between the furrows
2. Irrigation efficiency lower than with sprinkler
3. Uniformity of distribution of water not high
4. Furrows must be renewed every year, and sometimes after each cultivation
5. May cause soil erosion if slopes are too great
6. Requires more land leveling than for sprinkler irrigation
7. Not suitable for light, sandy soils because of the length of the furrows is limited
8. Not suitable for the application of very small amounts of irrigation water
9. Not suited for automation
10. Not suitable for germinating seeds

Table 4. Recommended lengths of furrows for different soil types, furrow slopes, and
depths of irrigation
                   
FURROW LENGTH (m)
CLAYS LOAMS SANDS
Furrow Slope (%)
Net Depth of Water Application (cm)
7.5 15   5 10   5 7.5 10

64
0.05 300 400 120 270 60 90 150
0.10 350 440 180 330 90 120 190
0.20 370 470 220 370 120 190 250
0.30 390 500 280 400 150 225 280
0.50 380 500 280 370 120 190 250
1.00 270 400 250 300 90 150 220
                   

Unnecessary water loses result from deep percolation if the furrows are too long. The
initial stream should be large enough to run through the furrow rapidly without erosion, but
should be reduced so that excessive runoff will not occur during the remainder of the irrigation
period.

Period flow of water in the furrows is of utmost importance for the efficient use of
irrigation water. The most uniform distribution is usually obtained by starting the irrigation with
the largest unit flow that can be safely carried in the furrows. With short, flat furrows, a large
flow is use to fill furrows quickly, then the flow is shut off, and the ponded water is allowed to
infiltrate the soil. With long sloping furrows, it is generally necessary to regulate the flow
during irrigation.

The maximum flow rate allowable at the start of the irrigation is determined by the
need to prevent overtopping of the beds and soil erosion. In Table 5 are recommended
infiltration rates for safe furrow inflows.

The use of plastic tubing to siphon water is applicable. The head requirement is small
to sufficiently drive the water into the furrow. For a head of about 30 cm using a 5-cm plastic
hose, the discharge is approximately 0.49 li/min. Shown in Table 6 are ranges of pipe flows
for specific heads.

(LAB EXERCISE ON FIELD TEST TO DETERMINE THE LENGTH OF FURROW P 340


CRC)

Table 5. Soil Infiltration rates and suitable furrow inflows per


100 m of furrow length at a furrow spacing of 1 m.

     
SOIL INFILTRATION RATE FURROW INFLOW
  (mm/hr) (li/sec)

Clay 1-5 0.03 - 0.15


Clay loam 5 -10 0.15 - 0.30
Silt loam 10 - 20 0.30 - 0.50
Sandy loam 10 - 30 0.50 - 0.80
Sand 30 - 100 0.80 - 2.70
     

Table 6. Siphon pipe flows (li/sec) for various heads


         

65
DIAMETER OF SIPHON
HEAD (cm)
(mm)
  5 10 15 20

10 0.05 0.07 0.08 0.09


20 0.19 0.26 0.32 0.73
30 0.42 0.59 0.73 0.84
40 0.75 1.06 1.29 1.49
50 1.17 1.65 2.02 2.33
         

CORRUGATION IRRIGATION
A variation of the furrow method is the use of small rills or corrugations for irrigating
closely spaced crops, such as small grains and pastures. The water seeps laterally through
the soil, wetting the area between corrugations.

This method is used on fine-textured soils that take in water slowly and on lands that
are moderately steep and irregular. Corrugations are often used on soils which seal the crust
when flooded. The spacing and size of corrugations vary with the soil (Tables 7 and 8). In
general, the more porous the soil, the more closely spaced the corrugations should be permit
wetting between them without excessive percolation losses. Their lengths depends upon the
soil and the slope. They should be short enough to avoid over-irrigation of the upper end of
the field by the time the lower end receives sufficient water. Corrugations are often used to
establish stands of perennial crops on border strips, but after the first season they generally
serve no useful purpose.

Table 7. Length and spacing of corrugations for different soils (Sterns, 1979).
                 
  CLAYS   LOAMS   SANDS
SLOPE
Length Spacing Length Spacing Length Spacing
(%)    
  (m) (m) (m) (m) (m) (m)

Deep Soils 0.2 180 0.75 130 0.75 70 0.60


0.4 120 0.65 90 0.75 45 0.65
0.6 90 0.55 75 0.65 40 0.50
0.8 85 0.55 60 0.55 30 0.45
1.0 75 0.50 50 0.50

Shallow Soils 0.2 120 0.60 90 0.60 45 0.45


0.4 85 0.55 60 0.55 30 0.45

66
0.6 70 0.55 50 0.50
0.8 60 0.50 45 0.45
1.0 55 0.45 40 0.45
                 

Table 8. Soil infiltration rates and suitable corrugation inflows per 100 m
of corrugation length with corrugating spacing of 0.65 m.

     
Infiltration Rate Corrugation Inflow
SOIL
(mm/hr) (li/sec)

Clay 1-5 0.02 - 0.09


Clay loam 5 - 10 0.09 - 0.20
Silt loam 10 - 20 0.20 - 0.40
Sandy loam 20 - 30 0.40 - 0.50
Sand 30 - 100 0.50 - 1.90
     

FLOODING
Can be accomplished in areas which have uniform and gentle slopes. It is generally
practiced in areas where irrigation water is abundant and inexpensive. This could be done
either by (1) ordinary flooding, (2) border strip flooding, (3) level-border and basin irrigation,
and (4) contour ditch irrigation.

Ordinary Flooding. Water is applied from ditches to guide its flow. It is difficult to
attain high irrigation efficiency in ordinary flooding. Much of its success depends primarily on
the smoothness of the land surface, the proper size of irrigation stream, and the attention and
skill of the irrigator. The water is brought to the field in permanent supply ditches which are
spaced apart, depending on the grade of the land, the texture and depth of the soil, stream
size, and nature of the crops.

The chief advantage of the ordinary flooding over the methods is its low initial cost of
preparing the land. However, it is more advisable to prepare the land for controlled flooding
when land, water, or labor is expensive; when soil is deep and not likely to crust badly; and
when the land is not too rough or steep.

Border-Strip Flooding. A field is divided into a series of strips, 5 to 15 m wide and


75 to 300 m long, by borders or ridges running down the predominant slope or on the contour.
To irrigate, water is released into the head of the border; it advances, confined and guided, by
two adjacent borders in a thin sheet toward the lower end of the strip. The object is to allow a
sheet of water to advance down the narrow strip of land, allowing it to enter the soil as the
water advances (Fig 13). Although the border method is suitable for a wide range of soil
texture, it is generally recommended for fine-textured soils with low intake rates. Table 9 gives
recommended dimensions for border strips of different soil textures. Border irrigation is
generally used for hay, and grain crops on land having a slope of up to 3 percent.

67
Figure 13. A schematic sketch illustrating the layout of an impounding border-irrigation system

In practice, a large stream is turned into a border strip. This stream is then turned into
subsequent strips when the advancing stream approaches the lower end of the field. The
stream should be large enough to form an advancing water front throughout the width of the
border, regardless of whether this is for a flat or for a steep-grade run. A small stream would
collect and concentrate in depressions along the run, and furrow channels which may develop
into a gully without properly watering the width of the border. Only a large stream, that will
completely cover the area between borders and flow in an advancing sheet, will cause no
erosion damage. On steep slopes, the water sheet may pass too rapidly and, thus, will not
provide adequate time for the water to enter the soil. A second application has to be done
over the border to fill in deficiency.

Table 9. Suitable dimensions for border strips.


           
SOIL
SLOPE LENGTH FLOW
TEXTURE INFILTRATION WIDTH (m)
(%) (m) (li/sec)
RATE (mm/hr)

Sand 25 and over 0.2 15 - 30 60 - 90 220 - 450


0.4 10 -12 60 - 90 100 - 120
0.8 5 - 10 75 30 - 70

Loams 7 - 25 0.2 15 - 30 250 - 300 70 - 140


0.4 10 - 12 90 - 180 40 - 50
0.8 5 - 10 90 12 - 25

Clays 2.5 - 7 0.2 15 - 30 350 - 800 45 - 90


0.4 10 - 12 180 - 300 30 - 40
           

68
To construct a border system, it is necessary to mark the location of the border
ridges. These ridges can be constructed manually with animal-drawn implements, or with
mechanical equipment like the border disc. They can be constructed also by earth-moving
machinery while leveling the land. The ridges should be constructed sufficiently high and firm,
so that water will not easily scour or overtop them. If border ridges are rounded, they can be
planted; and only little land maybe rendered unproductive. On steep slopes, the border ridges
should be close together. On more gentle slopes, they can be wider apart; and large streams
of water may be used. The strip of land between ridges should be leveled, so that water will
evenly spread over the surface. It is also desirable to have a uniform slope down the strip for
an even application of water.

Because of its wide adaptability and ease of water application, careful though should
be given to this method of irrigation. Physical properties of the soil, the shape of the land, and
anticipated land use are important considerations in preparation for the border irrigation.
These factors along with the size of the irrigation stream available, determine the suitability of
this method and the proper width and length of the border strips. Three important
requirements of this method are:
(1) relatively large irrigation streams,
(2) gentle topography, and
(3) careful leveling.

Level-border and Basin Irrigation. The basin method of irrigation consists of


supplying water to level plots surrounded by dikes or levees. It is especially adapted to nearly
level lands, and may be used on a wide range of soil textures and crops. This method is
particularly useful on fine-textured soil with low permeability, where it is necessary to hold the
water on the surface to secure adequate penetration. It is often used for leaching salts by
deep percolation in the reclamation of saline soils. Pasture, hay grain, orchards and many row
crops can be irrigated this way, but because of the cost involved in leveling the plots, its use
is restricted only to level lands. Small basins used in the irrigation of orchards are commonly
referred to as checks.

Level borders or basins may be square, rectangular, or irregular in shape, and may
vary in size from 2 m2 or more. Tables 10 and 11 give the recommended design for the level
border of basins. During irrigation, the basins are rapidly filled with water to a depth sufficient
to refill the soil reservoirs. In areas with heavy rainfall, it may be necessary to provide some
means of surface drainage.

Table 10. Suitabale areas for basins with various soil textures.

         
AREA (M )2
FLOW
(li/sec) Sandy
Sand Clay Loam Clay
Loam

10 65 200 400 700


20 130 400 800 1,400
50 325 1,000 2,000 3,500
100 650 2,000 4,000 7,000
         

69
Table 11. Suitable spacing of terrace steps for basins.

   
LAND SLOPE
SPACING (m)
(%)

0.1 60 - 150
0.2 30 - 75
0.5 12 - 30
1.0 6 - 15
1.5 4 - 10
2.0 3 - 7.5
3.0 2-5
4.0 1.5 - 3.75

Contour-ditch Irrigation. Controlled flooding from field ditches along the contour of
the land allows the water to flood down on the slope between field ditches without employing
dikes or other means that guide or restrict its movement. Because water naturally seeks the
lowest places leaving the higher areas dry, the field ditches must be placed close enough to
keep the water uniformly spread. Frequent ditch checks, outlet openings, and spreader
furrows or siphons in the ditch are needed for uniform distribution of water over the field.
Some irrigators prefer checking the water so that it will flow over the sodded ditch bank, while
others provide turnout boxes, siphons, or spills at high points.

The contour-ditch method is often used for close growing crops as pasture on sloping
and rolling land not easily utilized for better method of irrigation. Also, steep slopes may be
irrigated with contour furrows for such crops as citrus, deciduous tress, and vegetables. Much
time is required in applying water efficiently with this method of irrigation.

SUB-IRRIGATION
This requires complete control of the water table so that the root zone is kept
relatively free of excess water but is continually supplied with capillary moisture during the
cropping season. Lands suitable for this method of irrigation are rather limited and usually
consists of peat. These soils must permit rapid lateral and downward movement of water yet
is capable of moving the moisture from the water table throughout the major position of the
root zone. The topography of the land should be smooth, uniform, and approximately parallel
to the water table. Many crops, except in orchards, are adapted to this method of irrigation.

DRIP IRRIGATION
The application of water to the soil through small orifices is known as drip or trickle
irrigation. The small orifices, often called emitters, are designed to discharge water at rates of
1 to 8 li/hr. Water is delivered to the orifices through plastic pipelines which are generally laid
on the soil surface or buried. The rate of the discharge is determined by the size of the orifice
and the pressure in the pipelines. The pressure may vary from 0.15 atm. To 2 atm.

Drip irrigation is particularly beneficial for young orchards, vineyards, closed spaced
perennials, and other crops of high value and in areas where water is scarce or has high salt
content (Fig. 14). A highly efficient water utilization can be achieved with this method, but it is
very expensive.
70
Figure 14. Drip irrigation

Drip irrigation systems (or in short, drip systems) are classified according to:
 Emitter type
 System location (on or under the surface)
 Operation pressure (“low”, about 5 m or less: or “high”, round 1 atm [10 m] or more)
 Portability ( non portable, semi portable, or portable)

Orifice – or long path emitters – a solid system placed on the ground surface, operating
at a pressure of about 1 atm.

Sub surface irrigation – drip irrigation was called originally.

Sub irrigation – a method, which really consists of the manipulation of the ground water
level by means of a ditch or buried piped system ( “reverse drainage”), and has nothing in
common with drip irrigation.

Dr. Pietro Celestra of Pisa, who has been experimenting with the method developing it
and publicizing it since 1951. (an Italian pioneer of drip irrigation)

Cameron Irrigation. Ltd. (England) – introduced a screw – threaded metal nozzle


which had the water passing a helical path around the “screw” and supplied about 1 ph at
1.3 m pressure; in 1967 it was replace by plastic nozzle with twice the output (2.3 ph).

1956 Volmatic system – invented by Denmark; this system had the water passing
through a long (85 cm) 0.8 mm diameter capillary tube.

1962 S. Blass in Israel – developed the “high” pressure long path emitter, where 3 m
long fine tubes, of 1.2 – 1.4 mm inner diameter, stuck into the wall of a distribution pipe,
dissipated the 1 atm. Tubes were wound around the distributing pipe.

Improved manufacturing methods resulted in:


a. precisely – size
b. uniform emitters
c. more dependable
d. cheaper
e. long lasting

Reports on the extent of drip irrigation in the world:


a. 290,000 da drip irrigation in the U.S
b. 250,000 da in Australia
c. 65,000 each in Israel and Mexico
71
Advantages and Disadvantages of Drip Irrigation

Advantages
Soil moisture – Drip irrigation keeps soil moisture at a constant, optimal level by
renewing the water supply to the root zone at the same rate it is used up.

Water saving – water is applied most efficiently, exactly as much as required when
required and where required, so there are minimum losses due to deep percolation, wetting of
areas not under crop, or evaporation from land surface, from foliage, or in air.

Hydraulics – water control is easy and more complete. Since irrigations are slow and
spread over a long time, peak discharges are reduced, thus requiring smaller pipes, pumps,
and fitting, and causing less wear and longer life of network elements
.
Irrigation operation – work is comparatively easy and requires less manpower.

Agricultural operation – crops often show higher yields, better quality and earlier
maturity than under “conventional irrigation”.

Fertilizing – application of fertilizers can –and- should – be done in a dissolved form


through the irrigation water, thus making possible a constant nutrient supply, regulated in the
rate and in composition according to the plant’ s age and requirements, with reduced
fluctuations in availability and less losses.

Economics – Drip irrigation equipment is not cheap, but compares well with the cost
of solid systems of other methods.

Disadvantages, Problems and Some Solutions


Agricultural – the location water application causes the development of dense, but
limited in volume, root mass.

Technical – the main problem in Drip irrigation is clogging of the emitters by


suspended materials (sand and salt), precipitated salts (mainly carbohydrates), rust and
other iron oxides, and organic materials including plant roots, live organisms – such as algae
and minute animals – and inert matter.

Salinity – an important problem in Drip irrigation is the accumulation of salts in the


interface between the irrigated and nonirrigated zones in the soil, whenever there is any
appreciable salinity of the soil and/ or the irrigation water.

Other problems – it include the cost of the systems which is high in comparison with
surface or portable sprinkling irrigation – but generally nit high enough to be prohibitive.

Drip Irrigation System Elements

Main pipe – a rigid pipeline, generally steel, asbestos cement, concrete, or other
similar materials almost always buried underground conveying the water from the source
(such as well, lake, regional pipeline or canal with a pumping plant).

Control head – the control irrigation and it’s operation point of the system, consisting
of valves, discharge and pressure meters (means for control and regulation of discharges and
pressures, including no-return valves and air vents), automation equipment and control,
fibers, and dissolved fertilizer applications
.

72
Submain (or secondary) pipe – a many valved pipe, distributing the water to the
various subunits within the unit.

Auxiliary (manifold) pipe – a flexible or rigid pipe, generally of 20 to 75 mm


diameter, distributing the water between the laterals that belong to a simple subunit. The
manifold and it’s lateral are designed and operated as a single unified system, the Unit Drip
System (UDS), which is controlled by a single valve.

Lateral (pipe) – as a rule, is a flexible polyethylene or PVC pipe, laid on top of the
ground, carrying the emitters. Its diameter will generally be 12 to 25 mm, and its pressure
rating 4 atm (unless the system is portable, when structural strength may dictate a 6 atm
pipe.

Emitter (dripper) – a device for reducing the line of water pressure to atmospheric
pressure, providing water at a low, controlled discharge (2 to 4 l/hr).

The Emitters

The emitter, or dripper, is a device or system element which makes drip irrigation
possible by providing irrigation water at low flows and atmospheric pressure. Since the
performance of the emitters determines to a large extent the efficiency of the whole system,
and since their number is high (varying, typically, between 200 and 2000 hectare), they
should be:

1. Cheap
2. Uniform in structure and operation
3. Simple to manufacture, install and maintain
4. Minimum problems (such as clogging) at standard operating pressure (3 to
30 m)
5. Able to produce a constant low rate water supply as insensitive as possible to
line pressure variations.

Emitters may be classified according to various criteria. The most important are:

Emitter principle – Orifice, long path, perforated pipe, double-wall pipe.

Flow regime – laminar or turbulent (or partially turbulent); this is a strong related to
emitter principle.

Operation pressure – “Low” (2 to 5 m) or “High” (8 to 15 m).

Path cross- section – “narrow” (below 0.8 mm), “medium” or “wide” (above 1.5 mm);
related to sensitivity to clogging.

Discharges – “low” (below 4 l/hr), medium (4 to 10 l/hr) and “high” (15 l/hr and more).

Orifice emitters – an orifice producing a jet of water that strikes a cap which acts as
a pressure dissipator.

Long – path emitters – are principle long tubes allowing pressure drop by friction
along the tubes and low discharges.

Porous tubes – are pipes with walls having minute pores, through which water may
be drawn out by the soil suction.
73
Double-wall pipes – are two pipes, made of flexible polyethylene, one within the
other.

Drip Irrigation Equipment manufacturers

Emitters are manufactured by the following companies. Some of the more well-
known are mentioned below.

Country Company Type


U.S Drip EZE long path emitters
Submatic, Water saver Orifices emitters
Subterrain, Spears Self-regulating
England Cameron Orifices, long path emitters and micro tubes
Israel Netafim, Lego Side and in line screw
NAAN Nozzle with screw path plug.
Denmark Volmark Micro tubes orifices with perforated plastic.

Reynolds number, Re – which represents the radio of inertia forces to viscous


forces during flow.

There are 4 Regions of Flow regime according to Reynolds number:

a. Laminar flow – is inversely proportional to Re, and unaffected by e/D (i.e unaffected
by the the roughness).

R= o4
Re
b. Unstable flow – in this region discharge is unstable, and f cannot be calculated from
either Re or e/D. Use of flow in this region is undesirable.
c. Partially turbulent flow, and fully turbulent flow – f depends on both Re and e/ D.
Use is made of the Colebrook-White formula.

74
Relationship of pressure and discharge - Rewriting Equation 5 we can get, for a given
pipe (i.e., for given L and D).

DRAINAGE ENGINEERING
CHAPTER 7
INTRODUCTION

Drainage means the removal of excess water from a given place.

Two types of drainage can be identified:


i) Land Drainage: This is large scale drainage where the objective is to drain surplus
water from a large area by such means as excavating large open drains, erecting
dykes and levees and pumping. Such schemes are necessary in low lying areas and
are mainly Civil Engineering work.
Field Drainage - This is the drainage that concerns us in agriculture. It is the removal
of excess water from the root zone of crops.

THE NEED FOR DRAINAGE


Drainage Problem – excess water on the surface or subsurface (below the soil root
zone) of the soil.
Subsurface – characterize by high water table.

75
76
After Rainfall

77
PURPOSE OF DRAINAGE
 Main purpose – to provide a root environment that is suitable for the maximum growth
of plants
 Objective – increase production and for sustainable farm production.

Decrease in crop production due to poor drainage – limited amount of soil, inadequate
root system, and lack of nutrients.
To bring soil moisture down from saturation to field capacity. At field capacity, air is
available to the soil and most soils are mesophites ie. like to grow at moisture less
than saturation.
ii) Drainage helps improve hydraulic conductivity: Soil structure can collapse under
very wet conditions and so also engineering structures.
iii) In some areas with salt disposition, especially in arid regions, drainage is used to
leach excess salt.
iv) In irrigated areas, drainage is needed due to poor application efficiency which
means that a lot of water is applied.
v) Drainage can shorten the number of occasions when cultivation is held up waiting
for soil to dry out.

EFFECTS OF POOR DRAINAGE ON SOIL AND PLANTS


SOIL
 Displaces and obstructs the movement of air and gases.
- limited dissolved oxygen. DO is virtually absent in submerged soils,
apparently because of microbiological activity.
- Nonexistence of O2 causes the anaerobic decomposition of organic matter,
resulting in the production of reduced organic compounds such as methane,
methyl compounds, and complex aldehydes. Mineral substances in the soil
are altered from the oxidized state to reduced state.
- extremely slow rate of diffusion of gases.
- gas exchange is confined to a fraction of the top inch of soil.
- Waterlogging generally leads to a deceleration in the rate of decomposition of
organic matter holding the release of nitrogen.
PLANTS
 Decline in the transpiration rate of plants related to the deficient O 2 and increase
concentrations of CO2.
 Inhibited uptake of nutrients by plants. Symptoms: yellowing, reddening, or a
scorched or stippled (spotted) appearance of the leaves.
 Methane – found to inhibit the growth of tomato.
 Fluctuating water tables cause plants to rot off.

OTHER BENEFITS FROM DRAINAGE


 Reduced crop damage due to scalding.
 More healthful and pleasant environment for human habitation. Prevents breeding of
mosquitoes.
 Prevents soil compaction by animals and machines.
 Proper germination of seeds. A drained soil is a warm soil.
 Prevents onset of diseases.
 Prevents soil salinization.

Planning Drainage System


78
A. Section Advance organizer
The section will discuss steps to follow in planning a drainage system.

B. Section Intended Learning Outcome


The student must be able to:
1. Gather and evaluate the information necessary for planning a drainage system;
2. Anticipate drainage problems and consider these when a design is needed also for
an irrigation system; and
3. Comply with laws or regulations related to discharge of waste water.

C. Course Content
Drainage systems should be always complement irrigation systems unless the plan
covers drainage only such as for upland crop cultivation, orchard protection, reclamation or
urban drainage. Feasibility of a drainage systems greatly depend on economic
considerations. Benefits must exceed costs. Information that determine feasibility helps avoid
unnecessary costs.
The most useful criteria for appraising existing and potential drainage projects are:
Soil characteristics
Subsoil and substrata hydraulic conductivity
Topography
Depth and fluctuation of groundwater
Water quality
Soil salinity
Location of land with respect to major outlets

Drainage requires an adequate outlet. No amount of planning can solve the problem.
If the outlet easily clogs with sediments, it should be protected by proper erosion control
measures.

Planning Surveys
A reconnaissance survey is the first step to determine the general feasibility of
draining the land, the kind of system needed and rough estimate of costs. When the drainage
project proceeds, it is necessary to do additional surveys to locate right-of-ways and to carry
out construction works.

Surveys may not be necessary when drainage systems are only being rehabilitated.
Perhaps drains need only be deepened, enlarged or simply cleaned up. Thus, surveys may
be limited to profiles and cross-sections of existing ditches or pipelines.

Principal Information for Drainage Survey


1. Map of the area showing watershed boundary, area, slope, vegetation cover of area
tributary to each proposed ditch. The map should show ditch location, roads, and
other physical features.
2. Profile on or near the proposed or pipeline centerline with elevation and information
on existing channels, ditches, outlets, crossing and other structures.
3. Outlet conditions including adequate capacity, low and high water elevations,
frequency of flooding and authority to use of outlet, if necessary.
4. Land capability which should include agricultural land values, possible yield levels of
adaptable crops and clearing or development costs anticipated;
5. Soil investigations along proposed drain and levees and at structure sites to
determine characteristics of excavation works and construction materials with soil
borings to sufficient depth.;
6. Type and estimated costs of farm drainage required;
7. Condition which affect costs of construction such as right of way, accessibility, kind of
equipment and availability of contractor;
8. Benchmarks should be established along the proposed lines at desired intervals, on
permanent structures, properly referenced to facilitate relocation of future reference.
79
Drainability investigations
Soil survey and inspection may be used to determine the drainability of a piece of
land. However, tests may be needed especially if the watertable stays shallow. An auger hole
or perimeameter test determines in-situ permeability below the water table. Water is bailed or
pumped out and the rate at which water returns toward static level provides basis for
calculating permeability or hydraulic conductivity. Observation wells consisting of small
diameter pipes may also be used to monitor the fluctuation of the water table.

Designing of Drainage Ditches or Pipelines


An experienced agricultural engineer or civil engineer should design drainage ditches
or pipelines. Before negotiating with an engineer all the information helps one estimate the
engineer’s fee. An engineer could use a computer to solve design problems.

The engineer should be consulted often as the design work progresses.

Construction Phase

The engineer usually prepares the plans, specifications, and contract documents.
They should be clear, complete and should provide the following information:
1. location and amount of each item of work;
2. working condition for clearing and excavation;
3. time limits for starting and completion of the work and, if specified the penalties;
4. provision of setting stakes, inspection and tolerances or allowable limits as required;
and
5. method of payments.

Simple Drainage Plan


Fig. 11.4 shows a farm drainage plan indicating contours and location of drainage
lines. Analysis of contours will show that a lateral ditch is located at the lowest portion below
elevation 140. Another sublateral ditch, as shown by the arrow, passes from elevation 150 to
elevation 140. The arrow always point toward a lower elevation.

Localized Drainage Problems


If a drainage problem is observed to be localized these steps may be followed:
1. Determine if an outlet exists and if the capacity is adequate to discharge the excess
water;
2. identify the sources of excess water; proper water management can solve excessive
application of water. If excess water comes from upland areas, perhaps an
interceptor drainage canal or pipeline may be necessary;
3. Undertake soil borings too determine subsurface conditions. An underground pipe or
tile drain may be necessary to minimize surface disturbances that may create
passageway problems;
4. Undertake topographic survey or leveling work to be able to plan drainage systems
required by the situation;
5. Determine if excess water can be drained by gravity canal or by pumping scheme;
and
6. Prepare rough estimates of benefits and probable costs as basis for deciding whether
the proposed drainage improvement will be economically feasible.

80
Compliance with Drainage Laws and Regulations
In the Philippines, there are very few laws or regulations that regulate directly or
indirectly the discharge of excess water. Most of the regulations are mainly on discharge of
polluted water into rivers. In any case, it is important that prior to undertaking a drainage
project the local environmental office must be consulted to determine if there are laws and
regulations that should be observed.

Learning experiences
1. Instructional Foci
Layout of a drainage system on a topographic map.

2. General Teaching Strategy


During the lecture, the drainage system should be identified in the layout of the
irrigation system, and discussed. The important concepts from Fig. 11.4 should be
explained.
3. Suggested time to spend
4. Self-assessment questions:
a. if drainage is provided what would happen to land values in the area. Justify
your answer?
b. Under what conditions would you plan for a surface drainage system?

Unit Summary
Several factors to consider in planning for irrigation and drainage systems discussed.
Two most crucial factors – costs and presence of outlet, in the case of drainage system – are
emphasized. Furthermore, topographic mapping must be thought up well before it is carried
out. It may be involve unnecessary costs.
After the drainage plan come the design and construction, phases which should be
left to experienced engineers to see trough, especially for large areas. For small, manageable
areas it may not be necessary to engage the services of an engineer or contractor. Practical
knowledge and common sense of the farmer may be sufficient .
81
Unit Assessment Questions
1. In planning irrigation systems is it always necessary to do land leveling or land
forming? Why?
2. If a farm is located at an elevation higher than the source of water supply, give two
options for getting irrigation water.
3. With a topographic map to be provided by the teacher prepare a layout for a drainage
system which would be most efficient under specified soil and cropping practices.
4. A farmer in your area desires to drain 100 hectares of alluvial sandy loam land. If the
topographic survey shows that the maximum slope is 0.07 percent, what would be the
size of the drainage canal including the soil bank?

CHAPTER 8
HYDRAULICS AND THE MEASUREMENT OF WATER

BRIEF REVIEW ON HYDRAULICS

THE FLOW OF WATER IN OPEN CHANNELS

 Water flow: surface slope, gravity (source of energy)

82
DEVELOPMENT OF UNIFORM FLOW AND ITS FORMULAS

Qualifications for Uniform Flow


1. The depth, water area, velocity and discharge at every section of the channel
reach are constant; and
2. The energy line, water surface, and the channel bottom are parallel; that is their
slopes are equal.
3. Uniform flow cannot occur at very high velocity (ultrarapid)
 Results obtained from uniform-flow condition are approximates and general, but they
offer a relatively simple and satisfactory solution to many practical problems.

83
Establishment of Uniform Flow
When flow occurs in an open channel, resistance is encountered by the water as it
flows downstream. This resistance is generally counteracted by the components of gravity
forces acting on the body of the water in the direction of motion. A uniform flow will be
developed if the resistance is balanced by the gravity forces.

Figure:

Expressing the Velocity of a Uniform Flow


General form of most practical uniform-flow formulas:

x y
V =CR S

Where:
V = mean velocity
R = hydraulic radius
S = energy slope
C = factor of flow resistance, varying with the mean velocity,
hydraulic radius, channel roughness, viscosity, and many
other factors.

The Chezy Formula


Probably the first uniform-flow formula developed by French engineer Antoine Chezy
(1769).

V =C √ RS
C = Chezy’s C
Figure:

84
Assumptions:
1. The force resisting the flow per unit area of the stream bed is proportional to the
square of the velocity. The surface of contact of the flow with the stream bed is equal
to the product of the wetted perimeter and the length of the channel reach. The total
force resisting the flow = KV2PL.
2. In uniform flow, the effective component of gravity force causing the flow must be
equal to the total force of resistance.

Derivation:

KV2PL = ωALS

Where:
K = constant of proportionality
V = velocity
P = wetted perimeter
L = channel reach
ω = unit weight of water
A = water area
θ = slope angle
S = channel slope

Let A/P = R and C = (ω/K)1/2

Then: V =C √ RS
Determination of Chezy’s Resistance Factor

A. The G.K. Formula.

0 . 00281 1 .811
41 .65+ +
S n
C=
(
1+ 41 . 65+
0 .00281 n
S √R )
Developed by two Swiss engineers, Ganguillet and Kutter (1869). It was derived
elaborately from flow-measurement data in channels of various types.

B. The Bazin Formula

157 . 6
C=
m
1+
√R
m = coefficient of roughness

The Bazin formula was developed primarily from data collected from small
experimental channels; hence, its general application is found to be less satisfactory from the
G.K. formula.

C. The Powell Formula (Powell, 1950)

85
C=−42log ( 4CR + Rε )
Where:
R = Reynolds number
ε = Measure of the channel roughness

For rough channels, the flow is generally so turbulent that R becomes very large

compared to C; thus:

C=42log ( Rε )
For smooth channels, the surface roughness may be slight that ε becomes negligible
compared with R; then:

C=42log ( 4CR )
Since Chezy’s C is expressed implicitly in the Powell formula, the solution of the
formula of C requires a trial and error procedure. The Powell formula was developed from
limited laboratory experiments on smooth and rough channels and from the theoretical
velocity distribution. The application of this formula is limited, since further investigation is
needed for determination of the proper values of ε.

The Reynolds Number (R)

VL
V=
ν
Where:
= velocity of flow
V
L = characteristic length = hydraulic radius
 = Kinematic viscosity = 1.08 x 10-5 ft2/s

Example 2.1
Compute the velocity and discharge in the trapezoidal channel shown below using
the Chezy’s formula. Kutter’s n = 0.015, and S = 0.005 (units in meters).

Solution:
Solve A and R
Solve C then V and Q

86
The Manning’s Formula

1 . 49 2/3 1/2
V= R S (English unit )
n

1
V = R 2/3 S1/2 (Metric unit )
n
This formula was developed from seven different formulas, based on Bazin’s
experimental data, and further verified by 170 observations. Owing to its simplicity of form and
to the satisfactory results it lends to practical applications, the Manning’s formula has become
the most widely used of all uniform-flow formulas for open-channel flow computations.

For practical purposes, the Manning’s n and Kutter’s n values may be considered
identical when the slope is equal to or greater than 0.0001 and the hydraulic radius is
between 1.0 and 30ft.

Determination of Manning’s Roughness Coefficient

General approaches:
1. to understand the factors that affect the value of n and thus to acquire a basic
knowledge of the problem and narrow the wide range of guesswork,
2. to consult a table of typical n values for channels of various types,
3. to examine and become acquainted with the appearance of some typical channels
whose roughness coefficients are known, and
4. to determine the value of n by an analytical procedure based on the theoretical
velocity distribution in the channel cross section and on the data of either velocity or
roughness measurement.

Factors Affecting the Manning’s Roughness Coefficient

1. Surface Roughness
2. Vegetation
3. Channel Irregularity
4. Channel Alignment
5. Silting and Scouring
6. Obstruction
7. Size and Shape of Channel
8. Stage and Discharge
9. Seasonal Change
10. Suspended Material and Bed Load

Cowan developed a procedure for estimating the value of n:


n = (no+n1+n2+n3+n4)m5 Refer to Table 5-5 (Ven Te Chow, 1959, p.109)
Where:
no = is a basic n value for a straight, uniform, smooth channel in
the natural materials involved
n1 = is a value added for variations in shape and size of the
channel cross section
n2 = is a value for the variations in shape and size of the channel
cross section
n3 = is a value for obstructions
n4 = is a value for vegetation and flow conditions
87
m5 = is a correction factor for meandering channel
Refer to Table 5-6 for Manning’s roughness coefficient n values (Ven Te Chow, 1959,
p.110-113)

Refer to Fig 5-5 for typical channels showing different n values (Ven Te Chow, 1959,
p.116-123).

EXERCISE PROBLEMS

Exercise Problem 2.1. Compute the velocity and discharge of flow of new earth canal having
the same shape, size, slope and depth of flow as the channel given in Example 2.1.
Use (a) the G.K. formula, assuming Kutter’sn = 0.022; (b) Bazin formula, selecting a
proper value of m; and (c) the Powell formula, selecting a proper value of .

Exercise Problem 2.2. For the conditions given in Example 2.1, compute the values of
Bazin’s m and Powell’s .

Exercise Problem 2.3. Taking Manning’s n as the given value of Kutter’s n, solve Example
2.1 by the Manning formula.

Exercise Problem 2.4. If the coefficient of roughness n is unknown for the channel in
Example 2.1, but a discharge of 2,000 cfs is observed under the given conditions,
compute the values of Kutter’s n and Manning’s n.

Exercise Problem 2.5. Bazin’s tests was made on a rectangular plank flume 6.44 ft wide
with wooden strips 1 cm thick and 2.7 cm wide nailed crosswise on the bottom and sides
at a spacing of 3.7 cm center to center of strips. The flume gave a mean velocity of 3.33
fps at a flow depth of 1.02 ft and a slope of 0.0015. The temperature reading was 8.5 oC.
Determine the Manning’s n, and compute (a) Chezy’s C, (b) Kutter’s n, (c) Bazin’s m, and
(d) Powell’s .

Note: Change all English units to Metric units.

88
CHAPTER 9
SOILS

Definition:
- a material which supports foundations, roads, airports. Is considered to be all of
the material which covers the rock of the earth’s crust. (Engineering)
- Surface mantle of material limited to the first few feet to a depth of 5 to 10 ft (1.8 –
3m) subject to the forces of weather and climate. (Pedologist)
- Surface layers that support plant growth and deeper layers that transmits water.
(Drainage Engr.)
TERMS:
1. Soil Structure –
2. Cation exchange capacity – the total cations held on 100 grams of clay.
3. Flocculation – is the coalescing of individual clay particles to form larger particles or
flocs. (porosity is increased hence K is increased)
4. Dispersion (Deflocculation) – separation into individual particles.
5. Exchangeable Sodium Percentage (ESP) – the relative proportions of Na on a
particular clay particle to the cation exchange capacity. It is a measure of the amount
of sodium that is held on the cation exchange complex.
100(−0 . 0126+0 . 01475 SAR)
ESP=
1+(−0 . 0126+0 . 01475 SAR )

6. Sodium Absorption Ratio (SAR) – is a measure of the properties of the solution. It


is given by the formula:

89
Na
SAR =

√ ( C a+ M g)
2

COMPOSITION OF SOILS
a) Soil air (gases)
b) Soil Moisture (water)
c) Solid materials (inorganic or mineral materials and organic matter at various
proportions)

THE SOLID PHASE OF SOILS


- The solid phase of soils consists of the primary particles of sand, silt, and clay
along with OM and various chemical compounds.

A. Clay – clay particles are less than 5 microns in diameter. Clay particles possess a
negative charge

Two Groups of Clay


1. Kaolinite – the kaolin group of clays are characterized by having 1 silica and 1
alumina layer. The sheets are very compact. It does not shrink or swell with
water.
2. Montmorillonite – it has expanding lattices. The crystal lattice of the
montmorillonite expands and contracts as the amount of water that is present in
the lattice changes.

Measurement of the Electrolyte Content of Water


Electrical resistance (ohms)

Conductance (mhos); micromho = EC x 106


millimho = EC x 103

90
The Electrical Conductivity (EC, in mhos/cm) of the soil solution gives a rough
estimation of the salt content of the soil.

Conversion Factors:
a) EC to meq/li: meq/li = 10 x EC x 103 for irrigation water and soil
extracts in the range from 0.1 to 5.0 mmhos/cm.

b) EC to ppm: ppm = 0.64 x EC x 106 for irrigation waters in the


range 5,000 micromhos/cm

d) ppm to tons of salt per acre-ft of water : taf = 0.00136 x ppm

B. Silt – 0.005 – 0.05 mm. Silt is generally non-reactive. It does not swell when
exposed to water, nor does not have any appreciable amount of CEC. Silt particles
can easily slide over one another when wet which make it undesirable material for
construction purposes (i.e. open ditches).

C. Sand – sand particles may be as coarse as rock salt, as fine as powdered sugar,
or any size between those extremes ( ). The presence of sand in soil is noted by a
grittiness and roughness when a moist soil is worked between the finger. Akin to silt,
it shows some instability.

SOIL CLASSIFICATION

A. Engineers Classification

Unified soil-classification system – recognizes the fact that natural soils rarely exist
separately as sand, gravel or any single component, but are usually mixtures.

Engineering Properties of Soil


1. Particle size
2. Gradation
3. Particle shape
4. Density
5. Consistency

Particle Size
Coarse gravel – comparable in size to a lemon, an egg, or a walnut.

Fine gravel – about the size of a pea.


 The particles which pass a No. 200 sieve are called fines which are divided
into silt and clay.
 Mechanical analysis – the physical process of separating soil into its particle-
size groups. The engineer presents the results in a form of grain-size
distribution curve. This provides ready visualization of the distribution of
particle sizes or gradation.
 Wet mechanical analysis – used to determine the distribution of particles
sizes smaller than 0.05 mm which employs the principle of sedimentation.

Gradation – the distribution of particle sizes in a soil.


Hazen’s effective size (D10)– the grain size which corresponds to 10% on a
grain-size distribution curve. The effective size of clean sands and gravels can be
related to their permeability and is, therefore, of importance in drainage engineering.

91
Uniformity coefficient – ratio between the grain diameter that corresponds
to 60 % on the curve, and D10. The coefficient of gradation is given by the expression:

Well-graded soils – are those which have a reasonably large spread


between the largest and smallest particles, and have no marked deficiency in any one
size.

Pedologists’ Classification

Soil Profile – is a vertical section through the soil mass. It is classified into two
horizons:
1. Eluviation (A Horizon) – is the zone which is being leached. Chemicals and clays
are being dissolved and carried to greater depths in the soil.

2. Illuviation (B Horizon or subsoil) – refers to the zone in which these chemicals or


clays are being deposited and are accumulating.

 Below the A and B horizons lies the parent material or the raw
unweathered material from the soil has been formed.
 Residual soil or primary soil – if the parent material is rock.
 Secondary or transported soil – if the material has been transported by
wind or water.
 The “age” of soil is a measure of its development. The greater the
accumulation of clay or solid chemicals in the B Horizon the “older” is the
soil.
 Stratification – can be recognized by the abruptness of the change from
one stratum to another.
 Soil texture – the relative proportion of sand, silt and clay (with important
link to soil permeability).

The Soil Series


Soils which have common properties are grouped together into soil series. A name
assigned to each series and this name is usually a geographic place name of a locality
associated with the soil.

The Soil Type


The textural classification of soil is called soil type, and the complete identification of
the soil is made by specifying both series name as well as the type.

92
93
Size limits:
International Society of Soil Science United States Department of Agriculture
Separate Diameter Separate Diameter
Limits, mm Limits, mm
Very coarse sand 2.00 – 1.00
Coarse sand 2.00 – 0.20 Coarse sand 1.00 – 0.50
Medium sand 0.50 – 0.25
Fine sand 0.20 – 0.02 Fine sand 0.25 – 0.10
Very fine sand 0.10 – 0.05
Silt 0.02 – 0.002 Silt 0.05 – 0.002
Clay Below 0.002 Clay Below 0.002

CHAPTER 10
STATICS OF SOIL WATER

THE NATURE OF WATER

FORCES HOLDING WATER IN SOILS

Inbibitional Forces – forces which bind the water to the soil solids.
Interplanar water – causes the swelling and shrinkage of clay
94
Water hydration – shell of water molecules which surrounds some cations like sodium

Capillary Forces – causes the rise of water in tubes due to the difference in the attractive
forces of water to water (cohesive forces) and water to solid (adhesive forces).

h=
gρr
Derivation:

F=2 Πrσ cosθ


F=G=mg but m=ρV
2Πrσ cosθ=gρV
2Πrσ cosθ=gρΠr 2 h at θ=0; cosθ=1

h=
gρr

The Energy Concept of Soil Moisture


- In the region above the water table the soil-moisture pressure in negative. The
soil exerts an attraction of the soil water.
- Water Table – locus of points at atmospheric pressure. It represents the upper
limit of the zone of positive forces.

Negative and Positive Soil Moisture Pressures


Negative – water will be extracted from the soil. The more negative pressure applied,
the more water will be removed.

Positive – water will be forced out of the soil.

THE MOISTURE CHARACTERISTIC


Moisture characteristic curve – the plot of soil moisture content against the negative
pressure exerted. At 0 pressure (water table pressure), the MC = 47. 1% (“saturated” – up to
90% due to trapped air)

THE DRAINABLE PORE VOLUME


The drainable pore volume represents the volume of water that can be drained from a
unit volume of soil when the soil-moisture pressure is decreased from the atmospheric
pressure to some specific negative pressure.

95
Figure: Hypothetical moisture profiles showing two stages of falling water in the soil. The
shaded area represents the soil-water yield and is equal to the water which enters the drain.

Example

Soil Sample:
Assumptions: Soil bulk density = 1.40 (1cc of dry soil will weigh 1.4 g)

Initial Condition:
MC= 35% dry weight (saturated)
Soil moisture pressure = 0

Final condition:
Soil moisture capillary pressure is increased to 100 cm head
MC = 30%
Moisture concentrations:
IC: 1.40 x 0.35 = 0.49g of water @ density of water = 1g/cc
= 0.49 cc of water per cc of soil

FC: 1.40 x 0.30 = 0.42 g of water


= 0.42 cc of water per cc of soil.

Therefore: Drainable pore volume = 0.49 – 0.42 = 0.07 cc of water per cc of soil

In general: the quantity of water draining out of a unit column is given by:

96
h1 h1
a
q=∫ f (h )dh or q=∫ ah dh=
2 ( 1 2 22 )
h −h
h2 h2

Specific yield – the amount of water expressed on a volume basis that is released from the
soil when the water table drops a unit distance.

CHAPTER 11
PERMEABILITY

Darcy’s Law
The one-dimensional flow of water through a saturated homogenous soil can be
computed by the Darcy equation (Darcy’s Law):

More than 100 years ago, Henry Darcy studied the flow trough a bed of sand. He
reported in 1856 that the flow rate of water through a porous medium is proportional to the
head loss and inversely proportional to the length of the pathway of the flow. Water is made to
flow at a constant flow rate (Q) through a sand column with two manometers L apart (Figure
4). The fluid potentials or energy heads (h) at the manometer taps (points 1 and 2) may be
expressed in terms of the pressure head (P/) and elevation heads (z) (assuming constant
velocity or negligible change in velocity).
Darcy observed that:
(1) Q is proportional to the head loss, h; Q  h
(2) Q is proportional to the cross-sectional area of flow, A = r2; Q  A
(3) Q is inversely propotional to the length of the flow path, L; Q  1/L
(4) Q is influenced by the nature of the sand; Q  c, where c is flow medium
parameter.
Figure 3.0:

97
In equation form, these observations can be expressed as:
Δh
Q α cA
ΔL
Q Δh Δh
α c =k
A ΔL ΔL
or
Q Δh
v= =k
A ΔL

where: v = average velocity, “apparent” velocity


A = entire flow area
k = combined flow medium parameter and proportionality factor
(hydraulic conductivity)
h = change in head or elevation
L = change in flow distance along the flow path

The above equation is universally known as the Darcy’s law. It is valid for laminar
flows such as those of fluids through porous media. It is similar to many other basic
equations describing flow of masses and their properties(e.g. Ohm’s law, Dalton’s law and
heat transfer and gas diffusion equations). Darcy’s law serves as a basis for most equations
governing movement of water through soils.

Darcy’s law is in reality a model or mathematical expression for macroscopic flow of


water through a porous system. In reality, the flow through individual pores is non-uniformly
distributed along the flow direction. The head (h) varies continuously as pores change in
diameter and the flow through them converges or diverges. If the head is, however, averaged
at a number of points along the flow path, the change in head or hydraulic gradient and the
total discharge of the great number of pores are reasonably uniform. On a macroscale,
Darcy’s law has been shown to be mathematically valid. In general form, Darcy’s law may be
written as:

∂h
v =k
∂s (one dimension)

98
Example 1:

Given the following, compute the flow rate Q:

h1

Q=? h2

Datum line
100 m

x direction

h1 = 100 m
h2 = 50 m
K = 0.2 m/day
b = cx + d
Solution:

99
dh
Q=−KA
dx
dh
Q=−0 .2 ( cx +d ) consider unit Q
dx
x h

∫ Q( cx +d ) =−0 . 2∫ dh separate variables


0 hi

[ ]
x
Q h
ln ( cx+d ) =−0 . 2 [ h ]h 1
c 0
When x =0 ; h = h 1=100
x = 100 ; h = h 2 =50

[ ]
100
Q 50
ln ( cx+d ) =−0. 2 [ h ] 100
c o

c
ln (
Q 100 c+d
d )
=−0 .2 ( 50−100 )
10
Q=
1
c [(
ln
100 c +d
d )]
Drainage Flow Problems

1. Given the following flow and boundary conditions:

h1 = 100m

h 2 = 90m
K1 = (1 + 5x) m/day K2 = 10 m/day

X=0 x = 40 m x = 100m

Determine the following:


100
a) the flow rate through the porous media
b) the head at x = 40m
c) equations for the head as a function of x from x = 0m to x = 40m and from x = 40m to
x = 100m
d) Show that your equations in (c) satisfy the boundary conditions, i.e, the head at x =
0m, x = 40m and x = 100m and that the Q is the same for both sections. Also show
that the governing equation dQ/dx = 0 is satisfied in both sections.

2. Consider the following flow regions:

h1 = 100 m K = (100 + 1.0x)

b = (10 – 0.05x)m h 2 = 100 m

x=0 x(m)

a) determine the flow rate

b) Derive the equation for the head as a function of x from x = 0 to x = 100m

c) Show that your equation in (b) satisfies the boundary conditions (the head at x = 0
and x = 100m) and that you can reproduce the Q in (a) from Darcy’s Law. Also show
that the governing equation dQ/dx = 0, is satisfied.

101
CHAPTER 12
SEEPAGE ANALYSIS

Fundamental Flow Equation

All pressure analysis techniques are derived from solutions to the partial differential equations
that describe the flow of fluids through porous media under various boundary conditions. This
mathematical description of fluid flow is based on three physical principles: (1) the law of
conservation of mass, (2) Darcy's law, and (3) equations of state.

The law of conservation of mass and Darcy's law apply regardless of the fluid type and are
the basis of the flow equation and, therefore, of pressure analysis. The equations of state
apply to a specific fluid, e.g., oil, gas or water, and describe the relationship between that
fluid's density, pressure, and temperature.

In differential terms the law of conservation of mass can be expressed as

δ δt (ϕρ)=−1 r δ δt (rρV r ) (1)

Equation 1 is also termed the continuity equation.

Darcy's law states that the volumetric rate of flow per unit surface area at any point in a
porous medium is proportional to the potential gradient in the direction of flow at that point.

Mathematically, Darcy's law for radial flow is

V r =−k μ δP δr (2)

To obtain the general flow equation, the continuity and Darcy's equations (Equations 1 and 2)
are combined to yield:

δ δt (ϕρ)=1 r δ δr [rρk μ δp δr ] (3)

The general flow equation (Equation 3) describes the flow of a fluid in the reservoir at all times
and is independent of fluid type. This equation is called the radial diffusivity equation. The
derivation of this equation is based on the following assumptions:

 Darcy flow exists, i.e., no turbulence effects exist.


 Single-phase fluid flow exists.
 Gravitational effects are negligible.
 Reservoir is homogeneous and isotropic.
 Permeability and porosity are constant.
 Flow is isothermal (of constant temperature).

102
The Dupuit assumption

 The Dupuit assumption holds that groundwater moves horizontally in an unconfined


aquifer, and that the groundwater discharge is proportional to the saturated aquifer
thickness. It was first designed by Jules Dupuit in 1863 to simplify the groundwater
flow equation for analytical solutions.
 The Dupuit assumption requires that the water table is relatively flat, and that the
groundwater is hydrostatic (i.e., the equipotential lines are vertical):

 where is the vertical pressure gradient, is the specific weight, is the


density of water, is the standard gravity, and is the vertical hydraulic
gradient.

Flow Nets

A flownet is a graphical representation of two-dimensional steady-state groundwater flow


through aquifers. Construction of a flownet is often used for solving groundwater flow
problems where the geometry makes analytical solutions impractical. The method is often
used in civil engineering, hydrogeology or soil mechanics as a first check for problems of flow
under hydraulic structures like dams or sheet pile walls. As such, a grid obtained by drawing a
series of equipotential lines is called a flownet. The flownet is an important tool in analysing
two-dimensional irrotational flow problems.

Basic method

The method consists of filling the flow area with stream and equipotential lines, which are
everywhere perpendicular to each other, making a curvilinear grid. Typically there are two
surfaces (boundaries) which are at constant values of potential or hydraulic head (upstream
and downstream ends), and the other surfaces are no-flow boundaries (i.e., impermeable; for
example the bottom of the dam and the top of an impermeable bedrock layer), which define
the sides of the outermost streamtubes (see figure 1 for a stereotypical flownet example).

Mathematically, the process of constructing a flownet consists of contouring the two harmonic
or analytic functions of potential and stream function. These functions both satisfy the Laplace
equation and the contour lines represent lines of constant head (equipotentials) and lines
tangent to flowpaths (streamlines). Together, the potential function and the stream function
form the complex potential, where the potential is the real part, and the stream function is the
imaginary part.

The construction of a flownet provides an approximate solution to the flow problem, but it can
be quite good even for problems with complex geometries by following a few simple rules
(initially developed by Philipp Forchheimer around 1900, and later formalized by Arthur
Casagrande in 1937) and a little practice:

 streamlines and equipotentials meet at right angles (including the boundaries),


103
 diagonals drawn between the cornerpoints of a flownet will meet each other at right
angles (useful when near singularities),
 streamtubes and drops in equipotential can be halved and should still make squares
(useful when squares get very large at the ends),
 flownets often have areas which consist of nearly parallel lines, which produce true
squares; start in these areas — working towards areas with complex geometry,
 many problems have some symmetry (e.g., radial flow to a well); only a section of the
flownet needs to be constructed,
 the sizes of the squares should change gradually; transitions are smooth and the
curved paths should be roughly elliptical or parabolic in shape.

Example flownets

The first flownet pictured here (modified from Craig, 1997) illustrates and quantifies the flow
which occurs under the dam (flow is assumed to be invariant along the axis of the dam —
valid near the middle of the dam); from the pool behind the dam (on the right) to the tailwater
downstream from the dam (on the left).

There are 16 green equipotential lines (15 equal drops in hydraulic head) between the 5 m
upstream head to the 1m downstream head (4 m / 15 head drops = 0.267 m head drop
between each green line). The blue streamlines (equal changes in the streamfunction
between the two no-flow boundaries) show the flowpath taken by water as it moves through
the system; the streamlines are everywhere tangent to the flow velocity.

Example flownet 2,

The second flownet pictured here (modified from Ferris, et al., 1962) shows a flownet being
used to analyze map-view flow (invariant in the vertical direction), rather than a cross-section.
Note that this problem has symmetry, and only the left or right portions of it needed to have
been done. To create a flownet to a point sink (a singularity), there must be a recharge
boundary nearby to provide water and allow a steady-state flowfield to develop.

104
Flownet results

Darcy's law describes the flow of water through the flownet. Since the head drops are uniform
by construction, the gradient is inversely proportional to the size of the blocks. Big blocks
mean there is a low gradient, and therefore low discharge (hydraulic conductivity is assumed
constant here).

An equivalent amount of flow is passing through each streamtube (defined by two adjacent
blue lines in diagram), therefore narrow streamtubes are located where there is more flow.
The smallest squares in a flownet are located at points where the flow is concentrated (in this
diagram they are near the tip of the cutoff wall, used to reduce dam underflow), and high flow
at the land surface is often what the civil engineer is trying to avoid, being concerned about
soil piping or dam failure.

CHAPTER 13
QUANTITATIVE DETERMINATION OF DEPTH AND SPACING OF DRAINS

See attached powerpoint

105
106
107
Drain Depth and Spacing

L is drain spacing; h is mid drain water table height (m) above drain level; Do is depth of
aquifer from drain level to impermeable layer(m); q is the water input rate(m/day) = specific
discharge or drainage coefficient; K is hydraulic conductivity(m/day); H is the depth to water
table.

Design Water table depth (H):


This is the minimum depth below the surface at which the water table should be
controlled and is determined by farming needs especially crop tolerance to water.
Typically, it varies from 0.5 to 1.5 m.

Design Depth of Drain


The deeper a drain is put, the larger the spacing and the more economical the design
becomes.
Drain depth, however, is constrained by soil and machinery limitations.
 
Table : Typical Drain Depths(D)
Soil Type Drain Depth (m)
108
Sand 0.6
Sandy loam 0.8 - 1.0
Silt loam 0.8 - 1.8
Clay loam 0.6 - 0.8
Peat 1.2 - 1.5

Drain Spacing (L)

This is normally determined using the Hooghoudt equation. It states that Hooghoudt
equation states that for ditches reaching the impermeable layer:
 
L2 = 8 K Do h + 4 K h
q q
(See definitions of terms above)
For tube drains which do not reach the impermeable layer, the equation can be
modified as:
L = 8Kdh + 4 K h2
q q
Where d is called the Houghoudt equivalent d. The equation for tube drains can be
solved using trial and error method or the graphical method.

Example:
For the drainage design of an irrigated area, drain pipes with a radius of 0.1 m are used.
They are placed at a depth of 1.8 m below the soil surface. A relatively impermeable soil
layer was found at a depth of 6.8 m below the surface. From auger hole tests, the hydraulic
conductivity above this layer was estimated as 0.8 m/day. The average irrigation losses,
which recharge the groundwater, are 40 mm per 20 days so the average discharge of the
drain system amounts to 2 mm/day.
Estimate the drain spacing, if the depth of the water table is 1.2 m.

Solutiion:

Trial One

109
Assume L = 75 m, from Houghout d table provided, with L = 75 m, and Do = 5 m, d =
3.49 m.
From equation (1), L2 = (1920 x 3.49) + 576 = 7276.8; L = 85.3 m
Comment: The L chosen is small since 75 < 85.3 m
 
Try L = 100 m, from table, d = 3.78
From (1), L 2 = (1920 x 3.78) + 576 = 7833.6 ;
L = 88.51m
Comment: Since 88.51 < 100, try a smaller L; L should be between 75 and 100 m.

110
Try L = 90 m, d =3.49 + 15/25(3.78 - 3.49) = 3.66 m
L2 = (1920 x 3.66) + 576 = 7603.2 m ; L = 87 m
111
Comment: Since 87 < 90, try a smaller L; L should be between 75 and 90.
 
Try L = 87 m, d = 3.49 + 12/25(3.78 - 3.49) =3.63 m
L2 = (1920 x 3.63) + 576 = 7545.6; L = 86.87 m
Comment: The difference between the assumed and calculated L is <1, so : Drain
Spacing = 87 m.

Graphical Solution

Calculate 4 K h2 and 8Kh


q q
 
4 K h2 = 4 x 0.8 x 0.62 = 576;
q 0.002
8 K h = 8 x 0.8 x 0.6 = 1920
q 0.002  

Locate the two points on graph given and join.


For a value of Do = 5 m; produce downwards to meet the line. Read off the spacing
on the diagram
L = 87 m

112
Drain Diameters and Gradients
There are two approaches to design:
(a) Transport approach: Assumes that pipes are flowing full from top to end of
field. Assumes uniform flow. Widely used in United States, Canada and Germany.
Used to design collector drains.
(b) Drainage approach: Assumes that water enters the pipe all down the length as
it is perforated. This is more realistic. Widely used in United Kingdom, Holland and
Denmark. This is used to design lateral drainage pipes.

Parameters Required to use Solution Graphs


(a) Types of pipes: Pipes can be smooth or rough. Clay tiles and smooth plastic
pipes are smooth while corrugated plastic pipes are rough.
(b) Drainable area: The area drained by one lateral and is equal to the maximum
length of a lateral multiplied by drain spacing.
The whole area drained by the laterals discharging into a collector represents the
drainable area of the collector.
c) Specific discharge: Earlier defined. Same as drainage coefficient.
d) Silt safety factors: Used to account for the silting of pipes with time by making
the pipes bigger. 60, 75 and 100 % pipe capacity factors are indicated. This means
allowing 40, 25 and 0% respectively for silting.
e) Average hydraulic gradient(%): normally the soil slope.

Example:
The drainage design of a field is drain spacing = 30 m, length of drain lines = 200 m,
slope = 0.10%, specific discharge = 10 mm/day. Estimate drain diameter. Assume
60% silt factor and clay tiles.
 Solution: Area to be drained by one lateral = 30 m x 200 m = 6000 m 2 = 0.6 ha
Slope = average hydraulic gradient = 0.10% ; q = 10 mm/day
Using chart for smooth drains, nearest diameter = 70 mm inside diameter.

Drainage Filters
Filters for tile drains are permeable materials eg. gravel placed around the drains for
the purpose of improving the flow conditions in the area immediately surrounding the
drains as well as for improving bedding conditions.
Filters provide a high hydraulic conductivity around the drains which stabilizes the
soil around and prevent small particles from entering the lateral drains since they are
perforated.
a) Filters are needed to be gravel with same uniformity with the soil to be protected.
b) D15 Filter < 5 D85 Soil ; D15 Filter < 20 D15 Soil ; D50 Filter < 25 D50
Soil.
These are the filtration criteria.
To give adequate hydraulic conductivity, D85 Filter > 5 D15 Soil.
These criteria are difficult to achieve and should serve as guidelines.

Soils that Need Filters


 Uniform soils will cause problems while non-uniform ones since they are widely
distributed stabilize themselves.
b) Clays have high cohesion so cannot be easily moved so require no filters.
c) Big particles like gravel can hardly be moved due to their weight.
* Fine soils are then the soils that will actually need filters especially if they are
uniform.

113
CHAPTER 14
SURFACE DRAINS

DESIGN OF SURFACE DRAINAGE SYSTEMS:


Surface drainage involves the removal of excess water from the surface of the soil.
This is done by removing low spots where water accumulates by land forming or by
excavating ditches or a combination of the two.

Land forming is mechanically changing the land surface to drain surface water.
This is done by smoothing, grading, bedding or leveling.
Land smoothing is the shaping of the land to a smooth surface in order to eliminate
minor differences in elevation and this is accomplished by filling shallow depressions.
There is no change in land contour. Smoothing is done using land levelers or
planes

114
Land grading is shaping the land for drainage done by cutting, filling and smoothening to
planned continuous surface grade e.g. using bulldozers or scrapers.

CHAPTER 15
SUBSURFACE DRAINAGE SYSTEMS

DESIGN OF SUB-SURFACE DRAINAGE SYSTEMS


Sub-surface drainage is the removal of excess groundwater below the soil surface.
It aims at increasing the rate at which water will drain from the soil, and so lowering
the water table, thus increasing the depth of drier soil above the water table.
Sub-surface drainage can be done by open ditches or buried drains.

Sub-Surface Drainage Using Ditches

115
Ditches have lower initial cost than buried drains;
There is ease of inspection and ditches are applicable in some organic soils where
drains are unsuitable.
Ditches, however, reduce the land available for cropping and require more
maintenance that drains due to weed growth and erosion.

Sub-Surface Drains Using Buried Drains

Buried drains refer to any type of buried conduits having open joints or perforations,
which collect and convey drainage water.
116
They can be fabricated from clay, concrete, corrugated plastic tubes or any other
suitable material.
The drains can be arranged in a parallel, herringbone, double main or random
fashion.

Arrangements of Sub-Surface Drains

Sub-Surface Drainage Designs

The Major Considerations in Sub-surface Drainage Design Include:


Drainage Coefficient; - This is the rate of water removal used in drainage design to
obtain the desired protection of crops from excess surface or sub-surface water and
can be expressed in mm/day , m/day etc.
Drainage is different in Rain-Fed Areas and Irrigated Areas

In irrigated areas, water enters the groundwater from:


Deep percolation,
Leaching requirement,
Seepage or
Conveyance losses from watercourses and canals and
Rainfall for some parts of the world.

Drain Depth and Spacing;


Drain Diameters and Gradient;
Drainage Filters.

Example:
In the design of an irrigation system, the following properties exist: Soil field capacity
is 28% by weight, permanent wilting point is 17% by weight; Bulk density = 1.36 g/cm 3
; root zone depth is 1 m; peak ET is 5 mm/day; irrigation efficiency is 60%, water
conveyance efficiency is 80%, 50 % of water lost in canals contribute to seepage;
rainfall for January is 69 mm and evapotranspiration is 100 mm; salinity of irrigation

117
water is 0.80 mm hos/cm while that acceptable is 4 mmhos/cm. Compute the
drainage coefficient. Readily available moisture is 75% of AM.
Solution:
Readily available moisture (RAM) = ½ (FC - PWP) = 1/2(28 - 17) = 5.5%. In depth,
RAM = 0.055 x 1.36 x 1000 mm= 74.8 mm = Net irrigation
Shortest irrigation interval = RAM/peak ET = 74.8/5 = 15 days
With irrigation efficiency of 60 %, Gross irrigation requirement = 74.8/0.6 = 124.7
mm. This is per irrigation.
(a) Water losses = Gross - Net irrigation = 124.7 - 74.8 = 49.9 mm
Assuming 70% is deep percolation while 30% is wasted on the soil surface (Standard
assumption), deep percolation = 0.7 x 49.9 = 34.91 mm
b)Seepage
Conveyance efficiency, Ec = Water delivered to farm
Water released at dam
= 0.8
Water delivered to farm = Gross irrigation =124.7 mm
i.e. Water released = 124.7/0.8 = 155.9 mm
Excess water or water lost in canal = 155.9 - 124.7
= 31.2 mm
Since half of the water is seepage (given), the rest will be evaporation during
conveyance
Seepage = 1/2 x 31.2 mm = 15.6 mm
Leaching Reqd. = Ecirrig (ET - Rain ) = 0.8 (100 -69)
Ec accep 4
= 7.75 mm
This is for one month; for 15 days, we have 3.88 mm
 
(d) Rainfall = 69 mm; for 15 days, this is 34.5 mm
 
Note: In surface irrigation systems, deep percolation is much higher than leaching
requirement so only the former is used in computation.
It is assumed that excess water going down the soil as a result of deep percolation
can be used for leaching. In sprinkler system, leaching requirement may be greater
than deep percolation and can be used instead.
Neglecting Leaching Requirement, Total water input into drains is equal to:
34.91 + 15.6 + 34.5 = 85.01 mm
This is per 15 days, since irrigation interval is 15 days
Drainage coefficient = 85.01/15 = 5.67
= 6 mm/day.

118

You might also like