You are on page 1of 168

CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C.

Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

MOI UNIVERSITY
SCHOOL OF ENGINEERING
DEPARTMENT OF CIVIL AND STRUCTURAL
ENGINEERING

Lecture Notes
CVS 545E: IRRIGATION ENGINEERING

Instructor: Prof. Eng. Emmanuel C. Kipkorir (PhD, PE, MIEK),


Moi University,
School of Engineering,
Department of Civil and Structural Engineering,
P.O. Box 3900, Eldoret.
E-mail: emmanuel.kipkorir@yahoo.com
Tel 0722-598207.

October, 2016

1
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

CVS 545E: IRRIGATION ENGINEERING (3 Units)

Course Objectives
The purpose of this course is to introduce the students to agro‐hydrological principles and
equip them with knowledge and skills to design, implement and manage irrigation and
drainage infrastructure.

Expected Learning Outcomes of the course


At the end of the course the student should be able to:
1. Define the agricultural aspects of irrigation and be able to calculate the water demand of
different crops;
2. Design the facilities supplying water to the irrigation scheme;
3. Define different irrigation systems and be able to select the appropriate system;
4. Design a gravity irrigation system as far as structures, canal and field irrigation facilities
are concerned;
5. Design a sprinkler irrigation system as far as field irrigation facilities are concerned;
6. Design a drip irrigation system as far as field irrigation facilities are concerned;
7. Design surface and sub-surface drainage systems; and
8. Assess Environmental Impacts of Irrigation projects.

Course Content
Water demand: Crop water requirements: ETo, ETc, cropping calendar, rainfall, crop
irrigation requirement, irrigation efficiency, scheme irrigation requirement. Plant-Soil-Water
System: the soil, soil moisture contents, plant-soil-water system. Irrigation scheduling and
irrigation efficiency. Water supply: from rivers, lakes and groundwater, by artificial
facilities such as mobile weirs, reservoirs and pumps. Irrigation systems description:
Surface systems: furrow, basin, border; facilities for field irrigation by gravity. Pressure
systems: sprinkling, centre pivot and drip. Water quality: agricultural requirements with
regard to hygiene: salinity and alkalinity, flushing of the soil, choice of irrigation systems and
crop selection. Use of treated wastewater for irrigation: irrigation systems, crop types and
effects on public health. Water supply: from rivers, lakes and groundwater, by artificial
facilities such as mobile weirs, reservoirs and pumps. Design of irrigation systems: surface:
furrow, basin, border; sprinkler and drip. Design of conveyance and control structures:
river-weir, main intake structure, and sand trap; water level and discharge control structures;
canals: erosion, sedimentation, seepage, and lining; irrigation outlets and drop structures; and
pumping station on the riverbank including calculation of pressure lines and storage
reservoir. Water delivery methods: continuous flow, rotational, on-demand, semi-demand,
canal rotational and free demand and proportional supply. Design of surface and sub-
surface drainage systems: design of ditches, canals and pipe system. Environmental
issues: impacts of irrigation and drainage development on the environment.

Course outline
1. Crop water requirements: ETo, ETc, cropping calendar, rainfall, crop irrigation
requirement, irrigation efficiency, scheme irrigation requirement;
2. Plant-Soil-Water System: the soil, soil moisture contents, plant-soil-water system;
3. Irrigation schedule: Definitions; Real time scheduling and Practical irrigation schedules;
Adjustments of irrigation schedules to environment, Deficit irrigation
4. Irrigation systems: (i) Surface systems: furrow, basin, border, (ii) facilities for field
irrigation by gravity, (iii) Pressure systems: sprinkling, centre pivot and drip;

2
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

5. Agricultural water quality requirements: with regard to hygiene, salinity and alkalinity,
flushing of the soil, choice of irrigation systems and crop selection;
6. Water supply: from rivers, lakes and groundwater, by artificial facilities such as mobile
weirs, reservoirs and pumps;
7. Design of irrigation systems: (i) surface: furrow, basin, border, (ii) sprinkler, (iii) drip;
8. Design of: (i) river-weir, main intake structure, and sand trap, (ii) water level and
discharge control structures, (iii) canals: erosion, sedimentation, seepage, and lining, (iv)
irrigation outlets and drop structures, (v) pumping station on the riverbank including
calculation of pressure lines and storage reservoir;
9. Water delivery methods: continuous flow, rotational, on-demand, semi-demand, canal
rotational and free demand and proportional supply.
10. Design of surface and sub-surface drainage system: ditches and canal and pipe system;
and
11. EIA: Environmental Impacts of Irrigation projects and mitigation measures.

Teaching and Learning Methods


i. Lectures
ii. Tutorials
iii. Demonstrations and Presentations

Teaching Materials and Aids


i. Lecture hand book
ii. Text books
iii. Computer software simulation exercises
iv. Scientific calculators
v. Computers and internet

Course Assessment
i. Continuous Assessment and Tests 20%
ii. Computer laboratory Practicals 10%
iii. End of semester Examination 70%

Course References/Reference Texts


i. Melby Pete. 1995. Simplified Irrigation Design, 2nd Edition. John Wiley & Sons
ii. Kurt Hall. 2004. Basic Irrigation Systems and Design. Water Management
Specialists.
iii. Irrigation Association. 2011. Irrigation, 6th Edition.
iv. Allen, R.G., Pereira, L.S., Raes, D., Smith, M. 1998. Crop Evapotranspiration:
Guidelines for Computing Crop Water Requirements. FAO Irrigation and Drainage
Paper 56, Rome, Italy, p. 301.
v. Melby Pete. 1995. Simplified Irrigation Design, 2nd Edition. John Wiley & Sons.

Course Further Readings


i. Irrigation Association. 2010. Principles of Irrigation, 2nd Edition.

3
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Course Schedule
Week Topic Remarks
1 and 2 Crop water requirements: ETo, ETc, cropping calendar, Assignment 1-
rainfall, crop irrigation requirement, irrigation efficiency, Calculation of ETo
scheme irrigation requirement. using software.
Assignment 2-
Frequency analysis
of rainfall.
3 Plant-Soil-Water System: the soil, soil moisture contents,
plant-soil-water system.
4 Irrigation schedule: Definitions; Real time scheduling and Assignment 3-
Practical irrigation schedules; Adjustments of irrigation Irrigation schedule
schedules to environment, Deficit irrigation. computer exercise.
5 Overview of irrigation systems:
(i) Surface systems: furrow, basin, border,
(ii) facilities for field irrigation by gravity,
(iii) Pressure systems: sprinkling, centre pivot and drip.
6 Agricultural water quality requirements: with regard to
hygiene, salinity and alkalinity, flushing of the soil, choice
of irrigation systems and crop selection;

Water supply: from rivers, lakes and groundwater, by


artificial facilities such as mobile weirs, reservoirs and
pumps.
7, 8 and 9 Design of irrigation systems: Assignment 4-
(i) surface: furrow, basin, border; Design of sprinkler
(ii) sprinkler; and and drip irrigation
(iii) drip. system
10 Design of:
(i) river-weir, main intake structure, and sand trap,
(ii) water level and discharge control structures, CAT
(iii)canals: erosion, sedimentation, seepage, and lining,
(iv) irrigation outlets and drop structures,
(v) pumping station on the riverbank including
calculation of pressure lines and storage reservoir.
11 Water delivery methods: continuous flow, rotational, on- Submission of
demand, semi-demand, canal rotational and free demand Assignments 1, 2, 3
and proportional supply. and 4
12 Design of surface and sub-surface drainage system:
ditches and canal and pipe system.
13 EIA: Environmental Impacts of Irrigation projects and
mitigation measures.
14 or 15 Examination Three hour
written exam

4
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Table of Contents
Chapter 1-Introduction ........................................................................................................................................... 9
1.1 Need for irrigation ...................................................................................................................................... 9
1.2 Need for drainage ....................................................................................................................................... 9
1.3 Why drainage lags behind irrigation? .................................................................................................... 10
Chapter 2 - Introduction to evapotranspiration ..................................................................................................... 11
2.1 Evapotranspiration process ..................................................................................................................... 11
2.1.1 Evaporation ........................................................................................................................................ 11
2.1.2 Transpiration ..................................................................................................................................... 11
2.1.3 Evapotranspiration (ET) ................................................................................................................... 12
2.2 Factors affecting evapotranspiration ...................................................................................................... 12
2.2.1 Weather parameters .......................................................................................................................... 12
2.2.2 Crop factors........................................................................................................................................ 12
2.2.3 Management and environmental conditions ................................................................................... 13
2.3 Evapotranspiration concepts ................................................................................................................... 13
2.3.1 Reference crop evapotranspiration (ET o) ........................................................................................ 13
2.3.2 Crop evapotranspiration under standard conditions (ETc) ........................................................... 15
2.3.3 Crop evapotranspiration under non-standard conditions (ETc adj) ............................................... 15
2.4 Determining evapotranspiration ............................................................................................................. 16
2.4.1 ET measurement ................................................................................................................................ 16
2.4.2 ET computed from meteorological data .......................................................................................... 17
2.4.3 ET estimated from pan evaporation ................................................................................................ 17
2.5 Assignment 1 ............................................................................................................................................. 19
Chapter 3 ETc - Single crop coefficient (Kc) ........................................................................................................ 20
3.1 Length of growth stages ........................................................................................................................... 20
3.2 Crop coefficients ....................................................................................................................................... 25
3.2.1 Tabulated Kc values ........................................................................................................................... 25
3.3 Construction of the Kc curve.................................................................................................................... 31
3.3.1 Annual crops ...................................................................................................................................... 31
3.3.2 Forage crops ....................................................................................................................................... 31
3.3.3 Fruit trees ........................................................................................................................................... 31
3.4 Calculating ETc ......................................................................................................................................... 32
3.4.1 Graphical determination of Kc ......................................................................................................... 32
3.4.2 Numerical determination of Kc......................................................................................................... 33
Chapter 4 - Rainfall .............................................................................................................................................. 34
4.1 Definitions.................................................................................................................................................. 34
4.1.1 Historical or actual rainfall data ...................................................................................................... 34
4.1.2 Average or normal rainfall ............................................................................................................... 34
4.1.3 Dependable rainfall ........................................................................................................................... 35
4.1.4 Dry, normal and wet conditions ....................................................................................................... 35
4.1.5 Effective rainfall................................................................................................................................. 36
4.2 Frequency analysis.................................................................................................................................... 36
4.3 Effective rainfall........................................................................................................................................ 38
4.3.1 Infiltration and soil water storage .................................................................................................... 38
4.3.2 Estimation procedure ........................................................................................................................ 38
4.4 Assignment 2 ............................................................................................................................................. 39
Chapter 5 - Irrigation requirement ........................................................................................................................ 40
5.1 Definitions.................................................................................................................................................. 40
5.1.1 Crop irrigation requirement ............................................................................................................. 40
5.1.2 Scheme irrigation requirement ......................................................................................................... 40
5.2 First assessment of net irrigation requirement ...................................................................................... 41
5.2.1 Seasonal net irrigation requirement ................................................................................................. 41
5.2.2 Peak irrigation requirement ............................................................................................................. 41
5.2.3 Ground water contribution (Ge) ....................................................................................................... 41
5.2.4 Soil water contribution (Wb) ............................................................................................................. 42
5.2.5 Salinity control ................................................................................................................................... 42
5.3 Crop irrigation requirement .................................................................................................................... 44
5.4 Scheme irrigation requirement ................................................................................................................ 44
5.4.1 Irrigation efficiencies ......................................................................................................................... 44

5
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

5.5 Scheme water requirement ...................................................................................................................... 47


5.5.1 Seasonal requirement ........................................................................................................................ 47
5.5.2 Peak supply ........................................................................................................................................ 47
Chapter 6- Soil-Water-Plant Relationships ........................................................................................................... 48
6.1 The soil ....................................................................................................................................................... 48
6.1.1 Soil texture.......................................................................................................................................... 48
6.1.2 Soil structure ...................................................................................................................................... 49
6.1.3 Particle density ................................................................................................................................... 49
6.1.4 Bulk density ........................................................................................................................................ 49
6.2 Soil moisture .............................................................................................................................................. 49
6.2.1 Dry mass basis .................................................................................................................................... 50
6.2.2 Volume basis ...................................................................................................................................... 50
6.2.3 Equivalent depth ................................................................................................................................ 51
6.3 Soil moisture contents ............................................................................................................................... 52
6.3.1 Water retention curve ....................................................................................................................... 52
6.3.2 Saturation water content ................................................................................................................... 52
6.3.3 Field capacity ..................................................................................................................................... 52
6.3.4 Permanent wilting point .................................................................................................................... 53
6.3.5 Total available soil moisture ............................................................................................................. 53
6.4 Plant-Soil-Water system ........................................................................................................................... 54
6.4.1 The root system and rooting depth................................................................................................... 54
6.4.2 Readily Available Moisture (RAM) ................................................................................................. 56
6.5 Yield response to water ............................................................................................................................ 57
6.6 Actual evapotranspiration ....................................................................................................................... 59
6.7 Soil water balance ..................................................................................................................................... 60
Chapter 7- Irrigation scheduling ........................................................................................................................... 63
7.1 Definitions.................................................................................................................................................. 63
7.1.1 Root zone depletion............................................................................................................................ 63
7.1.2 Maximum depletion ........................................................................................................................... 63
7.1.3 Water application depth (DA) ........................................................................................................... 63
7.1.4 Water application duration (WAD) ................................................................................................. 64
7.1.5 Unit flow (qu) ...................................................................................................................................... 64
7.1.6 Water application time (ti) and Irrigation interval (INT) .............................................................. 65
7.2 Real time scheduling ................................................................................................................................. 66
7.2.1 Plant observation ............................................................................................................................... 66
7.2.2 Soil moisture meters .......................................................................................................................... 66
7.2.3 Cumulative pan evaporation............................................................................................................. 67
7.2.4 Soil water budget ............................................................................................................................... 67
7.3 Practical irrigation schedules .................................................................................................................. 67
7.3.1 Variable interval(s), fixed amount(s) ............................................................................................... 67
7.3.2 Fixed interval(s), variable amount(s) ............................................................................................... 67
7.3.3 Fixed intervals and fixed amounts.................................................................................................... 68
7.4 Adjustments of irrigation schedules to environment ............................................................................. 68
7.4.1 Various weather conditions............................................................................................................... 68
7.4.2 Different soil types ............................................................................................................................. 68
7.5 Deficit irrigation........................................................................................................................................ 68
Chapter 8.0 - Irrigation methods ........................................................................................................................... 69
8.1 Surface irrigation ...................................................................................................................................... 69
8.1.1 Types of surface irrigation methods ................................................................................................. 69
8.1.2 Capabilities and limitations .............................................................................................................. 73
8.1.3 A comparison ..................................................................................................................................... 76
8.2 Drip Irrigation .......................................................................................................................................... 76
8.2.1 Capabilities and limitations .............................................................................................................. 77
8.3 Sprinkler Irrigation .................................................................................................................................. 80
8.3.1 Wetting pattern .................................................................................................................................. 80
8.3.2 Types of sprinkler systems ................................................................................................................ 80
8.3.3 Capabilities and limitations .............................................................................................................. 83
8.4 Buried clay pot irrigation ......................................................................................................................... 85
Chapter 9- Irrigation water quality ....................................................................................................................... 87

6
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 10- Water Sources and Water Availability.............................................................................................. 90


10.1 Introduction ............................................................................................................................................ 90
10.1.1 Water sources ................................................................................................................................... 90
10.1.2 Water availability ............................................................................................................................ 90
10.1.3 Methods of tapping water from water sources .............................................................................. 91
10.2 Rivers ....................................................................................................................................................... 91
10.2.1 Introduction ..................................................................................................................................... 91
10.2.2 Tapping water from rivers .............................................................................................................. 92
10.2.3 Availability of river flow ................................................................................................................. 92
10.3 Reservoirs and lakes ............................................................................................................................... 93
10.3.1 Introduction ..................................................................................................................................... 93
10.3.2 Water availability in lakes and reservoirs ..................................................................................... 94
10.3.3 Water tapping from a reservoir...................................................................................................... 95
10.4 Groundwater ........................................................................................................................................... 95
10.4.1 Introduction ..................................................................................................................................... 95
10.4.2 Groundwater availability ................................................................................................................ 96
10.4.3 Pumping from wells ......................................................................................................................... 96
Chapter 11- Design of surface irrigation systems ................................................................................................. 97
11.1 General .................................................................................................................................................... 97
11.2 Principles of surface irrigation .............................................................................................................. 98
11.2.1 Hydraulic phases .............................................................................................................................. 98
11.2.2 Process and purpose ........................................................................................................................ 99
11.2.3 Method of operation ...................................................................................................................... 100
11.3 Computer applications in surface irrigation ...................................................................................... 101
11.3.1 Mathematical modelling ................................................................................................................ 101
11.3.2 Computer programs ...................................................................................................................... 101
11.3.3 Design, operation and evaluation ................................................................................................. 102
11.4 Parameters and variables involved ..................................................................................................... 102
11.4.1 Field parameters ............................................................................................................................ 103
11.4.2 Decision variables .......................................................................................................................... 108
11.4.3 Evaluation variables ...................................................................................................................... 109
11.5 The SURDEV package ......................................................................................................................... 113
Chapter 12 - Sprinkler Irrigation Design ............................................................................................................ 114
12.1 Farm systems vs Field systems ............................................................................................................ 114
12.2 Outline of sprinkler design procedure ................................................................................................ 114
12.3 Summary ............................................................................................................................................... 115
12.4 Design Example ..................................................................................................................................... 115
12.4 Exercise .................................................................................................................................................. 116
Chapter 13: Drip Irrigation System Design ........................................................................................................ 119
13.1 Basic Guidelines .................................................................................................................................... 119
13.1.1 Application uniformity .................................................................................................................. 119
13.1.2 Peak water consumption ............................................................................................................... 119
13.1.3 Durability ....................................................................................................................................... 119
13.1.4 Economic considerations ............................................................................................................... 119
13.2 The Design Procedure .......................................................................................................................... 119
13.2.1 Selection of emitter type and layout ............................................................................................. 119
13.2.2 Checking alternative layouts ......................................................................................................... 119
13.2.3 Water flow velocity ........................................................................................................................ 121
13.2.4 Spacing of laterals and emitters.................................................................................................... 121
13.2.5 Choosing emitters and laterals ..................................................................................................... 121
13.3 Design of drip irrigation system for row crops .................................................................................. 122
13.4 Design if irrigation system in greenhouses ......................................................................................... 129
Chapter 14 – Design of distribution network ...................................................................................................... 130
14.1 Introduction .......................................................................................................................................... 130
14.1.1 Conveyance structures .................................................................................................................. 130
14.1.2 Control structures .......................................................................................................................... 130
14.1.3 Outlets ............................................................................................................................................. 131
14.2 Open-channel structures ...................................................................................................................... 131
14.2.1 Conveyance structures .................................................................................................................. 131

7
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

14.2.2 Control structure ........................................................................................................................... 131


14.2.3 Best Hydraulic Section .................................................................................................................. 134
14.2.4 Outlets ............................................................................................................................................. 134
14.3 Pipe line system ..................................................................................................................................... 135
14.3.1 Conveyance structures .................................................................................................................. 135
14.3.2 Control structures .......................................................................................................................... 136
14.3.3 Outlets ............................................................................................................................................. 137
Chapter 15 - Water delivery ............................................................................................................................... 139
15.1 Delivery methods .................................................................................................................................. 139
15.1.1 Classification .................................................................................................................................. 139
15.1.2 Delivery flow rate, frequency and duration ................................................................................. 139
15.1.3 Operation ........................................................................................................................................ 140
15.2 Water delivery methods ....................................................................................................................... 142
15.2.1 Continuous Flow ............................................................................................................................ 142
15.2.2 Rotational ....................................................................................................................................... 143
15.2.3 On-demand ..................................................................................................................................... 145
15.2.4 Semi-demand .................................................................................................................................. 146
15.2.5 Canal Rotational and Free Demand ............................................................................................. 147
15.2.6 Proportional Supply ...................................................................................................................... 148
Chapter 16 - Surface and sub-surface drainage system ...................................................................................... 149
16.1 Introduction .......................................................................................................................................... 149
16.2 Design flows for land drainage ............................................................................................................ 149
16.3 Drainage ditches.................................................................................................................................... 150
16.4 Under-drains ......................................................................................................................................... 151
16.5 Flow of groundwater to drains ............................................................................................................ 151
16.6 Layout of a tile-drain system ............................................................................................................... 153
16.7 Design of a land-drainage system ........................................................................................................ 154
Chapter 17: Environmental Impacts of Irrigation Projects ................................................................................. 155
17.1 Introduction .......................................................................................................................................... 155
17.2 Water, a scarce resource ...................................................................................................................... 155
17.3 Sources of Environmental Impacts in Irrigated Agriculture ............................................................ 157
17.3.1 Environmental impact derived from the construction of irrigation projects ........................... 157
17.3.2 Environmental impact derived from water supply and operation of irrigation projects ........ 158
17.3.3 Environmental impact derived from irrigated agriculture management practices ................. 160
17.4 How can we mitigate the Environmental Impact of Irrigation? ....................................................... 164
17.4.1 Policy Interventions ....................................................................................................................... 165
17.4.2 Engineering interventions ............................................................................................................. 166
17.4.3 System management interventions ............................................................................................... 166
17.4.4 Irrigation/agronomic practices interventions .............................................................................. 166
17.5 Final Remarks ....................................................................................................................................... 166

8
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 1-Introduction
According to a Hindu Epic hymn “No grain is ever produced without water, but too much
water tends to spoil the grain. An inundation is as injurious to growth as the scarcity of
water”. To harvest good quality grain when rainfall is not sufficient, as is the case in arid and
semi-arid areas, effective irrigation system need to be designed and constructed using
engineering principles followed by good irrigation management. It is always emphasised in
the reports of Drainage Boards that irrigation and drainage should go together.

1.1 Need for irrigation

During the 20th century the rate of water withdrawals in the world increased much faster than
population, owing mostly to irrigation and industrial development. Although the current
irrigated area of the world covers about 17 percent of the total cultivated area - 2/3 is in the
developing countries - and contributes about one third of the world food production, it
actually accounts for 73 percent of the world water consumption today. For developing
countries this figure is even as high as 85 percent and is seen as the main factor behind
increasing global water crisis.

In the next few decades the world’s population is expected to grow from 6.2 billion today to
at least 8 billion by the year 2025, with about 90 percent of the increase being added to the
developing world. It is therefore clear that achieving food security will continue to pose
major challenges to decision-makers in the next few decades. At a time when industrial and
municipal users are in competition with agriculture for the limited water, farmers have to
grow more food crops with the same or smaller amount of water, to feed the increasing
population.

The key to increase world's food production without deepening the global water scarcity is
improving the efficiency in the use of irrigation water. At irrigation scheme and farm level,
irrigation efficiency can be sometimes as low as 30 percent. The gains from introducing
effective water management and improved irrigation and water-control techniques and
technology can be tremendous in terms of water saving as well as in increase in productivity,
stabilization of erratic food production and environment conservation. If there is still a water
shortage, the limited amount of water to the crops should be distributed optimally.

1.2 Need for drainage

Irrigation is usually needed to get economically viable crop yields in irrigated lands of the
(semi-) arid areas. Application of irrigation water is usually associated with soil salinisation.
Therefore, an extra amount of water, in addition to actual evapotranspiration, is required to
leach the salts below the root zone before it affects the crop yields adversely. During this
process, water applied in excess of the soil water holding capacity of the root zone percolates
to the aquifer. In absence of natural drainage, the ground water table rises into the root zone
causing water logging. During fallow periods, this often leads to increase in capillary rise,
resulting to salinity in the root zone. Therefore, in irrigated areas, drainage is regarded as a
complementary activity to irrigation.

Drainage in irrigated lands of (semi-) arid areas is often practised to protect crops from
excess soil water conditions during rainy or irrigation periods and to prevent water table

9
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

induced soil salinisation during the fallow periods. The success of a drainage system depends
on its proper design. For irrigated areas, no specific drainage design criteria is available. It
has been common practice to use design criteria of humid areas to irrigated areas of (semi-)
arid regions.

1.3 Why drainage lags behind irrigation?

The practice to start the irrigation planning and execution with the available funds in the hope
that either drainage might not be needed or funds in future would be available for this activity
once the project starts yielding income continued to be in vogue. Following are some of the
reasons why drainage has not received due importance in the irrigation projects:
- To keep the benefit-cost ratio of the irrigation projects at a respectable level, drainage
component was neglected or inadequate provisions were made.
- Funds at the disposal of the planners were inadequate to provide for both irrigation and
drainage.
- The benefits of drainage were not as eloquently documented as that of irrigation.
Although irrigation and drainage go together was known, yet appreciation for drainage
was lacking both amongst the planners and the farmers.
- Besides these points, it has been the option of many experts that drainage may not be
needed immediately upon introduction of irrigation. A promise of providing drainage was
there but it was not implemented till the adverse effects appeared and farmers raised a hue
and cry.

10
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 2 - Introduction to evapotranspiration


This chapter explains the concepts of and the differences between reference crop
evapotranspiration (ETo) and crop evapotranspiration under standard conditions (ETc) and
various management and environmental conditions (ETc adj). It also examines the factors that
affect evapotranspiration, the units in which it is normally expressed and the way in which it
can be determined.

2.1 Evapotranspiration process

The combination of two separate processes whereby water is lost on the one hand from the
soil surface by evaporation and on the other hand from the crop by transpiration is referred to
as evapotranspiration (ET).

2.1.1 Evaporation
Evaporation is the process whereby liquid water is converted to water vapour (vaporization)
and removed from the evaporating surface (vapour removal). Water evaporates from a variety
of surfaces, such as lakes, rivers, pavements, soils and wet vegetation.

Energy is required to change the state of the molecules of water from liquid to vapour. Direct
solar radiation and, to a lesser extent, the ambient temperature of the air provide this energy.
The driving force to remove water vapour from the evaporating surface is the difference
between the water vapour pressure at the evaporating surface and that of the surrounding
atmosphere. As evaporation proceeds, the surrounding air becomes gradually saturated and
the process will slow down and might stop if the wet air is not transferred to the atmosphere.
The replacement of the saturated air with drier air depends greatly on wind speed. Hence,
solar radiation, air temperature, air humidity and wind speed are climatological parameters to
consider when assessing the evaporation process.

Where the evaporating surface is the soil surface, the degree of shading of the crop canopy
and the amount of water available at the evaporating surface are other factors that affect the
evaporation process. Frequent rains, irrigation and water transported upwards in a soil from a
shallow water table wet the soil surface. Where the soil is able to supply water fast enough to
satisfy the evaporation demand, the evaporation from the soil is determined only by the
meteorological conditions. However, where the interval between rains and irrigation becomes
large and the ability of the soil to conduct moisture to near the surface is small, the water
content in the topsoil drops and the soil surface dries out. Under these circumstances the
limited availability of water exerts a controlling influence on soil evaporation. In the absence
of any supply of water to the soil surface, evaporation decreases rapidly and may cease
almost completely within a few days.

2.1.2 Transpiration
Transpiration consists of the vaporization of liquid water contained in plant tissues and the
vapour removal to the atmosphere. Crops predominately lose their water through stomata.
These are small openings on the plant leaf through which gases and water vapour pass. The
water, together with some nutrients, is taken up by the roots and transported through the
plant. The vaporization occurs within the leaf, namely in the intercellular spaces, and the
vapour exchange with the atmosphere is controlled by the stomatal aperture. Nearly all water
taken up is lost by transpiration and only a tiny fraction is used within the plant.

11
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Transpiration, like direct evaporation, depends on the energy supply, vapour pressure
gradient and wind. Hence, radiation, air temperature, air humidity and wind terms should be
considered when assessing transpiration. The soil water content and the ability of the soil to
conduct water to the roots also determine the transpiration rate, as do waterlogging and soil
water salinity. The transpiration rate is also influenced by crop characteristics, environmental
aspects and cultivation practices. Different kinds of plants may have different transpiration
rates. Not only the type of crop, but also the crop development, environment and
management should be considered when assessing transpiration.

2.1.3 Evapotranspiration (ET)


Evaporation and transpiration occur simultaneously and there is no easy way of
distinguishing between the two processes. Apart from the water availability in the topsoil, the
evaporation from a cropped soil is mainly determined by the fraction of the solar radiation
reaching the soil surface. This fraction decreases over the growing period as the crop
develops and the crop canopy shades more and more of the ground area. When the crop is
small, water is predominately lost by soil evaporation, but once the crop is well developed
and completely covers the soil, transpiration becomes the main process. At sowing nearly
100% of ET comes from evaporation, while at full crop cover more than 90% of ET comes
from transpiration.

Units
The evapotranspiration rate is normally expressed in millimetres (mm) per unit time. The rate
expresses the amount of water lost from a cropped surface in units of water depth. The time
unit can be an hour, day, decade, month or even an entire growing period or year.

As one hectare has a surface of 10,000 m2 and 1 mm is equal to 0.001 m, a loss of 1 mm of


water corresponds to a loss of 10 m3 of water per hectare. In other words, 1 mm day-1 is
equivalent to 10 m3 ha-1 day-l.

2.2 Factors affecting evapotranspiration

Weather parameters, crop characteristics, management and environmental aspects are factors
affecting evaporation and transpiration.

2.2.1 Weather parameters


The principal weather parameters affecting evapotranspiration are radiation, air temperature,
humidity and wind speed. Several procedures have been developed to assess the evaporation
rate from these parameters. The evaporation power of the atmosphere is expressed by the
reference crop evapotranspiration (ETo). The reference crop evapotranspiration represents the
evapotranspiration from a standardized vegetated surface.

2.2.2 Crop factors


The crop type, variety and development stage should be considered when assessing the
evapotranspiration from crops grown in large, well-managed fields. Differences in resistance
to transpiration, crop height, crop roughness, reflection, ground cover and crop rooting
characteristics result in different ET levels in different types of crops under identical
environmental conditions. Crop evapotranspiration under standard conditions (ETc) refers to
the evaporating demand from crops that are grown in large fields under optimum soil water,

12
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

excellent management and environmental conditions, and achieve full production under the
given climatic conditions.

2.2.3 Management and environmental conditions


Factors such as soil salinity, poor land fertility, limited application of fertilizers, the presence
of hard or impenetrable soil horizons, the absence of control of diseases and pests and poor
soil management may limit the crop development and reduce the evapotranspiration. Other
factors to be considered when assessing ET are ground cover, plant density and the soil water
content. The effect of soil water content on ET is conditioned primarily by the magnitude of
the water deficit and the type of soil. On the other hand, too much water will result in
waterlogging which might damage the root and limit root water uptake by inhibiting
respiration.

When assessing the ET rate, additional consideration should be given to the range of
management practices that act on the climatic and crop factors affecting the ET process.
Cultivation practices and the type of irrigation method can alter the microclimate, affect the
crop characteristics or affect the wetting of the soil and crop surface. A windbreak reduces
wind velocities and decreases the ET rate of the field directly beyond the barrier. The effect
can be significant especially in windy, warm and dry conditions although evapotranspiration
from the trees themselves may offset any reduction in the field. Soil evaporation in a young
orchard, where trees are widely spaced, can be reduced by using a well-designed drip or
trickle irrigation system. The drippers apply water directly to the soil near trees, thereby
leaving the major part of the soil surface dry, and limiting the evaporation losses. The use of
mulches, especially when the crop is small, is another way of substantially reducing soil
evaporation. Anti-transpirants, such as stomata-closing, film-forming or reflecting material,
reduce the water losses from the crop and hence the transpiration rate.

Where field conditions differ from the standard conditions, correction factors are required to
adjust ETc. The adjustment reflects the effect on crop evapotranspiration of the environmental
and management conditions in the field.

2.3 Evapotranspiration concepts

Distinctions are made (Figure 2.1) between reference crop evapotranspiration (ETo), crop
evapotranspiration under standard conditions (ETc) and crop evapotranspiration under non-
standard conditions (ETc adj). ETo is a climatic parameter expressing the evaporation power of
the atmosphere. ETc refers to the evapotranspiration from excellently managed, large, well-
watered fields that achieve full production under the given climatic conditions. Due to sub-
optimal crop management and environmental constraints that affect crop growth and limit
evapotranspiration, ETc under non-standard conditions generally requires a correction.

2.3.1 Reference crop evapotranspiration (ETo)


The evapotranspiration rate from a reference surface, not short of water, is called the
reference crop evapotranspiration or reference evapotranspiration and is denoted as ET o. The
reference surface is a hypothetical grass reference crop with specific characteristics. The use
of other denominations such as potential ET is strongly discouraged due to ambiguities in
their definitions.

13
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

The concept of the reference evapotranspiration was introduced to study the evaporative
demand of the atmosphere independently of crop type, crop development and management
practices. As water is abundantly available at the reference evapotranspiring surface, soil
factors do not affect ET. Relating ET to a specific surface provides a reference to which ET
from other surfaces can be related. It obviates the need to define a separate ET level for each
crop and stage of growth. ETo values measured or calculated at different locations or in
different seasons are comparable as they refer to the ET from the same reference surface.

Figure 2.1. Reference (ETo), crop evapotranspiration under standard (ETc) and non-standard
conditions (ETc adj)
The only factors affecting ETo are climatic parameters. Consequently, ETo is a climatic
parameter and can be computed from weather data. ETo expresses the evaporating power of

14
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

the atmosphere at a specific location and time of the year and does not consider the crop
characteristics and soil factors. The FAO Penman-Monteith method is recommended as the
sole method for determining ETo. The method has been selected because it closely
approximates grass ETo at the location evaluated, is physically based, and explicitly
incorporates both physiological and aerodynamic parameters. Moreover, procedures have
been developed for estimating missing climatic parameters.

2.3.2 Crop evapotranspiration under standard conditions (ETc)


The crop evapotranspiration under standard conditions, denoted as ETc, is the
evapotranspiration from disease-free, well-fertilized crops, grown in large fields, under
optimum soil water conditions, and achieving full production under the given climatic
conditions.

The amount of water required to compensate the evapotranspiration loss from the cropped
field is defined as crop water requirement. Although the values for crop evapotranspiration
and crop water requirement are identical, crop water requirement refers to the amount of
water that needs to be supplied, while crop evapotranspiration refers to the amount of water
that is lost through evapotranspiration. The irrigation water requirement basically represents
the difference between the crop water requirement and effective precipitation. The irrigation
water requirement also includes additional water for leaching of salts and to compensate for
non-uniformity of water application.

Crop evapotranspiration can be calculated from climatic data and by integrating directly the
crop resistance, albedo and air resistance factors in the Penman-Monteith approach. As there
is still a considerable lack of information for different crops, the Penman-Monteith method is
used for the estimation of the standard reference crop to determine its evapotranspiration rate,
i.e., ETo. Experimentally determined ratios of ETc/ETo, called crop coefficients (Kc), are used
to relate ETc to ETo or ETc = Kc ETo.

Differences in leaf anatomy, stomatal characteristics, aerodynamic properties and even


albedo cause the crop evapotranspiration to differ from the reference crop evapotranspiration
under the same climatic conditions. Due to variations in the crop characteristics throughout
its growing season, Kc for a given crop changes from sowing till harvest.

2.3.3 Crop evapotranspiration under non-standard conditions (ETc adj)


The crop evapotranspiration under non-standard conditions (ETc adj) is the evapotranspiration
from crops grown under management and environmental conditions that differ from the
standard conditions. When cultivating crops in fields, the real crop evapotranspiration may
deviate from ETc due to non-optimal conditions such as the presence of pests and diseases,
soil salinity, low soil fertility, water shortage or waterlogging. This may result in scanty plant
growth, low plant density and may reduce the evapotranspiration rate below ETc.

The crop evapotranspiration under non-standard conditions is calculated by using a water


stress coefficient Ks and/or by adjusting Kc for all kinds of other stresses and environmental
constraints on crop evapotranspiration.

15
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

2.4 Determining evapotranspiration

2.4.1 ET measurement
Evapotranspiration is not easy to measure. Specific devices and accurate measurements of
various physical parameters or the soil water balance in lysimeters are required to determine
evapotranspiration. The methods are often expensive, demanding in terms of accuracy of
measurement and can only be fully exploited by well-trained research personnel. Although
the methods are inappropriate for routine measurements, they remain important for the
evaluation of ET estimates obtained by more indirect methods.

Soil water balance


Evapotranspiration can be determined by measuring the various components of the soil water
balance. The method consists of assessing the incoming and outgoing water flux into the crop
root zone over some time period. Irrigation (I) and rainfall (P) add water to the root zone. Part
of I and P might be lost by surface runoff (RO) and by deep percolation (DP) that will
eventually recharge the water table. Water might also be transported upward by capillary rise
(CR) from a shallow water table towards the root zone or even transferred horizontally by
subsurface flow in (SFin) or out of (SFout) the root zone. In many situations, however, except
under conditions with large slopes, SFin and SFout are minor and can be ignored. Soil
evaporation and crop transpiration deplete water from the root zone. If all fluxes other than
evapotranspiration (ET) can be assessed, the evapotranspiration can be deduced from the
change in soil water content ( SW) over the time period:

ET = I + P - RO - DP + CR ±  SF ±  SW

Some fluxes such as subsurface flow, deep percolation and capillary rise from a water table
are difficult to assess and short time periods cannot be considered. The soil water balance
method can usually only give ET estimates over long time periods of the order of week-long
or ten-day periods.

Lysimeters
By isolating the crop root zone from its environment and controlling the processes that are
difficult to measure, the different terms in the soil water balance equation can be determined
with greater accuracy. This is done in lysimeters where the crop grows in isolated tanks filled
with either disturbed or undisturbed soil. In precision weighing lysimeters, where the water
loss is directly measured by the change of mass, evapotranspiration can be obtained with an
accuracy of a few hundredths of a millimetre, and small time periods such as an hour can be
considered. In non-weighing lysimeters the evapotranspiration for a given time period is
determined by deducting the drainage water, collected at the bottom of the lysimeters, from
the total water input.

A requirement of lysimeters is that the vegetation both inside and immediately outside of the
lysimeter be perfectly matched (same height and leaf area index). This requirement has
historically not been closely adhered to in a majority of lysimeter studies and has resulted in
severely erroneous and unrepresentative ETc and Kc data.

As lysimeters are difficult and expensive to construct and as their operation and maintenance
require special care, their use is limited to specific research purposes.

16
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

2.4.2 ET computed from meteorological data


Owing to the difficulty of obtaining accurate field measurements, ET is commonly computed
from weather data. A large number of empirical or semi-empirical equations have been
developed for assessing crop or reference crop evapotranspiration from meteorological data.
Some of the methods are only valid under specific climatic and agronomic conditions and
cannot be applied under conditions different from those under which they were originally
developed.

Numerous researchers have analysed the performance of the various calculation methods for
different locations. As a result of an Expert Consultation held in May 1990, the FAO
Penman-Monteith method is now recommended as the standard method for the definition
and computation of the reference evapotranspiration, ETo. The ET from crop surfaces under
standard conditions is determined by crop coefficients (Kc) that relate ETc to ETo. The ET
from crop surfaces under non-standard conditions is adjusted by a water stress coefficient
(Ks) and/or by modifying the crop coefficient.

2.4.3 ET estimated from pan evaporation


Evaporation from an open water surface provides an index of the integrated effect of
radiation, air temperature, air humidity and wind on evapotranspiration. However, differences
in the water and cropped surface produce significant differences in the water loss from an
open water surface and the crop. The pan has proved its practical value and has been used
successfully to estimate reference evapotranspiration by observing the evaporation loss from
a water surface and applying empirical coefficients (pan coefficient-Kp) to relate pan
evaporation (Eo) to ETo. ETo = KpEo.

In selecting the appropriate pan coefficient, not only the pan type, but also the ground cover
in the station, its surroundings as well as the general wind and humidity conditions, should be
checked. The siting of the pan and the pan environment also influence the results. This is
particularly so where the pan is placed in fallow rather than cropped fields. Two cases are
commonly considered: Case A where the pan is sited on a short green (grass) cover and
surrounded by fallow soil; and Case B where the pan is sited on fallow soil and surrounded
by a green crop (Figure 2.2).

Figure 2.2: Two cases of evaporation pan siting and their environment.

17
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Depending on the type of pan and the size and state of the upwind buffer zone (fetch), pan
coefficients will differ. The larger the upwind buffer zone, the more the air moving over the
pan will be in equilibrium with the buffer zone. At equilibrium with a large fetch, the air
contains more water vapour and less heat in Case A than in Case B. Pan coefficients for the
Class A pan for different ground cover, fetch and climatic conditions are presented in Table
2.1.

Table 2.1: Pan coefficients (Kp) for Class A pan for different pan siting and environment and
different levels of mean relative humidity and wind speed (FAO Irrigation and
Drainage Paper No. 24)

The Class A Evaporation pan (Figure 2.3) is circular, 120.7 cm in diameter and 25 cm deep.
It is made of galvanized iron (22 gauge) or Monel metal (0.8 mm). The pan is mounted on a
wooden open frame platform which is 15 cm above ground level. The soil is built up to
within 5 cm of the bottom of the pan. The pan must be level. It is filled with water to 5 cm
below the rim, and the water level should not be allowed to drop to more than 7.5 cm below
the rim. The water should be regularly renewed, at least weekly, to eliminate extreme
turbidity. The pan, if galvanized, is painted annually with aluminium paint.

Screens over the pan are not a standard requirement and should preferably not be used. Pans
should be protected by fences to keep animals from drinking. The site should preferably be
under grass, 20 by 20 m, open on all sides to permit free circulation of the air. It is preferable
that stations be located in the centre or on the leeward side of large cropped fields.

Pan readings are taken daily in the early morning at the same time that precipitation is
measured. Measurements are made in a stilling well that is situated in the pan near one edge.
The stilling well is a metal cylinder of about 10 cm in diameter and some 20 cm deep with a
small hole at the bottom.

18
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Figure 2.3: Details of Class A pan

2.5 Assignment 1
Calculate with the ETo calculator software for each of the 12 months of the year mean monthly reference
evapotranspiration (ETo) for Eldoret
1. Introduction (explaining the ETo concept and the objective of the assignment);
2. Material and methods (Input data, FAO Penman-Monteith equation, assessment of missing data);
3. Results and discussion (a) ET assessed with Tmin and Tmax; (b) previous climatic data + vapour
pressure; (c) previous climatic data + solar radiation; (d) previous climatic data + wind speed.
4. Conclusions.

Data Given: Climatic data


Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
Tmax (OC) 23.6 24.4 24.6 23.5 22.6 21.8 21.3 21.2 22.6 22.6 22.0 22.6
Tmin (OC) 9.6 9.4 10.0 10.6 10.4 9.5 9.4 9.3 9.0 9.9 10.2 9.9
RHmean (%) 63 61 62 68 71 75 77 76 69 68 69 66
Sun hours (hrs) 8.90 8.80 8.50 7.60 7.90 7.00 5.50 6.30 7.50 7.50 7.70 8.60
Wind (km/d) 143 136 152 126 102 86 82 87 103 127 157 163

Station location:
Latitute = 0O 32’ N
Longitude = 35O 17’ E
Altitude = 2240 m asl.

19
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 3 ETc - Single crop coefficient (Kc)


This chapter deals with the calculation of crop evapotranspiration (ETc) under standard
conditions. No limitations are placed on crop growth or evapotranspiration from soil water
and salinity stress, crop density, pests and diseases, weed infestation or low fertility. ETc is
determined by the crop coefficient approach whereby the effect of the various weather
conditions are incorporated into ETo and the crop characteristics into the Kc coefficient:

ETc = Kc ETo (3.1)

The effect of both crop transpiration and soil evaporation are integrated into a single crop
coefficient. The Kc coefficient incorporates crop characteristics and averaged effects of
evaporation from the soil. For normal irrigation planning and management purposes, for the
development of basic irrigation schedules, and for most hydrologic water balance studies,
average crop coefficients are relevant and more convenient than the Kc computed on a daily
time step using a separate crop and soil coefficient.

The calculation procedure for crop evapotranspiration, ETc, consists of:


- identifying the crop growth stages, determining their lengths, and selecting the
corresponding Kc coefficients;
- adjusting the selected Kc coefficients for frequency of wetting or climatic
conditions during the stage;
- constructing the crop coefficient curve (allowing one to determine Kc values
for any period during the growing period); and
- calculating ETc as the product of ETo and Kc.

3.1 Length of growth stages

FAO Irrigation and Drainage Paper No. 24 provides general lengths for the four distinct
growth stages and the total growing period for various types of climates and locations. This
information has been supplemented from other sources and is summarised in Table 3.1.

In some situations, the time of emergence of vegetation and the time of effective full cover
can be predicted using cumulative degree-based regression equations or by more
sophisticated plant growth models. These types of models should be verified or validated for
the local area or for a specific crop variety using local observations.

20
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

TABLE 3.1. Lengths of crop development stages* for various planting periods and climatic regions (days)
Crop Init. Dev. Mid Late Total Plant Date Region
(Lini) (Ldev) (Lmid) (Llate)
a. Small Vegetables
Broccoli 35 45 40 15 135 Sept Calif. Desert, USA
Cabbage 40 60 50 15 165 Sept Calif. Desert, USA
Carrots 20 30 50/30 20 100 Oct/Jan Arid climate
30 40 60 20 150 Feb/Mar Mediterranean
30 50 90 30 200 Oct Calif. Desert, USA
Cauliflower 35 50 40 15 140 Sept Calif. Desert, USA
Celery 25 40 95 20 180 Oct (Semi) Arid
25 40 45 15 125 April Mediterranean
30 55 105 20 210 Jan (Semi) Arid
1
Crucifers 20 30 20 10 80 April Mediterranean
25 35 25 10 95 February Mediterranean
30 35 90 40 195 Oct/Nov Mediterranean
Lettuce 20 30 15 10 75 April Mediterranean
30 40 25 10 105 Nov/Jan Mediterranean
25 35 30 10 100 Oct/Nov Arid Region
Onion (dry) 15 25 70 40 150 April Mediterranean
20 35 110 45 210 Oct; Jan. Arid Region; Calif.
Onion (green) 25 30 10 5 70 April/May Mediterranean
20 45 20 10 95 October Arid Region
30 55 55 40 180 March Calif., USA
Onion (seed) 20 45 165 45 275 Sept Calif. Desert, USA
Spinach 20 20 15/25 5 60/70 Apr; Mediterranean
Sep/Oct
20 30 40 10 100 November Arid Region
Radish 5 10 15 5 35 Mar/Apr Medit.; Europe
10 10 15 5 40 Winter Arid Region
b. Vegetables - Solanum Family (Solanaceae)
Egg plant 30 40 40 20 130\1 October Arid Region
30 45 40 25 40 May/June Mediterranean
Sweet peppers (bell) 25/30 35 40 20 125 April/June Europe and Medit.
30 40 110 30 210 October Arid Region
Tomato 30 40 40 25 135 January Arid Region
35 40 50 30 155 Apr/May Calif., USA
25 40 60 30 155 Jan Calif. Desert, USA
35 45 70 30 180 Oct/Nov Arid Region
30 40 45 30 145 April/May Mediterranean
c. Vegetables - Cucumber Family (Cucurbitaceae)
Cucumber 20 30 40 15 105 June/Aug Arid Region
25 35 50 20 130 Nov; Feb Arid Region
Pumpkin, Winter 20 30 30 20 100 Mar, Aug Mediterranean
squash 25 35 35 25 120 June Europe
Sweet melons 25 35 40 20 120 May Mediterranean
30 30 50 30 140 March Calif., USA
15 40 65 15 135 Aug Calif. Desert, USA
30 45 65 20 160 Dec/Jan Arid Region

21
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Water melons 20 30 30 30 110 April Italy


10 20 20 30 80 Mat/Aug Near East (desert)
d. Roots and Tubers
Beets, table 15 25 20 10 70 Apr/May Mediterranean
25 30 25 10 90 Feb/Mar Mediterranean & Arid
Cassava: year 1 20 40 90 60 210 Rainy Tropical regions
year 2 150 40 110 60 360 Season
Potato 25 30 30/45 30 115/130 Jan/Nov (Semi) Arid Climate
25 30 45 30 130 May Continental Climate
30 35 50 30 145 April Europe
45 30 70 20 165 Apr/May Idaho, USA
30 35 50 25 140 Dec Calif. Desert, USA
Sweet potato 20 30 60 40 150 April Mediterranean
15 30 50 30 125 Rainy seas. Tropical regions
Sugarbeet 30 45 90 15 180 March Calif., USA
25 30 90 10 155 June Calif., USA
25 65 100 65 255 Sept Calif. Desert, USA
50 40 50 40 180 April Idaho, USA
25 35 50 50 160 May Mediterranean
45 75 80 30 230 November Mediterranean
35 60 70 40 205 November Arid Regions
e. Legumes (Leguminosae)
Beans (green) 20 30 30 10 90 Feb/Mar Calif., Mediterranean
15 25 25 10 75 Aug/Sep Calif., Egypt, Lebanon
Beans (dry) 20 30 40 20 110 May/June Continental Climates
15 25 35 20 95 June Pakistan, Calif.
25 25 30 20 100 June Idaho, USA
Faba bean, broad bean 15 25 35 15 90 May Europe
20 30 35 15 100 Mar/Apr Mediterranean
- dry 90 45 40 60 235 Nov Europe
- green 90 45 40 0 175 Nov Europe
Green gram, cowpeas 20 30 30 20 110 March Mediterranean
Groundnut 25 35 45 25 130 Dry West Africa
35 35 35 35 140 Season High Latitudes
35 45 35 25 140 May/June Mediterranean
Lentil 25 35 70 40 170 Oct/Nov Arid Region
Peas 15 25 35 15 90 May Europe
20 30 35 15 100 Mar/Apr Mediterranean
Soybeans 15 15 40 15 85 Dec Tropics
20 30/35 60 25 140 May Central USA
f. Perennial Vegetables (with winter dormancy and initially bare or mulched soil)
Artichoke 40 40 250 30 360 Apr (1st yr) California
20 25 250 30 325 May (2nd (cut in May)
yr)
Asparagus 50 30 100 50 230 Feb Warm Winter
90 30 200 45 365 Feb Mediterranean
g. Fibre Crops
Cotton 30 50 60 55 195 Mar-May Egypt; Pakistan; Calif.
45 90 45 45 225 Mar Calif. Desert, USA

22
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

30 50 60 55 195 Sept Yemen


30 50 55 45 180 April Texas
h. Oil Crops
Castor beans 25 40 65 50 180 March (Semi) Arid Climates
20 40 50 25 135 Nov. Indonesia
Safflower 20 35 45 25 125 April California, USA
25 35 55 30 145 Mar High Latitudes
35 55 60 40 190 Oct/Nov Arid Region
Sunflower 25 35 45 25 130 April/May Medit.; California
i. Cereals
Barley/Oats/Wheat 15 25 50 30 120 November Central India
20 25 60 30 135 March/Apr 35-45 °L
15 30 65 40 150 July East Africa
40 30 40 20 130 Apr
40 60 60 40 200 Nov
20 50 60 30 160 Dec Calif. Desert, USA
Winter Wheat 202 602 70 30 180 December Calif., USA
30 140 40 30 240 November Mediterranean
160 75 75 25 335 October Idaho, USA
Grains (small) 20 30 60 40 150 April Mediterranean
25 35 65 40 165 Oct/Nov Pakistan; Arid Reg.
Maize (grain) 30 50 60 40 180 April East Africa (alt.)
25 40 45 30 140 Dec/Jan Arid Climate
20 35 40 30 125 June Nigeria (humid)
20 35 40 30 125 October India (dry, cool)
30 40 50 30 150 April Spain (spr, sum.); Calif.
30 40 50 50 170 April Idaho, USA
Maize (sweet) 20 20 30 10 80 March Philippines
20 25 25 10 80 May/June Mediterranean
20 30 50/30 10 90 Oct/Dec Arid Climate
30 30 30 103 110 April Idaho, USA
20 40 70 10 140 Jan Calif. Desert, USA
Millet 15 25 40 25 105 June Pakistan
20 30 55 35 140 April Central USA
Sorghum 20 35 40 30 130 May/June USA, Pakis., Med.
20 35 45 30 140 Mar/April Arid Region
Rice 30 30 60 30 150 Dec; May Tropics; Mediterranean
30 30. 80 40 180 May Tropics
j. Forages
Alfalfa 4 1st cutting 10 20 20 10 60 Jan Apr Calif., USA.
cycle (last - 4°C)
Alfalfa 4, other cutting 5 10 10 5 30 Mar Calif., USA.
cycles 5 20 10 10 45 Jun Idaho, USA.
Bermuda for seed 10 25 35 35 105 March Calif. Desert, USA
Bermuda for hay 10 15 75 35 135 --- Calif. Desert, USA
(several cuttings)
Grass Pasture 4 10 20 -- -- -- 7 days before last -4°C in
spring until 7 days after first
-4°C in fall
Sudan, 1st cutting 25 25 15 10 75 Apr Calif. Desert, USA

23
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

cycle
Sudan, other cutting 3 15 12 7 37 June Calif. Desert, USA
cycles
k. Sugar Cane
Sugarcane, virgin 35 60 190 120 405 Low Latitudes
50 70 220 140 480 Tropics
75 105 330 210 720 Hawaii, USA
Sugarcane, ratoon 25 70 135 50 280 Low Latitudes
30 50 180 60 320 Tropics
35 105 210 70 420 Hawaii, USA
l. Tropical Fruits and Trees
Banana, 1st yr 120 90 120 60 390 Mar Mediterranean
nd
Banana, 2 yr 120 60 180 5 365 Feb Mediterranean
Pineapple 60 120 600 10 790 Hawaii, USA
m. Grapes and Berries
Grapes 20 40 120 60 240 April Low Latitudes
20 50 75 60 205 Mar Calif., USA
20 50 90 20 180 May High Latitudes
30 60 40 80 210 April Mid Latitudes (wine)
n. Fruit Trees
Citrus 60 90 120 95 365 Jan Mediterranean
Deciduous Orchard 20 70 90 30 210 March High Latitudes
20 70 120 60 270 March Low Latitudes
Olives 30 90 60 90 2705 March Mediterranean
Pistachios 20 60 30 40 150 Feb Mediterranean
o. Wetlands - Temperate Climate
Wetlands (Cattails, 10 30 80 20 140 May Utah, USA; killing frost
Bulrush)
Wetlands (short veg.) 180 60 90 35 365 November frost-free climate
* Lengths of crop development stages provided in this table are indicative of general conditions, but may vary substantially from
region to region, with climate and cropping conditions, and with crop variety. The user is strongly encouraged to obtain
appropriate local information.
1
Crucifers include cabbage, cauliflower, broccoli, and Brussel sprouts. The wide range in lengths of seasons is due to varietal
and species differences.
2
These periods for winter wheat will lengthen in frozen climates according to days having zero growth potential and wheat
dormancy. Under general conditions and in the absence of local data, fall planting of winter wheat can be presumed to occur in
northern temperate climates when the 10-day running average of mean daily air temperature decreases to 17° C or December 1,
whichever comes first. Planting of spring wheat can be presumed to occur when the 10-day running average of mean daily air
temperature increases to 5° C. Spring planting of maize-grain can be presumed to occur when the 10-day running average of
mean daily air temperature increases to 13° C.
3
The late season for sweet maize will be about 35 days if the grain is allowed to mature and dry.
4
In climates having killing frosts, growing seasons can be estimated for alfalfa and grass as:
alfalfa: last -4° C in spring until first -4° C in fall (Everson, D. O., M. Faubion and D. E. Amos 1978. Freezing
temperatures and growing seasons in Idaho. Univ. Idaho Agric. Exp. station bulletin 494.18 p)
grass: 7 days before last -4° C in spring and 7 days after last -4° C in fall (Kruse E. G. and Haise, H. R. 1974. "Water
use by native grasses in high altitude Colorado meadows." USDA Agric. Res. Service, Western Region report ARS-
W-6-1974. 60 pages)
5
Olive trees gain new leaves in March. See footnote 24 of Table 12 for additional information, where the Kc continues outside of
the "growing period".

Primary source: FAO Irrigation and Drainage Paper 24 (Doorenbos and Pruitt, 1977), Table 22.

24
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

The lengths of the initial and development periods may be relatively short for deciduous trees
and shrubs that can develop new leaves in the spring at relatively fast rates.

The rate at which vegetation cover develops and the time at which it attains effective full
cover are affected by weather conditions in general and by mean daily air temperature in
particular. Therefore, the length of time between planting and effective full cover will vary
with climate, latitude, elevation and planting date. It will also vary with cultivar (crop
variety). Generally, once the effective full cover for a plant canopy has been reached, the rate
of further phenological development (flowering, seed development, ripening, and senescence)
is more dependent on plant genotype and less dependent on weather.

The end of the mid-season and beginning of the late season is usually marked by senescence
of leaves, often beginning with the lower leaves of plants. The length of the late season
period may be relatively short (less than 10 days) for vegetation killed by frost (for example,
maize at high elevations in latitudes > 40°N) or for agricultural crops that are harvested fresh
(for example, table beets and small vegetables).

High temperatures may accelerate the ripening and senescence of crops. Long duration of
high air temperature (> 35°C) can cause some crops such as turf grass to go into dormancy. If
severely high air temperatures are coupled with moisture stress, the dormancy of grass can be
permanent for the remainder of the growing season. Moisture stress or other environmental
stresses will usually accelerate the rate of crop maturation and can shorten the mid and late
season growing periods.

The values in Table 3.1 are useful only as a general guide and for comparison purposes. The
listed lengths of growth stages are average lengths for the regions and periods specified and
are intended to serve only as examples. Local observations of the specific plant stage
development should be used, wherever possible, to incorporate effects of plant variety,
climate and cultural practices. Local information can be obtained by interviewing farmers,
ranchers, agricultural extension agents and local researchers, by conducting local surveys, or
by remote sensing. When determining stage dates from local observations, the guidelines and
visual descriptions may be helpful.

3.2 Crop coefficients

Changes in vegetation and ground cover mean that the crop coefficient Kc varies during the
growing period. The trends in Kc during the growing period are represented in the crop
coefficient curve. Only three values for Kc are required to describe and construct the crop
coefficient curve: those during the initial stage (Kc ini), the mid-season stage (Kc mid) and at the
end of the late season stage (Kc end).

3.2.1 Tabulated Kc values


Table 3.2 lists typical values for Kc ini, Kc mid and Kc end for various agricultural crops. The
coefficients presented are organized by group type (i.e., small vegetables, legumes, cereals,
etc.) to assist in locating the crop in the table and to aid in comparing crops within the same
group. There is usually close similarity in the coefficients among the members of the same
crop group, as the plant height, leaf area, ground coverage and water management are
normally similar.

25
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

The coefficients in Table 3.2 integrate the effects of both transpiration and evaporation over
time. The effects of the integration over time represent an average wetting frequency for a
'standard' crop under typical growing conditions in an irrigated setting. The values for Kc
during the initial and crop development stages are subject to the effects of large variations in
wetting frequencies and therefore refinements to the value used for Kc ini should always be
made. For frequent wettings such as with high frequency sprinkler irrigation or rainfall, the
values for Kc ini may increase substantially.

TABLE 3.2. Single (time-averaged) crop coefficients, Kc, and mean maximum plant heights for non
stressed, well-managed crops in subhumid climates (RHmin  45%, u2  2 m/s) for use with
the FAO Penman-Monteith ETo.
Maximum Crop
Crop Kc ini Kc mid Kc end
Height (h) (m)
a. Small Vegetables 0.7 1.05 0.95
Broccoli 1.05 0.95 0.3
Brussel Sprouts 1.05 0.95 0.4
Cabbage 1.05 0.95 0.4
Carrots 1.05 0.95 0.3
Cauliflower 1.05 0.95 0.4
Celery 1.05 1.00 0.6
Garlic 1.00 0.70 0.3
Lettuce 1.00 0.95 0.3
Onions - dry 1.05 0.75 0.4
- green 1.00 1.00 0.3
- seed 1.05 0.80 0.5
Spinach 1.00 0.95 0.3
Radish 0.90 0.85 0.3
b. Vegetables - Solanum Family (Solanaceae) 0.6 1.15 0.80
Egg Plant 1.05 0.90 0.8
2
Sweet Peppers (bell) 1.05 0.90 0.7
Tomato 1.152 0.70-0.90 0.6
c. Vegetables – Cucumber Family (Cucurbitaceae) 0.5 1.00 0.80
Cantaloupe 0.5 0.85 0.60 0.3
Cucumber
- Fresh Market 0.6 1.002 0.75 0.3
- Machine harvest 0.5 1.00 0.90 0.3
Pumpkin, Winter Squash 1.00 0.80 0.4
Squash, Zucchini 0.95 0.75 0.3
Sweet Melons 1.05 0.75 0.4
Watermelon 0.4 1.00 0.75 0.4
d. Roots and Tubers 0.5 1.10 0.95
Beets, table 1.05 0.95 0.4
Cassava
- year 1 0.3 0.803 0.30 1.0
- year 2 0.3 1.10 0.50 1.5
Parsnip 0.5 1.05 0.95 0.4
Potato 1.15 0.754 0.6
Sweet Potato 1.15 0.65 0.4
Sugar Beet 0.35 1.20 0.705 0.5
e. Legumes (Leguminosae) 0.4 1.15 0.55

26
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Beans, green 0.5 1.052 0.90 0.4


Beans, dry and Pulses 0.4 1.152 0.35 0.4
Chick pea 1.00 0.35 0.4
Fababean (broad bean)
- Fresh 0.5 1.152 1.10 0.8
- Dry/Seed 0.5 1.152 0.30 0.8
Grabanzo 0.4 1.15 0.35 0.8
Green Gram and Cowpeas 1.05 0.60- 0.4
0.356
Groundnut (Peanut) 1.15 0.60 0.4
Lentil 1.10 0.30 0.5
Peas
- Fresh 0.5 1.152 1.10 0.5
- Dry/Seed 1.15 0.30 0.5
Soybeans 1.15 0.50 0.5-1.0
f. Perennial Vegetables (with winter dormancy and 0.5 1.00 0.80
initially bare or mulched soil)
Artichokes 0.5 1.00 0.95 0.7
Asparagus 0.5 0.957 0.30 0.2-0.8
Mint 0.60 1.15 1.10 0.6-0.8
Strawberries 0.40 0.85 0.75 0.2
g. Fibre Crops 0.35
Cotton 1.15-1.20 0.70-0.50 1.2-1.5
Flax 1.10 0.25 1.2
Sisal 8 0.4-0.7 0.4-0.7 1.5
h. Oil Crops 0.35 1.15 0.35
Castorbean (Ricinus) 1.15 0.55 0.3
Rapeseed, Canola 1.0-1.159 0.35 0.6
Safflower 1.0-1.159 0.25 0.8
Sesame 1.10 0.25 1.0
Sunflower 1.0-1.159 0.35 2.0
i. Cereals 0.3 1.15 0.4
Barley 1.15 0.25 1
Oats 1.15 0.25 1
10
Spring Wheat 1.15 0.25-0.4 1
Winter Wheat
- with frozen soils 0.4 1.15 0.25-0.410 1
- with non-frozen soils 0.7 1.15 0.25-0.410
Maize, Field (grain) (field corn) 1.20 0.60- 2
0.3511
Maize, Sweet (sweet corn) 1.15 1.0512 1.5
Millet 1.00 0.30 1.5
Sorghum
- grain 1.00-1.10 0.55 1-2
- sweet 1.20 1.05 2-4
Rice 1.05 1.20 0.90-0.60 1
j. Forages
Alfalfa Hay
- averaged cutting effects 0.40 0.9513 0.90 0.7

27
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

- individual cutting periods 0.4014 1.2014 1.1514 0.7


- for seed 0.40 0.50 0.50 0.7
Bermuda hay
- averaged cutting effects 0.55 1.0013 0.85 0.35
- Spring crop for seed 0.35 0.90 0.65 0.4
Clover hay, Berseem
- averaged cutting effects 0.40 0.9013 0.85 0.6
- individual cutting periods 0.4014 1.1514 1.1014 0.6
Rye Grass hay
- averaged cutting effects 0.95 1.05 1.00 0.3
Sudan Grass hay (annual)
- averaged cutting effects 0.50 0.9014 0.85 1.2
- individual cutting periods 0.5014 1.1514 1.1014 1.2
Grazing Pasture
- Rotated Grazing 0.40 0.85-1.05 0.85 0.15-0.30
- Extensive Grazing 0.30 0.75 0.75 0.10
Turf grass
- cool season 15 0.90 0.95 0.95 0.10
- warm season 15 0.80 0.85 0.85 0.10
k. Sugar Cane 0.40 1.25 0.75 3
l. Tropical Fruits and Trees
Banana
- 1st year 0.50 1.10 1.00 3
- 2nd year 1.00 1.20 1.10 4
Cacao 1.00 1.05 1.05 3
Coffee
- bare ground cover 0.90 0.95 0.95 2-3
- with weeds 1.05 1.10 1.10 2-3
Date Palms 0.90 0.95 0.95 8
Palm Trees 0.95 1.00 1.00 8
Pineapple 16
- bare soil 0.50 0.30 0.30 0.6-1.2
- with grass cover 0.50 0.50 0.50 0.6-1.2
Rubber Trees 0.95 1.00 1.00 10
Tea
- non-shaded 0.95 1.00 1.00 1.5
- shaded 17 1.10 1.15 1.15 2
m. Grapes and Berries
Berries (bushes) 0.30 1.05 0.50 1.5
Grapes
- Table or Raisin 0.30 0.85 0.45 2
- Wine 0.30 0.70 0.45 1.5-2
Hops 0.3 1.05 0.85 5
n. Fruit Trees
Almonds, no ground cover 0.40 0.90 0.6518 5
19
Apples, Cherries, Pears
- no ground cover, killing frost 0.45 0.95 0.7018 4
- no ground cover, no frosts 0.60 0.95 0.7518 4
- active ground cover, killing 0.50 1.20 0.9518 4

28
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

frost
- active ground cover, no frosts 0.80 1.20 0.8518 4
Apricots, Peaches, Stone Fruit 19, 20
- no ground cover, killing frost 0.45 0.90 0.6518 3
- no ground cover, no frosts 0.55 0.90 0.6518 3
- active ground cover, killing 0.50 1.15 0.9018 3
frost
- active ground cover, no frosts 0.80 1.15 0.8518 3
Avocado, no ground cover 0.60 0.85 0.75 3
21
Citrus, no ground cover
- 70% canopy 0.70 0.65 0.70 4
- 50% canopy 0.65 0.60 0.65 3
- 20% canopy 0.50 0.45 0.55 2
Citrus, with active ground cover or weeds 22
- 70% canopy 0.75 0.70 0.75 4
- 50% canopy 0.80 0.80 0.80 3
- 20% canopy 0.85 0.85 0.85 2
Conifer Trees 23 1.00 1.00 1.00 10
Kiwi 0.40 1.05 1.05 3
24
Olives (40 to 60% ground coverage by canopy) 0.65 0.70 0.70 3-5
Pistachios, no ground cover 0.40 1.10 0.45 3-5
19
Walnut Orchard 0.50 1.10 0.6518 4-5
o. Wetlands - temperate climate
Short Veg., no frost 1.05 1.10 1.10 0.3
Reed Swamp, standing water 1.00 1.20 1.00 1-3
Reed Swamp, moist soil 0.90 1.20 0.70 1-3
p. Special
Open Water, < 2 m depth or in subhumid climates or 1.05 1.05
tropics
Open Water, > 5 m depth, clear of turbidity, temperate 0.6525 1.2525
climate
1
These are general values for Kc ini under typical irrigation management and soil wetting. For frequent wettings such as with high
frequency sprinkle irrigation or daily rainfall, these values may increase substantially and may approach 1.0 to 1.2.
2
Beans, Peas, Legumes, Tomatoes, Peppers and Cucumbers are sometimes grown on stalks reaching 1.5 to 2 meters in height. In
such cases, increased Kc values need to be taken. For green beans, peppers and cucumbers, 1.15 can be taken, and for tomatoes,
dry beans and peas, 1.20. Under these conditions h should be increased also.
3
The midseason values for cassava assume non-stressed conditions during or following the rainy season. The Kc end values
account for dormancy during the dry season.
4
The Kc end value for potatoes is about 0.40 for long season potatoes with vine kill.
5
This Kc end value is for no irrigation during the last month of the growing season. The K c end value for sugar beets is higher, up to
1.0, when irrigation or significant rain occurs during the last month.
6
The first Kc end is for harvested fresh. The second value is for harvested dry.
7
The Kc for asparagus usually remains at Kc ini during harvest of the spears, due to sparse ground cover. The Kc mid value is for
following regrowth of plant vegetation following termination of harvest of spears.
8
Kc for sisal depends on the planting density and water management (e.g., intentional moisture stress).
9
The lower values are for rainfed crops having less dense plant populations.
10
The higher value is for hand-harvested crops.

29
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

11
The first Kc end value is for harvest at high grain moisture. The second Kc end value is for harvest after complete field drying of
the grain (to about 18% moisture, wet mass basis).
12
If harvested fresh for human consumption. Use Kc end for field maize if the sweet maize is allowed to mature and dry in the
field.
13
This Kc mid coefficient for hay crops is an overall average Kc mid coefficient that averages Kc for both before and following
cuttings. It is applied to the period following the first development period until the beginning of the last late season period of the
growing season.
14
These Kc coefficients for hay crops represent immediately following cutting; at full cover; and immediately before cutting,
respectively. The growing season is described as a series of individual cutting periods (Figure 35).
15
Cool season grass varieties include dense stands of bluegrass, ryegrass, and fescue. Warm season varieties include bermuda
grass and St. Augustine grass. The 0.95 values for cool season grass represent a 0.06 to 0.08 m mowing height under general turf
conditions. Where careful water management is practiced and rapid growth is not required, Kc's for turf can be reduced by 0.10.
16
The pineapple plant has very low transpiration because it closes its stomates during the day and opens them during the night.
Therefore, the majority of ETc from pineapple is evaporation from the soil. The Kc mid < Kc ini since Kc mid occurs during full
ground cover so that soil evaporation is less. Values given assume that 50% of the ground surface is covered by black plastic
mulch and that irrigation is by sprinkler. For drip irrigation beneath the plastic mulch, Kc's given can be reduced by 0.10.
17
Includes the water requirements of the shade trees.

18
These Kc end values represent Kc prior to leaf drop. After leaf drop, Kc end  0.20 for bare, dry soil or dead ground cover and Kc
end  0.50 to 0.80 for actively growing ground cover.
19
Refer to footnotes 21 and 22 for estimating Kc for immature stands.
20
Stone fruit category applies to peaches, apricots, pears, plums and pecans.
21
These Kc values can be calculated from Eq. 98 for Kc min = 0.15 and Kc full = 0.75, 0.70 and 0.75 for the initial, mid season and
end of season periods, and fc eff = fc where fc = fraction of ground covered by tree canopy (e.g., the sun is presumed to be directly
overhead). The values listed correspond with those in Doorenbos and Pruitt (1977) and with more recent measurements. The
midseason value is lower than initial and ending values due to the effects of stomatal closure during periods of peak ET. For
humid and subhumid climates where there is less stomatal control by citrus, values for Kc ini, Kc mid, and Kc end can be increased by
0.1 - 0.2, following Rogers et al. (1983).
22
These Kc values can be calculated as Kc = fc Kc ngc + (1 - fc) Kc cover where Kc ngc is the Kc of citrus with no active ground cover
(calculated as in footnote 21), Kc cover is the Kc, for the active ground cover (0.95), and fc is defined in footnote 21. The values
listed correspond with those in Doorenbos and Pruitt (1977) and with more recent measurements. Alternatively, K c for citrus with
active ground cover can be estimated directly from Eq. 98 by setting Kc min = Kc cover. For humid and subhumid climates where
there is less stomatal control by citrus, values for Kc ini, Kc mid, and Kc end can be increased by 0.1 - 0.2, following Rogers et al.
(1983).

For non-active or only moderately active ground cover (active indicates green and growing ground cover with LAI > about 2 to
3), Kc should be weighted between Kc for no ground cover and Kc for active ground cover, with the weighting based on the
"greenness" and approximate leaf area of the ground cover.
23
Confers exhibit substantial stomatal control due to reduced aerodynamic resistance. The Kc, can easily reduce below the values
presented, which represent well-watered conditions for large forests.
24
These coefficients represent about 40 to 60% ground cover. Refer to Eq. 98 and footnotes 21 and 22 for estimating K c for
immature stands. In Spain, Pastor and Orgaz (1994) have found the following monthly K c's for olive orchards having 60%
ground cover: 0.50, 0.50, 0.65, 0.60, 0.55, 0.50, 0.45, 0.45, 0.55, 0.60, 0.65, 0.50 for months January through December. These
coefficients can be invoked by using Kc ini = 0.65, Kc mid = 0.45, and Kc end = 0.65, with stage lengths = 30, 90, 60 and 90 days,
respectively for initial, development, midseason and late season periods, and using Kc during the winter ("off season") in
December to February = 0.50.
25
These Kc's are for deep water in temperate latitudes where large temperature changes in the water body occur during the year,
and initial and peak period evaporation is low as radiation energy is absorbed into the deep water body. During fall and winter
periods (Kc end), heat is released from the water body that increases the evaporation above that for grass. Therefore, Kc mid
corresponds to the period when the water body is gaining thermal energy and Kc end when releasing thermal energy. These Kc's
should be used with caution.

Primary sources: Kc ini: Doorenbos and Kassam (1979)


Kc mid and Kc end: Doorenbos and Pruitt (1977); Pruitt (1986); Wright (1981, 1982). Snyder et al., (1989)

30
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

The values for Kc mid and Kc end in Table 3.2 represent those for a sub-humid climate with an
average daytime minimum relative humidity (RHmin) of about 45% and with calm to
moderate wind speeds averaging 2 m/s. For more humid or arid conditions, or for more or
less windy conditions, the Kc coefficients for the mid-season and end of late season stage
should be modified as described in FAO 56 Chapter 6.

The values for Kc in Table 3.2 are values for non-stressed crops cultivated under excellent
agronomic and water management conditions and achieving maximum crop yield (standard
conditions). Where stand density, height or leaf area are less than that attained under such
conditions, the value for Kc mid and, for most crops, for Kc end will need to be modified (FAO
56, Chapters 8, 9 and 10).

3.3 Construction of the Kc curve

3.3.1 Annual crops


Only three point values for Kc are required to describe and to construct the Kc curve. The
curve is constructed using the following three steps:
1. Divide the growing period into four general growth stages that describe crop
phenology or development (initial, crop development, mid-season, and late season
stage), determine the lengths of the growth stages, and identify the three Kc values
that correspond to Kc ini, Kc mid and Kc end from Table 3.2.
2. Adjust the Kc values to the frequency of wetting and/or climatic conditions of the
growth stages as outlined in the previous section.
3. Construct a curve by connecting straight line segments through each of the four
growth stages. Horizontal lines are drawn through Kc ini in the initial stage and
through Kc mid in the mid-season stage. Diagonal lines are drawn from Kc ini to Kc
mid within the course of the crop development stage and from Kc mid to Kc end within
the course of the late season stage.

3.3.2 Forage crops


Many crops grown for forage or hay are harvested several times during the growing season.
Each harvest essentially terminates a 'sub' growing season and associated Kc curve and
initiates a new 'sub' growing season and associated Kc curve. The resulting Kc curve for the
entire growing season is the aggregation of a series of Kc curves associated with each sub-
cycle.

3.3.3 Fruit trees


Values for the crop coefficient during the mid-season and end of late season stages are given
in Table 3.2. As mentioned before, the Kc values listed are typical values for standard
climatic conditions and need to be adjusted where RHmin or u2 differ. As the mid and late
season stages of deciduous trees are quite long, the specific adjustment of Kc to RHmin and u2
should take into account the varying climatic conditions throughout the season. Therefore,
several adjustments of Kc are often required if the mid and late seasons cover several climatic
seasons, e.g., spring, summer and autumn or wet and dry seasons. The Kc ini and Kc end for
evergreen non dormant trees and shrubs are often not different, where climatic conditions do
not vary much, as happens in tropical climates. Under these conditions, seasonal adjustments

31
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

for climate may therefore not be required since variations in ETc depend mostly on variations
in ETo.

3.4 Calculating ETc


From the crop coefficient curve the Kc value for any period during the growing period can be
graphically or numerically determined. Once the Kc values have been derived, the crop
evapotranspiration, ETc, can be calculated by multiplying the Kc values by the corresponding
ETo values.

3.4.1 Graphical determination of Kc


Weekly, ten-day or monthly values for Kc are necessary when ETc calculations are made on
weekly, ten-day or monthly time steps. A general procedure is to construct the Kc curve,
overlay the curve with the lengths of the weeks, decade or months, and to derive graphically
from the curve the Kc value for the period under consideration. Assuming that all decades
have a duration of 10 days facilitates the derivation of Kc and introduces little error into the
calculation of ETc.

Example 3.1
An example application for using the Kc procedure is presented for dry bean crop planted on
23rd May. The development, mid-season and late season stage lengths are given as 25, 25, 30
and 20 days and the values for Kc ini, Kc mid and Kc end are given as 0.15, 1.19, and 0.35. The
Kc curve and the 10-day values for Kc and ETc for the dry bean crop are presented in Figure
3.2.

For all decades the Kc values can be derived directly from the curve. The value at the middle
of the decade is considered to be the average Kc of that 10-day period. Only the second
decade of June, where the Kc value changes abruptly, requires some calculation.

First five days of that decade, Kc = 0.15, while during the second part of the decade Kc varies
from 0.15 to 0.36 at the end of day 10. The Kc for that decade is consequently: 5/10 (0.15) +
5/10(0.15+0.36)/2 = 0.20.

Figure 3.2: Kc curve and ten-day values for Kc and ETc for the dry bean crop example

32
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

3.4.2 Numerical determination of Kc


The Kc coefficient for any period of the growing season can be derived by considering that
during the initial and mid-season stages Kc is constant and equal to the Kc value of the growth
stage under consideration. During the crop development and late season stage, Kc varies
linearly between the Kc at the end of the previous stage (Kc prev) and the Kc at the beginning of
the next stage (Kc next), which is Kc end in the case of the late season stage:
 i    L prev 
Kc i
 Kc pre
    Kc next
 Kc prev
 (3.2)
 L stage 

where
i day number within the growing season [1.. length of the growing season],
Kci crop coefficient on day i,
Lstage length of the stage under consideration [days],
 (Lprev) sum of the lengths of all previous stages [days].

Equation 3.2 applies to all four stages.

Example 3.2. Numerical determination of Kc


Determine Kc at day 20, 40, 70 and 95 for the dry bean crop
Crop growth stage Length (days) Kc
Initial 25 Kc ini = 0.15
Crop development 25 0.15... 1.19
mid-season 30 Kc mid = 1.19
late season 20 1.19 .. Kc end = 0.35
At i = 20: initial stage, Kc = Kc ini = 0.15 -
At i = 40 Crop development stage,
For:  (Lprev) = Lini = 25 days
and: Lstage = Ldev = 25 Days
From Eq. 3.2: Kc = 0.15 + [(40 - 25)/25](1.19 - 0.15) = 0.77 -
At i = 70: mid-season stage, Kc = Kc mid = 1.19 -
At i = 95 late season stage,
For:  (Lprev) = Lini + Ldev + Lmid = (25 + 25 + 30) = 80 Days
and: Lstage = Llate = 20 Days
From Eq. 3.2: Kc = 1.19 + [(95-80)/20](0.35-1.19) = 0.56 -
The crop coefficients at day 20, 40, 70 and 95 for the dry bean crop are 0.15, 0.77, 1.19 and 0.56
respectively.

33
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 4 - Rainfall
Precipitation is water originally contained by the air and reaching the earth surface in the
form of rain, drizzle, snow, hail fog condensation, or frost. In areas with irrigated agriculture,
rainfall is the main component of precipitation.

Rainfall distribution is highly variable in both space and time. Over large areas covering
countries and continents the times of year when rainfall is received and the amount received
varies. This forms the basis of classification of climates over the global ranging from deserts
through semi-arid to humid climate (Table 4.1).

Table 4.1: Major climatic zones


Climate Annual Rainfall (mm) Wet period (months)
Desert < 100 0-1
Arid 100-400 1-3
Semi-arid 400-600 3-4
Sub-humid 600-1200 4-6
Moist sub-humid 1200-1500 6-9
Humid > 1500 9-12

Irrigation is needed when the amount of precipitation going into the root zone is not enough
to meet the water requirements of crops. The ‘wet period’ in Table 4.1 shows the months of
the year when irrigation will not be needed in each of the climatic zones.

The total rainfall received in a given time period at a location varies from year to year and
therefore, rather than using mean rainfall data a dependable level of rainfall should be
selected. For instance, the dependable rainfall may be the rainfall, which can be expected in 8
out of 10 years (80%). In order to estimate such dependable rainfall levels, a statistical
analysis needs to be made from long-term rainfall records. Dependable rainfall can only be
determined through calculation of the rainfall probability or frequency. The frequency
distribution provides the most complete description of the nature of the variation in rainfall
records. Methods of computing rainfall probabilities are not given in this course but the
student is requested to consult a statistical hand book for hydrology.

For preliminary planning purposes, monthly rainfall data can be used. When designing
irrigation systems and irrigation schedules however, monthly intervals are too long, as they
may hide dry spells of one or two weeks, which can be highly critical for the development
and yield of some crops. Decade (10-day) or weekly rainfall data should therefore be used.

4.1 Definitions

4.1.1 Historical or actual rainfall data


The historical rainfall data is the actual recorded rainfall during a specified period. Real time
irrigation scheduling requires this data in order to determine the timing of the next irrigation.
The data is also useful for evaluating purposes.

4.1.2 Average or normal rainfall


Average or normal rainfall is the arithmetic mean derived from a record of several years of
historical rainfall data. Although the monthly or annual average rainfall is commonly

34
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

available, the average rainfall appears only as information to indicate the ‘normal’ amount of
rainfall one can expect. When compared with actual amount of rainfall, it gives an idea of the
departure of the actual amount from the normal value.

4.1.3 Dependable rainfall


Dependable rainfall is defined as the rainfall, which can be, expected a set number of years
out of a total number of years. For instance, the dependable rainfall may be the rainfall,
which can be expected in 9 years out of each 10 years (90%). The percentage (90%) gives the
probability, that the rainfall will be obtained or exceeded i.e. the probability that the actual
rainfall will be equal to or higher than the dependable rainfall. One year out of 10, the rainfall
amount will be smaller.

The determination of the probability level is related to the risk, one wants to accept. In the
case of expensive structures such as bridges or dams and intakes in rivers one may want to
restrict the risk, that the rainfall (causing the flood discharge) will exceed a certain value, to
once in 50 or once in 100 years. The corresponding probabilities of exceedance here are 2%
and 1% respectively.

For agriculture the risks involved are the reduction in or the loss of the yield once in so many
years. The selected dependable level of rainfall is the depth of rainfall that can be expected 3
out of 4 years or 4 out of 5 years. The probabilities of exceedance are respectively 75% and
80%. This ‘minimum’ rainfall is used as a design norm for the dimensioning of irrigation
system as well as for water management. A higher level of dependable rainfall (say 9 out of
10 years) may need to be selected during the period that crops are germinating or are most
sensitive to water stress and yields are severely affected.

4.1.4 Dry, normal and wet conditions


For management and planning purposes the information on the amount of rainfall which one
can expect in a specified period under dry, normal and wet conditions is important. The
period under consideration may be any period such as a week, a decade, month or year.

A period is dry if the rainfall received during that period will be exceeded 4 out of 5 years,
i.e. having a probability of exceedance of 80%. This information is of importance to ascertain
if rainfall is sufficient to support rainfed agriculture or if supplementary irrigation is needed.
When calculating the capacity of storage reservoirs and of main canals, the rainfall amount
with a 80% probability of exceedance is generally used.

The rainfall in a period is normal, if the rainfall received during that period will be exceeded
1 out of 2 years, i.e. each other year. The probability of exceedance is equal to 50% and the
rainfall corresponds to the average or mean rainfall.

A period is wet if the rainfall received during that period is only exceeded 1 out of 5 years,
i.e. having a probability of exceedance of 20%.

The three values are useful for the programming of irrigation supply and simulation of
irrigation management conditions. The rainfall amount which one can expect with 20, 50 and
80% probability in each of the decades of rainy period of Moi University Main Campus are
given as an example in Figure 4.1.

35
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Figure 4.1 Mean monthly reference evapotranspiration and dependable rainfall amounts for
each of the months of the rainy period in Moi University Main Campus with a 20,
50 and 80% probability of exceedance representing a wet, normal and dry month.

4.1.5 Effective rainfall


The effective rainfall is defined as that part of rainfall which is effectively used by the crop
after rainfall losses due to surface run off and deep percolation have been accounted for. The
effective rainfall is the rainfall ultimately used to determine the irrigation requirement of the
concerned crop.

4.2 Frequency analysis

Dependable rainfall can only be determined through calculation of the rainfall probability or
frequency. This is normally done through the use of statistical methods. The problem with
rainfall however is that it seldom satisfies the statistical theories on which these methods are
based. For instance (extremely) heavy rainfalls and (extremely) low rainfalls may have
different statistical probabilities than ‘normal’ rainfalls. Also, in mountainous areas local
climatic differences are great and data may show large departure from the mean. Most
methods give similar results near the middle (around 50%) but different results for low and
high values (extremes).

Another problem arises from the fact that the frequency distribution obtained is considered
representative for frequencies in the future and is therefore being used to predict the
probability of certain rainfall. However, the frequency distribution has been obtained from
series of rainfall data over limited period and this time period does not contain all (present
and future) rainfall depths.

36
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Now, if all rainfall depths form the population, the frequency distribution based on the
sample data available forms only an estimate of the probability distribution belonging to the
population. The longer the period of observation the more similar the frequency distribution
will be to the probability distribution (assuming that the climate will not change in the
future). A period of 30 years and over normally is thought to be satisfactory.
The procedure of frequency analysis based on depth ranking and assuming a normal
distribution is as follows:
- Rank the (n) data (Pi) in a descending order, the highest value first and the lowest value
last.
- Attach a serial rank number (r) to each value (Pi) with r =1 for the highest value (P1)
and r = n for the lowest value (Pn).
- Calculate the frequency of exceedance F(P>Pi) as:

Method Frequency of exceedance


California r/n
Hazen (r-0.5)/n
Weibull r/(n+1)
Gringorten (r-0.44)/(n+0.12)

The frequency of exceedance corresponds with the plotting position on the probability scale
of the probability paper. The Weibull and Gringorten plotting positions are theoretically
better sound. All four relationships give similar values near the centre of distribution, but may
vary considerable at the extreme ends.

- Plot the data on normal probability paper

After ranking the data and calculating the frequency of exceedance, the calculated
frequencies of exceedance are plotted on normal probability paper. If the plotted data fall in a
reasonable alignment, it can be assumed that the data can be approximated by the assumed
normal distribution. Since it would be rare for a set of data to plot exactly on a line, a
decision must be made as to whether or not the deviations from the line are random
deviations or represent true deviations indicating the data does not follow the given
probability distribution.

After fitting a straight line through the points, the magnitude of rainfall corresponding to
various probabilities is derived from the probability plot.

If the points in the probability plot do not fall in a reasonable alignment, it often means that
the data is not distributed as the selected normal distribution. To convert the distribution to
normality, transformation can be attempted by plotting on the normal probability paper, the
square root or logarithm of the same rainfall data. The transformed data might be closer to the
normal distribution.

37
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

4.3 Effective rainfall

4.3.1 Infiltration and soil water storage


Effective rainfall in its simplest sense means “useful” or “utilisable” rainfall. Rainfall is not
necessarily useful or desirable at the time, rate or amount in which it is received. Some of it
may be wasted while some may even be destructive.

Rainfall may evaporate in the atmosphere, be intercepted by vegetation or it may reach the
soil, where it either runs off over the surface or infiltrates into the soil. Infiltrated rainfall may
join the ground water and be lost as groundwater flow. Thus, a part of the rainfall will be lost
through surface run-off and percolation below the root zone and will not be available for
evapotranspiration.

Run-off will depend on the intensity and duration of rainfall, on its timing (related to earlier
rainfall), on soil characteristics (texture, structure), on the slope of the terrain and on the
vegetation cover.

Deep percolation will depend on initial soil moisture content, soil water holding capacity (soil
type) and depth of the rooting zone (crop type).

Effective rainfall may be defined as the fraction of rainfall, stored in the root zone and used
by plant-soil system for evapotranspiration.

This is purely technical definition: for a scheme manager rainfall will only be effective when
one or more irrigation gifts can be omitted. Also, rainfall, which falls on fallow land, will
only be considered effective in as far as it will be stored in the soil for use by the next crop.

For a farmer, rainfall which leaves his field as run-off is not effective, although it may be
effective to his neighbour receiving this runoff or to farmers further away who may re-use the
water for their irrigation scheme by deriving it from the river.

Furthermore, the decrease in radiation and temperature and thus in evapotranspiration equally
might be considered a useful effect of rainfall.

Thus, it is always important to consider in which setting and from whose point of view the
effectiveness of rainfall is considered. For agricultural purposes rainfall may be considered
effective in as far as it satisfies water needs for land preparation, consumption by crops, salt
leaching requirements and/or percolation needs (as for rice or fishponds).

4.3.2 Estimation procedure


Although run-off and deep percolation can be measured it is not practical to do this for all
meteorological stations or scheme areas as these measures are complicated, time consuming
and expensive.

A number of empirical formulae have been developed for the estimation of effective rainfall.
Most of these formulas take into account total precipitation and consumptive use or
(potential) evapotranspiration.

38
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Fixed percentage of rainfall


The effective rainfall (Peff) is calculated as a fixed percentage (a) of total rainfall (Ptot)

Peff = a.Ptot for a = 0.7 to 0.9

This method is rather crude and results will not be very satisfactory, as -among others- the
influence of the depth of rainfall is disregarded.

USDA SCS method


The Soil Conservation Service of the U.S. Department of Agriculture (USDA SCS) has
developed a procedure for estimating effective rainfall as a function of the
evapotranspiration/precipitation ratio by processing precipitation and soil moisture data for a
period of 50 years for 22 experimental stations representing different climatic and soil
moisture data. The results are given in Table 4.2.
Table 4.2 Average monthly effective rainfall as related to Average Monthly ETcrop and Mean Monthly Rainfall
(USDA SCS, 1969)
Monthly Monthly mean rainfall (mm)
ETcrop 12.5 25 37.5 50 62.5 75 87.5 100 112 125 138 150 162 175 188 200
(mm) Effective rainfall (mm)
25 8 16 24
50 8 17 25 32 39 46
75 9 18 27 34 41 48 56 62 69
100 9 19 28 35 43 52 59 66 73 80 87 94 100
125 10 20 30 37 46 54 62 70 76 85 92 98 107 116 120
150 10 21 31 39 49 57 66 74 81 89 97 104 112 119 127 133
175 11 23 32 42 52 61 69 78 86 95 103 111 118 126 134 141
200 11 24 33 44 54 64 73 82 91 100 109 117 125 134 142 150
225 12 25 35 47 57 68 78 87 96 106 115 124 132 141 150 159
250 13 25 38 50 61 72 84 92 102 112 121 132 140 150 158 167

The average monthly effective rainfall is based on a net water storage depth of 75 mm.
Correction factors to be used for different net water depth storage are:

Effective storage 20 25 37.5 50 62.5 75 100 125 150 175 200


Correction factor 0.73 0.77 0.86 0.93 0.97 1.00 1.02 1.04 1.06 1.07 1.08

4.4 Assignment 2
Given over 30 years rainfall data for a station. Task:
1. Graphical frequency analysis of annual rainfall data using Microsoft Excel;
2. Given that rainfall less than 750 mm results in a drought event, determine the return
period of drought;
3. Perform frequency analysis at decadal time step using RAINBOW Software;
4. Determine the decadal amount of rainfall that represents 80% dependable level for
design of an irrigation system;
5. Determine the decadal amount of rainfall that represents 20% and 50% dependable
levels for planning and management of an irrigation system; and
6. Determine graphical relationship of mean decadal reference evapotranspiration and
dependable rainfall amounts for each of the decades of the rainy period in the station
with a 20, 50 and 80% probability of exceedance representing a wet, normal and dry
month.

39
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 5 - Irrigation requirement


This chapter provides guidelines for estimation of the irrigation water requirement. The
requirement is the amount of water to be supplied to the plants to prevent stress and yield
reduction. Seasonal and short-term irrigation requirements for fields, farms and the entire
irrigation scheme are needed for the design and operation of an irrigation system. Knowledge
of the peak irrigation requirement is needed in determining the size of main canals, pipe
networks and pumping stations. Information on the seasonal irrigation requirement is needed
when designing water storage reservoirs and distribution systems or in determining the size
of the irrigation scheme given a limited amount of water. Short term irrigation requirements,
when combined with soil water holding characteristics, enables to specify when and how
much water to apply.

The calculation of the irrigation water requirements consists in determining, in a first step, the
net irrigation requirement of each of the crops cultivated. The requirement is obtained by
subtracting the expected gains of water from the crop water requirement (ETcrop). The first
gain to be considered is rainfall, but water transported to the root zone by capillary rise from
shallow ground water table may also contribute to the crop requirement. Part of the crop
water requirement may also be met by stored soil water at the start of the growing season.

In addition to meeting the net irrigation requirements, water may be required for leaching
accumulated salts out of the root zone and for cultural practices. Since irrigation is never 100
% efficient, allowance must be made for losses during conveyance and application of water.
The gross irrigation requirement of the scheme is finally obtained by adding up the individual
irrigation requirements of each of the crops.

5.1 Definitions

5.1.1 Crop irrigation requirement


It is the depth of water needed to meet the water loss through evapotranspiration of a disease-
free crop, growing in large fields under non-restricting soil conditions including soil water
and fertility and achieving full production potential under the given environment.

The net irrigation requirement (In) is obtained by subtracting from the crop water requirement
(ETcrop) the expected gains of water. The gains include effective rainfall (P eff), ground water
contribution (Ge) and stored soil water (Wb). To avoid any water deficit during the season, the
estimations are made with rainfall amounts that can be expected 3 out of 4 years or 4 out of 5
years.

In the calculations of the gross irrigation requirement, the amount of water needed for
leaching accumulated salts from the root zone and to compensate for water losses during
conveyance and application are included. The leaching requirement (LR) and irrigation
efficiency (E) are expressed as a fraction of the net irrigation requirement.

5.1.2 Scheme irrigation requirement


The scheme irrigation requirement is the sum of the individual irrigation requirements of
each of the crops.

40
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

5.2 First assessment of net irrigation requirement

5.2.1 Seasonal net irrigation requirement


For preliminary planning purpose, monthly data are normally used to determine the seasonal
irrigation requirement. The water requirements for the different months are computed by
taking into account on the one hand the mean crop water requirement (ETcrop) and on the
other hand the dependable effective rainfall (Peff). No other gain of water is considered.

Mean monthly evapotranspiration values are generally used. By nature evapotranspiration is


much more conservative for a given period in the year and at a given location than
precipitation, which usually exhibits large variations. Hence, rather than using mean monthly
rainfall data, a dependable level of rainfall should be selected e.g. 75 or 80% i.e. the monthly
rainfall that can be expected 3 out of 4 years or 4 out of 5 years. A higher level of dependable
rainfall (say 9 out of 10 years, i.e. 90%) may need to be selected during the period that crops
are germinating or are most sensitive to water stress and yields are severely affected. An
example of the calculation of seasonal net irrigation requirement is given in Table 5.1.

Table 5.1: First estimate of seasonal net irrigation requirement of maize


Crop: Maize, Length of growing season: 150 days, Sowing date: 1 March, Soil: Sandy loam
March April May June July Total
31 days 30 days 31 days 30 days 28 days
ETo (mm/day) 2.4 3.7 4.6 4.9 5.5 -
Kc 0.4 0.68 1.15 1.15 0.88 -
ETcrop (mm/day) 1.0 2.5 5.3 5.6 4.8 -
ETcrop (mm) 30 75 164 169 136 574
Rainfall (mm) 14 15 0 0 0 -
Peff (mm) 8 10 0 0 0 18
Inet (mm) 22 65 164 169 (peak) 136
Seasonal Net irrigation requirement (mm) 556

5.2.2 Peak irrigation requirement


For preliminary planning, the capacity of the engineering works can be obtained from supply
needed during the month of peak water use. In selecting ETcrop in the months of peak water
use, knowledge should be obtained on level and frequency at which high demands for water
can be expected. When sufficiently long climatic records are available (10 years or more) a
frequency analysis can be made similar to that for rainfall. The value of ETcrop selected for
design can then be based on a probability of 75 or 80 percent or highest ETcrop value out of 4
or 5 years.

5.2.3 Ground water contribution (Ge)

Ground water, if not too far below the root zone, may add water to the root zone by capillary
rise. The extent of capillary rise depends on:
- the depth of ground water table below the root zone,
- the soil type, i.e. its capillary properties. Capillary rise will be much higher in clay(ey)
soils than in sand(y) soils,
- the soil water content in the root zone i.e. the difference in water content at the ground
water table (saturation) and in the root zone. The larger the difference, the larger the
upward transport of water.

41
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

5.2.4 Soil water contribution (Wb)

Abundant rainfall during the wet season may bring the soil profile near or at field capacity at
the start of the growing season. The amount of water may be equivalent to one full irrigation
and hence should be deducted when determining the seasonal irrigation requirements. An
example of the calculation of the net irrigation requirement taking into account the soil water
contribution is worked out in Table 5.2.

Table 5.2: calculation of seasonal net irrigation requirement of maize, considering available
soil water at the start of the season.
Available soil water at the start of season 25 mm
ETcrop 574 mm
Effective rainfall -18 mm
Ground water contribution -31 mm
Soil water contribution -25 mm
Net irrigation requirement 500 mm or 5,000m³/ha

If no water is stored in the root zone, it should be checked if no extra water is required to
replenish depleted soil moisture prior to sowing/planting. Soil moisture depleted by the
preceding crop may not have been replenished and should preferably be brought to an
acceptable level, for instance up to field capacity level. As this “filling up” prior to
sowing/planting is not part of irrigation schedule, it constitutes an extra water requirement.

5.2.5 Salinity control

Accumulation of salts in the root zone is referred to as soil salinity. Salts contained in
irrigation water are left behind when water is taken up by plants or lost by evaporation. As a
result there is gradual accumulation of these salts as the season progresses and from one year
to the next year.

The level of salinity is affected by the quality (salt content) and quality of irrigation water, by
soil factors affecting drainage, by the availability of water (rainfall) to leach the profile, by
the method of irrigation and by the prevailing cultural practices.

The leaching requirement is the minimum amount of irrigation water that must percolate
below the root zone in order to maintain soil salinity at a given level. The level maintained
usually corresponds to the salinity level at which the particular crop would not suffer an
unacceptable reduction in yield level. The leaching requirement may be calculated by:
EC 1
LR 
w

5 EC e
 EC w
El
where,
LR = leaching requirement, expressed as the fraction of the total seasonal volume of
irrigation water supplied, which should be used for the leaching of salts,
ECw = electric conductivity of irrigation water [dS/m or mmhos/cm],
ECe = electric conductivity of the soil saturation extract corresponding to the yield level, that
can be tolerated for the particular crop [dS/m or mmhos/cm]. Crop salt tolerance for
different crops are given in Table 5.3,
ECl = leaching efficiency [fraction].

42
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Table 5.3: Crop salt tolerance levels for different crops


Crop Yield potential Max
100% 90% 75% 50% ECe
Ece ECw ECe ECw ECe ECw ECe ECw
Field crops
Barley1 8.0 5.3 10.0 6.7 13.0 8.7 18.0 12.0 28
Beans (field) 1.0 0.7 1.5 1.0 2.3 1.5 3.6 2.4 7
Broad beans 1.6 1.1 2.6 1.8 4.2 2.0 6.8 4.5 12
Maize 1.7 1.1 2.5 1.7 3.8 2.5 5.9 3.9 10
Cotton 7.7 5.1 9.6 6.4 13.0 8.4 17.0 12.0 27
Cowpeas 1.3 0.9 2.0 1.3 3.1 2.1 4.9 3.2 9
Flax 1.7 1.1 2.5 1.7 3.8 2.5 5.9 3.9 10
Groundnut 3.2 2.1 3.5 2.4 4.1 2.7 4.9 3.3 7
Rice (paddy) 3.0 2.0 3.8 2.6 5.1 3.4 7.2 4.8 12
Safflower 5.3 3.5 6.2 4.1 7.6 5.0 9.9 6.6 15
Sesbania 2.3 1.5 3.7 2.5 5.9 3.9 9.4 6.3 17
Sorghum 4.0 2.7 5.1 3.4 7.2 4.8 11.0 7.2 18
Soybean 5.0 3.3 5.5 3.7 6.2 4.2 7.5 5.0 10
Sugarbeet 7.0 4.7 8.7 5.8 11.0 7.5 15.0 10.0 24
Wheat1 6.0 4.0 7.4 4.9 9.5 6.4 13.0 8.7 20
Vegetable crops
Beans 1.0 0.7 1.5 1.0 2.3 1.5 3.6 2.4 7
Beets2 4.0 2.7 5.1 3.4 6.8 4.5 9.6 6.4 15
Broccoli 2.8 1.9 3.9 2.6 5.5 3.7 8.2 5.5 14
Cabbage 1.8 1.2 2.8 1.9 4.4 2.9 7.0 4.6 12
Cantaloupe 2.2 1.5 3.6 2.4 5.7 3.8 9.1 6.1 16
Carrot 1.0 0.7 1.7 1.1 2.8 1.9 4.6 3.1 8
Cucumber 2.5 1.7 3.3 2.2 4.4 2.9 6.3 4.2 10
Lettuce 1.3 0.9 2.1 1.4 3.2 2.1 5.2 3.4 9
Onion 1.2 0.8 1.8 1.2 2.8 1.8 4.3 2.9 8
Pepper 1.5 1.0 2.2 1.5 3.3 2.2 5.1 3.4 9
Potato 1.7 1.1 2.5 1.7 3.8 2.5 5.9 3.9 10
Radish 1.2 0.8 2.0 1.3 3.1 2.1 5.0 3.4 9
Spinach 2.0 1.3 3.3 2.2 5.3 3.5 8.6 5.7 15
Sweet maize 1.7 1.1 2.5 1.7 3.8 2.5 5.9 3.9 10
Sweet potato 1.5 1.0 2.4 1.6 3.8 2.5 6.0 4.0 11
Tomato 2.5 1.7 3.5 2.3 5.0 3.4 7.6 5.0 13
Forage crops
Alfalfa 2.0 1.3 3.4 2.2 5.4 3.6 8.8 5.9 16
Barley hay1 6.0 4.0 7.4 4.9 9.5 6.3 13.0 8.7 20
Bermuda grass 6.9 4.6 8.5 5.7 10.8 7.2 14.7 9.8 23
Clover, berseem 1.5 1.0 3.2 2.1 5.9 3.9 10.3 6.8 19
Maize (forage) 1.8 1.2 3.2 2.1 5.2 3.5 8.6 5.7 16
Harding grass 4.6 3.1 5.9 3.9 7.9 5.3 11.1 7.4 18
Orchard grass 1.5 1.0 3.1 2.1 5.5 3.7 9.6 6.4 18
Perennial rye 5.6 3.7 6.9 4.6 8.9 5.9 12.2 8.1 19
Soudan grass 2.8 1.9 5.1 3.4 8.6 5.7 14.4 9.6 26
Tall fescue 3.9 2.6 5.8 3.9 8.6 5.7 13.3 8.9 23
Tall wheat grass 7.5 5.0 9.9 6.6 13.3 9.0 19.4 13.0 32
Trefoil, big 2.3 1.5 2.8 1.9 3.6 2.4 4.9 3.3 8
Trefoil, small 5.0 3.3 6.0 4.0 7.5 5.0 10.0 6.7 15
Wheat grass 7.5 5.0 9.0 6.0 11.0 7.4 15.0 9.8 22
Fruit crops
Almond 1.5 1.0 2.0 1.4 2.8 1.9 4.1 2.7 7
Apple, pear 1.7 1.0 2.3 1.6 3.3 2.2 4.8 3.2 8
Apricot 1.6 1.1 2.0 1.3 2.6 1.8 3.7 2.5 6
Avocado 1.3 0.9 1.8 1.2 2.5 1.7 3.7 2.4 6
Date palm 4.0 2.7 6.8 4.5 10.9 7.3 17.9 12.0 32
Fig, olive, pomegranate 2.7 1.8 3.8 2.6 5.5 3.7 8.4 5.6 14
Grape 1.5 1.0 2.5 1.7 4.1 2.7 6.7 4.5 12
Grapefruit 1.8 1.2 2.4 1.6 3.4 2.2 4.9 3.3 8
Lemon 1.7 1.1 2.3 1.6 3.3 2.2 4.8 3.2 8
Orange 1.7 1.1 2.3 1.6 3.2 2.2 4.8 3.2 8
Peach 1.7 1.1 2.2 1.4 2.9 1.9 4.1 2.7 7
Plum 1.5 1.0 2.1 1.4 2.9 1.9 4.3 2.8 7
Strawberry 1.0 0.7 1.3 0.9 1.8 1.2 2.5 1.7 4
Walnut 1.7 1.1 2.3 1.6 3.3 2.2 4.8 3.2 8
1
During germination and seedling stage ECe should not exceed 4 or 5 mmhos/cm. Data may not apply to new semi-dwarf
varieties of wheat. 2During germination Ece should not exceed 3 mmhos/cm.

43
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Leaching efficiency is a function of soil drainage characteristics. Well drained sandy soils
have ECl as low as 30%. Other soils fall within this range. An example of the calculations of
the leaching requirement is worked out in Table 5.4.

Table 5.4: Estimate of leaching requirement


Location: Marigat
Maize (season: April – August)
Soil : sandy loam (ECl =0.9), Seasonal net irrigation requirement : 500 mm
ECw = 1 to 2.5 dS/m
For 100% yield
If ECw = 1 dS/m LR = 1/((5×1.7-1) ×0.9) = 0.15
If ECw = 2.5 dS/m LR = 2.5/((5×1.7-2.5) ×0.9) = 0.46

Leaching of salt can be done before, during or after the irrigation season depending on when
water is available and on the salt accumulation rate. For accumulation rate creating
intolerable salt levels in the root zone during the season, leaching has to be done concurrently
with irrigation. If not, leaching can be done before or after the season, depending on the
availability of water. Leaching is normally practised outside the peak period, but when very
saline water is used, this may need to be considered in the peak supply.

In case problems with salinity may be expected FAO Irrigation and Drainage Paper No. 29
“Water Quality for Agriculture” constitutes a good first reference on salinity.

5.3 Crop irrigation requirement


In
The irrigation requirement of a crop is given by:
1  LR 
Where In = net irrigation requirement (mm/season)
LR = leaching requirement

The net irrigation requirement of the crop is obtained by subtracting the gains of water from
the crop water requirement (ETcrop):
I n  ET crop
 ( P eff  G e  W b )

The gains of water include effective rainfall (Peff), ground water contribution (Ge) and stored
soil water (Wb). All variables are expressed in units of depth of water (mm).
AI n
The amount of water to be supplied can be obtained from: (m³) 10
1  LR 
Where A= acreage under the given crop (ha), the factor 10 appears due to conversion of mm
to m³/ha.

5.4 Scheme irrigation requirement

5.4.1 Irrigation efficiencies


Water losses occur at different levels:
- at the level of the plant i.e. when applying water to the soil,
- at the level of the field i.e. after water has entered the field,
- at the level of the canals i.e. during conveyance of the water between the main scheme
inlet to the field offtake.

44
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Water losses are normally expressed as irrigation efficiencies, whereby the concept
“efficiency” denotes that fraction of total amount of water, which will benefit the field
respectively the crop.

Field application efficiency (Ea)


At the field level losses may occur due to deep percolation below the root zone and unwanted
drainage (runoff) of water from the field. Deep percolation almost certainly will occur as it is
nearly impossible to achieve uniform water distribution within a field and the correct rate of
water application at the crop level. Field application efficiency is defined as:

Ea = (water stored in the root zone)/(water received at field inlet)

Ea is affected by the type of the irrigation system, soil type and the skill of the farmer.

Field canal efficiency (Eb)


The field canal efficiency is the efficiency of water conveyance in the canals within the
sector, block or sub-unit. The unit may be the tertiary, but may also be the quaternary unit or
even the sub-quaternary unit. The exact definition of the “field” unit depends on the
definition of the scheme as well as on irrigation organisational aspects i.e. farmers groups or
irrigation groups. The canals at this level are usually unlined and seepage losses along them
are high. Field canal efficiency is defined as:

Eb = (water received at field inlet)/(water received at block inlet)

Conveyance efficiency (Ec)


The conveyance efficiency is the efficiency of water conveyance in the (main) canal system,
which transports i.e. conveys water from the scheme head works to the various sectors,
blocks or sub-units. The scheme head works may be the intake at the river or storage
reservoir, but it may also be the head of tertiary unit in case the tertiary unit is considered as
the scheme.

Depending on the length of the canals and the porosity of the bed (lined, unlined, soil type)
seepage and evaporation can be high. Conveyance efficiency is defined as:

Ec = (water received at block inlet)/(water received at the head works)

Distribution efficiency (Ed)


The distribution efficiency is the efficiency of water conveyance and distribution between the
head (inlet) of the scheme and the farm or field offtake (turn-out/inlet). It covers all the losses
inherent to the ‘transport’ of the water and is a function of layout of the system, water
transport type (canals, pipes), nature of the canal bed, soil type, maintenance, irrigation
method and scheme management. Ed is independent of crop type and crop stage (although
these factors may influence the irrigation method).

Ed = Ec×Eb

Scheme efficiency (Ep)


The overall, scheme or project efficiency can be defined as:
Ep = (water stored in the root zone)/(water received at the head works)

45
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

The scheme efficiency can be calculated as:

Ep = Ea×Eb×Ec

Irrigation efficiencies must be established through research i.e. monitoring of irrigation in the
field. Not always such studies are or have been carried out, or adjustments in locally
established values are necessary due to lining of the canals or improved performance of the
farmers due to longer irrigation practice. If no values are locally available, use could be made
of Table 5.5 for an estimation of efficiencies.

Table 5.5: Conveyance, field canal, distribution, field application efficiencies


ICID/ILRI
Conveyance (Ec)
Continuous supply with no substantial change in flow 0.9
Rotational supply in projects of 3000 to 7000 ha and rotational areas of 70 – 0.8
300 ha with effective management
Rotational supply in large schemes (>10000 ha) and small schemes (<1000
ha) with respective problematic communication and less effective
management:
-based on predetermined schedule 0.7
-based on advanced request 0.65
Field canal efficiency (Eb)
Blocks larger than 20 ha -unlined 0.8
-lined or piped 0.9
Blocks below or up to 20 ha -unlined 0.7
-lined or piped 0.8
Distribution efficiency (Ed=Ec×Eb)
Average for rotational supply with management and communication:
-adequate 0.65
-sufficient 0.55
-insufficient 0.40
-poor 0.30
Field application efficiency (Ea)
USDA US(SCS) ICID/ILRI
Surface method:
-soil type -light soils 0.55
-medium soils 0.70
-heavy soils 0.60
-irrigation method -graded border 0.6-0.75 0.53
-basin and level border 0.6-0.80 0.58
-contour ditch 0.5-0.55
-furrow 0.55-0.7 0.57
-corrugation 0.50-0.7
Surface Up to 0.80
Sprinkler -hot, dry climate 0.60
-moderate climate 0.70
-humid, cool climate 0.8 0.67
Rice 0.32

46
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

5.5 Scheme water requirement

5.5.1 Seasonal requirement


Once the cropping pattern and intensity have been selected, the gross irrigation requirement
of the scheme is obtained by adding up the individual irrigation requirements of each of the
crops:
N
10 A i I n ,i
V 
E
 1 
 LR
p i 1 i

where, Ep = scheme efficiency


for crop i = 1 to N
A = cultivated area (ha)
In = net irrigation requirement (mm/season)
LR = leaching requirement

An example of the calculation of the scheme water requirement is worked out in Table 5.6.

Table 5.6: Yearly supply requirement of a scheme


Project size: 150 ha
Cropping intensity: 200%
-short rains: wheat (Area: 150 ha, In: 82 mm, LR: 0)
-long rains: maize (Area: 90 ha, In: 500 mm, LR: 0.15)
cotton (Area: 60 ha, In: 650 mm, LR: 0.07)
Surface irrigation, rotational supply to irrigation blocks of 20 ha, lined canal (Ep = 0.50)
Crop
Maize (10/0.5)×(90×500)/(1-0.15) = 1,058,824 m³
Cotton (10/0.5)×(60×650)/(1-0.07) = 838,710 m³
Wheat (10/0.5)×(150×82) = 246,000 m³
The yearly supply requirement: V = 2,143,534 m³
Similarly the monthly supply requirements can be determined.

5.5.2 Peak supply


For a first estimate on the capacity of engineering works, the peak supply can be based on
project supply of the month of highest irrigation demands:

 A 
10
V max  C i
I peak , i
E p i 1

Where C = flexibility factor


Ipeak = net irrigation requirement during peak month

To incorporate flexibility in the delivery capacity of the supply system as well as to allow for
future intensification and diversification of crop production, a flexibility factor C is
frequently added. This factor varies with the type of project and is generally higher for small
schemes as compared to large schemes. With monocultures such as orchards and permanent
pastures the factor is small.

47
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 6- Soil-Water-Plant Relationships


The growth of the crop involves the plant itself, the water and the soil. Plant-water-soil
relationships will be discussed here in as far as the soil is concerned. Main aspects considered
are: (i) how much water can be held by (or stored in) the soil, and (ii) how much water stored
in the soil is readily available for use by the crop.

6.1 The soil

6.1.1 Soil texture


The major part of (non-organic) soils consists of mineral soil particles: sand, silt and clay.
The mineral soil elements can be classified according to their size. The size distribution of the
ultimate soil particles is referred to as “texture”. It can be estimated in the fields or
determined in the laboratory.

There are several textural classifications. The most commonly used for agronomic purposes
is the classification of the U.S. Department of Agriculture (Figure 6.1). The main particle size
limits are given in Table 6.1.

Figure 6.1: Texture classification (Soil survey manual United States Department of
Agriculture)

Table 6.1: Particle size limits


Soil Sub-class Diameter (cm) Limits (cm)
Sand 2.00-0.050
Very coarse 2.00-1.00
Coarse 1.00-0.50
Medium 0.50-0.25
Fine 0.25-0.10
Very fine 0.10-0.05
Silt 0.050-0.002
Coarse 0.050-0.020
Fine 0.020-0.002
Clay < 0.002

48
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

6.1.2 Soil structure


The soil is a complex porous medium consisting of four components: mineral solid particles,
organic matter, water with dissolved substances and air.

The solid mineral components consists mainly of various proportions of sand, silt and clay.
Organic matter is usual about 5% or less (<1% in tropical and arid soils). There is also some
living organic matter. The mineral particles and organic matter form a loosely packed matrix
with spaces in between the particles.

The term “soil structure” refers to the three-dimensional arrangements of the primary soil
particles (sand, silt, and clay) and the secondary soil particles (micro-aggregates) into a
certain structural pattern (macro-aggregates). The aggregates of textural elements are held
together by colloids (mineral and organic) and separated from one another by cracks and
large pores. Structure is not a plant-growth factor in itself, but it defines water retention,
water movement, soil aeration, root penetration, micro-biological activities, resistance to
erosion, etc.

6.1.3 Particle density


Particle density is the mass per unit volume of soil particles, usually expressed on grams per
cm3 of soil particles. Instead of “particle density” the term specific gravity is often used.
Average specific gravity of distinct soil components are: organic matter 1.47 g/cm3; sand
2.66 g/cm3; clay 2.75 g/cm3. For the soil as a whole the particle density varies from 2.6 to 2.9
g/cm3 (2.65 g/cm3 in average).

6.1.4 Bulk density


Bulk density is the dry weight of a unit volume of soil in its field condition, usually expressed
in grams per cm3. Bulk density is also known as “volume weight” or “apparent density”. The
bulk density of the soil is obtained by weighting an oven dry undisturbed soil sample of
known volume. The bulk density, especially of the topsoil, should be checked regularly. It
usually ranges from 1.25 to 1.65 g/cm3. The finer the texture of the soil and the higher the
organic matter content the smaller the bulk density.

6.2 Soil moisture

A soil consists of mineral solid particles, organic matter and pore space. In the pore space, the
space between the solid particles, water can be stored. The pore space varies from 30% to
60% of the total soil volume. Water and air exist in the pore space in varying proportions
depending on the water status in the soil. In a dry soil, most of the pore space would be empty
and filled with air, while in a wet soil, water would occupy most of the pore space.

The soil moisture content can be expressed in several ways. When examining water budgets,
it is useful to express the water content on a volume basis or in terms of depths. When
defining soil moisture content, it is useful to visualise the soil as a container in which water
can be stored. In Figure 6.2 a cubic soil sample is presented which contains a certain amount
of water in the pore space. Consider that all the solid particles could be compressed together
without leaving any pore space between them. The soil water would settle above the solid and
the soil air would occupy the space above the soil water.

49
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

a
Air
d
B b Solution

Solid
c

Figure 6.2 Schematic presentation of a cubic soil sample of base area A with the three
considered soil phases: solids, solution and air.

6.2.1 Dry mass basis


The dry mass moisture content is given by the ratio of the mass of soil water to the mass of
dry soil (times 100 if expressed as a percentage):

mass . soil . water


 m
 100 [mass %]
mass . dry . soil

The gravimetric method is the standard method to determine the soil moisture content. The
method consists in weighing, drying and reweighing a soil sample. By means of an auger a
number of representative soil samples are taken. The difference in mass of the sample before
and after drying gives the mass of water. If ms+w is the wet mass (gram) of the sample and ms
the mass (gram) of oven dry sample, the dry mass water percentage is:

m sw  m s
 m
 100 [mass %]
ms

In order to remove all the water from the soil sample, the sample should be dried to constant
weight during 24 hours in an oven at 105 °C.

6.2.2 Volume basis


The volume water content is the ratio of the volume of the soil water to the bulk volume of
the soil (times 100 if expressed as a percentage):

volums . soil . water


  100 [vol%]
bulk .volume . soil

with reference to Figure 6.2, it is the ratio of (B a) to (B A).

50
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Given the wet mass of the sample and the mass of the oven dry sample, the volume moisture
percentage is given by:
m sw  m s
  100 [vol%]
 wV b
where (ms+w-ms)/ρw is the volume of water (cm³), ρw the density of water (g/cm³) and Vb the
bulk volume of the soil sample (i.e. BA in Figure 6.2). For all practical purposes, soil water
can be considered to have a density of equal to 1 g/cm³.

Measuring the volume water content of a soil sample implies that the bulk volume of the
sample is carefully preserved during sampling. Such undisturbed samples could be obtained
with core samplers. However, if the natural soil structure can not be maintained during the
sampling, the volume water content of the loose material can still be determined if the bulk
density of the soil is known. The bulk density ρb is defined as the ratio of dry mass of given
sample to bulk volume:
mass . dry . soil ms
b   [g/cm³]
bulk .volume . soil Vb

with reference to Figure 6.2, it is the ratio between (ρb c A) to (B A).


m sw  m s
Integrating the above equations yields:   100  b   b m
[vol %]
ms w

6.2.3 Equivalent depth


The water content can be expressed also as the equivalent depth of liquid water per unit soil
depth. With reference to Figure 6.2, the equivalent depth for a soil depth of B, is b. If
expressed in millimeter water per meter soil depth, the depth value is equal to 10 times the
value of the volume water percentage:
1000 b ( mm )
S   10  [mm (water)/m (soil depth)]
B (m )

The amount of water retained in a specified soil depth (e.g. the root zone) is consequently:
W = ZS = 10Zθ [mm (water)]

Where Z is the effective rooting depth in meter of the crop and θ the volume water percentage
(vol%). To express the water content as a depth of water is useful. It makes the adding and
subtracting of gains and losses of water straightforward since the various parameters of the
soil moisture budget are usually expressed in terms of equivalent water depth.

Example 6.1
A representative soil sample is taken in the root zone (Z = 0.6 m) of potatoes cultivated on a
loamy soil (ρb=1.40). The weight of the soil sample before and after drying is respectively
133 and 114 gram. Express the water content of the root zone as a depth of water.

Solution
Θm=100(133-114)/114 = 16.7 mass % ;
Θ = 1.4(16.7) = 23.3 vol%.
S = 10(23.3) = 233.3 mm(water)/m(soil depth);
W = 0.60(233.3) = 140.0 mm. Therefore the root zone contains 140.0 mm of water.

51
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

6.3 Soil moisture contents

6.3.1 Water retention curve


Water is stored in the soil within the pore spaces between the soil particles. It is held to the
particles by forces of attraction acting between the water molecules and the particles of the
soil matrix. The closer a water molecule is to a matrix particle the stronger is the force of
attraction binding it to the matrix. Water that is further away from the matrix is free to move
and be taken up by plants or be drained out of the root zone by gravity.

The forces holding water to the soil matrix are called matrix forces. The plant roots have to
overcome these forces to extract water from the soil. Water in a completely saturated soil can
be regarded as being free of all forces i.e. has a matrix potential of zero. As water is removed
from the soil the remaining water comes under increasing stronger matrix forces and will be
held stronger by the soil.

There is therefore a relationship between soil water content and its matrix potential. This
relationship can be shown in a soil water retention curve. The shape of the soil water
retention curve is strongly affected by the texture and structure of the soil. It is normally
determined experimentally using a pressure plate apparatus. Once the curve is made the water
content of a soil sample can be estimated if its matrix potential is known.

6.3.2 Saturation water content


When all the available pore space is filled with water, the soil is said to be saturated.
Saturated conditions exist in the upper soil layers immediately after a heavy rain or irrigation,
and also directly above the groundwater table. Since at saturation water has filled up all the
pore space, the saturation water content is equal to the porosity:
 b 
  100  1   [vol%]
sat
 p 
 
θsat = soil water content at saturation point (vol%)
ρb = soil bulk density (g/cm³)
ρp = soil particle density, which is the mass per unit volume of soil particles. For the
soil as a whole the particle density is about 2.65 g/cm³.

6.3.3 Field capacity


If the soil is saturated, the water that is not very strongly held to the soil matrix drains freely
under gravity. The water content at field capacity (θFC) is that soil water content, which is
held by the soil matrix against the gravitational forces. Field capacity will be reached within 1
to 2 days after saturation.

Example 6.2
The root zone of potatoes (Z = 0.6 m), cultivated on a loamy soil, contains 145.0 mm of
water. What will be its water content, reached within 1 to 2 days, after a heavy rainfall of 60
mm. The field capacity of the loamy soil is 31 vol%.

Solution
The root zone contains 145.0 mm of water
Rainfall +60.0
Total 205.0 mm of water
Field capacity is 31 vol%.

52
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

The corresponding equivalent depth is SFC = 10(31) = 310 mm (water)/m(soil depth). The
amount of water retained in the root zone at field capacity is WFC=0.60(310) =186 mm.
Hence, the root zone cannot contain more than 186 mm. The surplus (205-186 = 19 mm) will
drain freely out of the root zone under gravity. After 1 to 2 days, the root zone will contain
186 mm of water.

6.3.4 Permanent wilting point


Plants extract water freely and easily from soil at field capacity. As water uptake progresses
the amount of water in the soil decreases and the remaining water is held to the particles with
greater force, making it more difficult for the plant to extract it. A point is reached when the
plant can no longer extract water in sufficient quantities and rapidly enough to replace the
water being lost by transpiration. Leaves begin to wilt during the afternoon but may still
recover at night. Finally a stage is reached when the wilted plants do not recover at night, or
even when water is added to the soil. The plants have permanently wilted and soil water
content is said to be at the permanent wilting point (θwp). Beyond the permanent wilting point
water is no longer available to the plant.

6.3.5 Total available soil moisture


The amount of water held between field capacity and permanent wilting point is the total
available soil moisture, i.e. water available for the crop to use.

Sa = 10(θFC-θWP) [mm(water)/m(soil depth)]

The total amount of water that a crop can extract from its root zone of Z meter is

TAM = ZSa =10Z(θFC-θWP) [mm(water)]

The magnitude of Sa for any soil is very important. It shows the capacity of the soil to supply
water to the crop over time. A low value of Sa indicates early wilting point of crops and a
high frequency of irrigation will be required to keep soil moisture at acceptable levels.

The magnitude of Sa is a function of soil texture and structure. Coarse textured soils (sandy
soils) have lower Sa than fine textured clays, while loams are intermediate. This is because
the finer the particle size the larger is the surface area for water adsorption. Therefore, soils
with fine particles will have higher water contents at both field capacity and wilting point
than soils with coarser particles (Table 6.2). As an example of the calculation of the total
available water is given in Table 6.3.

Table 6.2: Average soil moisture constants and physical properties for different soil types
Textural ρb Θsat ΘFC ΘWP Sa
class g/cm³ (vol %) (vol %) (vol %) (mm/m)
Sand 1.65 38 15 7 80
Sandy loam 1.50 46 21 6 150
Loam 1.40 47 31 10 210
Clay loam 1.35 44 40 26 140
Silty clay 1.30 51 42 25 170
Clay 1.25 54 45 27 180

53
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Table 6.3: Total available soil moisture contained in the root zone
Calculate the total available soil moisture (TAM) contained in the root zone of
soybeans (Z=0.6m) cultivated respectively on a sandy, loamy and clay soil
Soil type FC (vol%) WP (vol%) Sa (mm/m) TAM (mm)
Sand 15 7 10(15-7)= 80 0.6(80)= 48
Loam 31 10 10(31-10)=210 0.6(210)=126
Clay 45 27 10(45-27)=180 0.6(180)=108

Given the soil texture in percentages the soil moisture content at SAT, FC and WP can be
estimated using the Soil Water Characteristics Hydraulic Properties Calculator freely
available at: http://wilkes.edu/~boram/soilwatr.htm

Exercise:
Use the Soil Water Characteristics Hydraulic Properties Calculator freely available at:
http://wilkes.edu/~boram/soilwatr.htm to compute the moisture content at FC, PWP and SAT
for the soil samples with soil texture for various profile depths given in the following Table
and also compute the TAW per profile depth. Also determine the textural class.

Profile Depth Soil Texture Textural class Moisture Content TAW


(cm) Clay Silt Sand FC PWP SAT (mm)
(%) (%) (%) (vol%) (vol%) (vol%)
0-25 56 23 21 Clay 45.8 32.4 54 34
25-45 45 21 34 Clay 37.3 25.0 51.8 25
45-65 47 23 30 Clay 39.2 26.3 52.4 26
65-90 35 29 36 Clay loam 32 19 50 33

6.4 Plant-Soil-Water system

6.4.1 The root system and rooting depth


The soil acts as a storage medium for water. Plants are anchored in the soil by their roots. The
roots also act as the organs of water absorption.

The root zone is the effective area ramified by roots which water uptake occurs. It defines the
lower limit of the soil volume effective as water storage for plants. The effective root zone
depends mainly on the crop type, its growth stage and the total depth of soil available for
rooting.

The effective root depth changes with crop growth (Figure 6.3). At sowing it is about twice
the seeding depth i.e. the zone from which the germinating seed and young seedling can
extract water. Next, the depth increases until a maximum is reached at the peak growth,
usually at the end of the development stage. The rate of root depth increase can be assumed to
be either linear or exponential with time depending on the crop. Table 6.4 gives the effective
rooting depths of main crops, i.e. the soil depth in which the bulk of the roots are
concentrated and which should be considered when designing irrigation systems. The base of
the rooting zone might be deeper.

54
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Figure 6.3: Effective rooting depth of maize crop

Root distribution is not uniform throughout the profile. The density is highest at the top 5 –
20 cm and decreases to zero at the base of the root zone. Water extraction generally follows
root distribution. The extraction of water follows a roughly linear pattern, with maximum
value at the top. Consequently the fraction of soil water available for plant growth (p) is not
uniform throughout the whole profile. As the use of different values of p complicates water
availability calculations, use is made of average (or mean) values, varying from 0.2 to 0.65
for most crops.

Table 6.4: Effective rooting depth (Z) and fraction of available soil water (p) for different
fully grown crops
Effective rooting depth, Z (m) p
Small vegetables 0.40 0.25
Onions 0.30-0.50 0.25
Carrots 0.40-0.70 0.35
Lettuce 0.20-0.40 0.30
Celery 0.30-0.50 0.20
Spinach 0.30-0.50 0.20
Strawberry 0.30-0.60 0.15
Cabbage 0.40-0.60 0.40
Vegetables (Leguminosae) 0.50 0.40
Beans (green) 0.50-0.80 0.45
Beans(pulses) 0.60-1.20 0.50
Groundnut 0.50-1.00 0.40
Peas 0.45-0.90 0.35
Soya bean 0.60-1.30 0.55
Vegetables (Solananaceae) 0.60 0.30
Tomatoes 0.60-1.20 0.40
Potatoes 0.40-0.60 0.30
Sweet peppers 0.50-1.00 0.25
Eggplants 0.40-0.60 0.30
Vegetables (Cucurbiaceae) 0.60 0.45
Cucumbers 0.50-1.00 0.40
Squash 0.50-1.00 0.40
Melon 0.60-1.00 0.35
Water melon 1.00-1.80 0.55
Pumpkin 0.90-1.20 0.40

55
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Fibre crops 1.20 0.60


Cotton 0.80-1.50 0.60
Flax 1.00-1.50 0.50
Sisal 1.00-2.00 0.80
Oil crops
Sunflower 0.80-1.50 0.60
Safflower 1.00-1.50 0.60
Cereals 1.20 0.55
Barley 0.90-1.20 0.60
Oats 0.60-0.75 0.60
Wheat 0.75-1.20 0.55
Maize 0.60-1.00 0.50
Sorghum 0.60-1.50 0.60
Millet 0.60-1.20 0.60
Pasture and Fodder crops 0.70 0.50
Pasture 0.40-0.80 0.50
Alfalfa 1.00-1.50 0.40
Clover 0.60-0.90 0.35
Sudan grass 0.90-1.20 0.40
Sugar cane 0.80-1.50 0.55
Grapes and Barries
Grapes 1.00-2.00 0.60
Coffee 0.80-1.50 0.40
Tea 0.80-1.50 0.30
Cacao 0.80-1.50 0.20
Fruit tree 1.50 0.50
Dates 1.50-2.50 0.60
Olives 1.20-1.70 0.65
Citrus 0.60-1.50 0.50
Deciduous trees 1.00-2.00 0.50
Palm trees 0.70-1.10 0.65
Avocado 0.80-1.20 0.40
Banana 0.50-0.90 0.35

6.4.2 Readily Available Moisture (RAM)

Critical moisture content


As has been pointed out earlier water uptake gets progressively more difficult as plants
absorb water between field capacity and permanent wilting point. Although water is
theoretically available up to wilting point the experience as shown that the evapotranspiration
rate drops below its potential level (ETact<ETcrop) and that growth reduction occurs before this
point is reached (Figure 6.4). The soil moisture content is said to be at the critical moisture
(θcritical) when the evapotranspiration starts to drop below ETcrop. Irrigation gifts therefore
should be planned before this point and rather much earlier to maintain full
evapotranspiration and optimal growth.

56
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Figure 6.4: Reduction in crop evapotranspiration due to water stress

Depletion factor
Readily available soil moisture (RAM) is the fraction (p) of total available soil moisture (Sa)
that a crop can extract from the soil without suffering water stress (ETact = ETcrop).

RAM = p.TAM = p.Z.Sa [mm(water)]

TAM = total available moisture = Z.Sa in mm(water)


Z = depth of the root zone in m

The critical moisture content is given by: Θcritical = θFC - p(θFC - θWP) [vol%]
θFC = soil moisture content at field capacity (vol%)
θWP = soil moisture content at wilting point (vol%)
Integrating the above equation yields: RAM = 10Z(θFC - θcritical)

RAM is a function of both the crop and the soil. It indicates the tolerance of the crop to water
stress because it gives the fraction of Sa that can be safely removed before stress occurs. If in
irrigation scheduling soil moisture is at field capacity just after the water gift, one can allow
the crop to take up water equal to RAM before providing the next water gift. In this sense
RAM is also called “allowable depletion”.

Likewise, the fraction (p) of available soil water is known as the “depletion factor”. The
magnitude of p has to be decided for each crop in a given soil. Generally (p) is determined by
the rooting characteristics of the crop (depth and volume of soil penetrated by the roots at a
given growth stage). Thus the value of p is chosen to minimise risk. The higher the risk (i.e.
sensitive crop or growing stage) the lower the value of p.

6.5 Yield response to water


With increasing water stress the yield starts to decrease. For many crops ETcrop shows a direct
relationship with dry matter production or yield. When ETact = ETcrop, water supply is
optimum. No stress exists and the crop produces an optimum yield. On the other hand when

57
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

ETact is reduced by water stress, the yield will be less than optimum. This forms the basis of
yield prediction based on the level of water supply. The relation is:
 Ya   ETa 
1    Ky  1  
 Ym   ETc 

Ya = actual harvested yield; Ym = maximum possible yield;


ETact = actual crop evapotranspiration; ETcrop = potential crop evapotranspiration;
Ky = yield response factor

Ym is the maximum possible yield that can be obtained in the given environment by a well
adapted, high yielding, variety under optimal water supply conditions. Ym is determined
through experiments. ETcrop is the total seasonal evapotranspiration that can be obtained by a
disease free crop with optimum water supply. It is either experimentally determined with
lysimeter or estimated from ETo and Kc. ETact is the actual evapotranspiration, which is either
measured using any of the soil moisture measurement techniques or estimated by a soil water
budget analysis. Finally, Ky expresses the effect of water deficit on yield for the given crop. It
depends on the drought resistance of the species and its growth stage (Ky <1, more resistance;
Ky >1, less resistant). Values of Ky for several crops at various growth stages are given in
Table 6.5. A detailed description of the procedure for the estimation of ETact, Ym and Ky is
given in FAO Irrigation and Drainage Paper No 33. An example of the calculation of yield
reduction is given in Table 6.6.

Table 6.5: Yield response factor Ky


Crop Vegetative period (1) Flowering Yield Ripening Total
Early Late total Period Formation Growing
(1a) (1b) (2) (3) (4) period
Alfalfa 0.7-1.1 0.7-1.1
Banana 1.2-1.35
Bean 0.2 1.1 0.75 0.2 1.15
Cabbage 0.2 0.45 0.6 0.95
Citrus 0.8-1.1
Cotton 0.2 0.5 0.25 0.85
Grape 0.85
Groundnut 0.2 0.8 0.6 0.2 0.7
Maize 0.4 1.5 0.5 0.2 1.25
Onion 0.45 0.8 0.3 1.1
Pea 0.2 0.9 0.7 0.2 1.15
Pepper 1.1
Potato 0.45 0.8 0.7 0.2 1.1
Safflower 0.3 0.55 0.6 0.8
Sorghum 0.2 0.55 0.45 0.2 0.9
Soybean 0.2 0.8 1.0 0.85
Sugarbeet
Beet 0.6-1.0
Sugar 0.7-1.1
Sugarcane 0.75 0.5 0.1 1.2
Sunflower 0.25 0.5 1.0 0.8 0.95
Tobacco 0.2 1.0 0.5 0.9
Tomato 0.4 1.1 0.8 0.4 1.05
Water melon 0.45 0.7 0.8 0.8 0.3 1.1
Wheat
Winter 0.2 0.6 0.5 1.0
Spring 0.2 0.65 0.55 1.15

58
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Table 6.6: Example of calculation of yield reduction


Crop: Maize Growing period: 1 May to 1 September
Seasonal water shortage is 130mm
Maximum yield (Ym) is 9 ton and Seasonal Ky is 1.25
May June July Aug Total
Growing stage Establishment Vegetative Vegetative Yield
Length (days) 25 30 30 35 120
ETcrop 90 190 280 270 830
Seasonal relative evapotranspiration deficit: (1-ETact/ETcrop)= 0.16; Ky =1.25
Relative yield reduction: (1-Ya/Ym) =1.25×0.16 = 0.2
Maximum Yield, Ym = 9 ton then Ya= (1.0- 0.2)×9 = 7.2 ton

6.6 Actual evapotranspiration


As long as the root zone depletion is smaller than the readily available soil water (RAM), the
soil supplies water fast enough to meet the atmospheric demand of the crop, and water uptake
equals:
ETc = Kc×ETo

where, ETc is crop evapotranspiration [mm/day]; Kc is crop coefficient [-] and ET0 is
reference crop evapotranspiration [mm/day].

When the soil water content decreases, water becomes more strongly bounded to the soil
matrix and is more difficult to extract. When the root zone depletion exceeds RAM, the root
zone depletion is high enough to limit evapotranspiration to less than potential values and the
crop evapotranspiration begins to decrease in proportion to the amount of water remaining in
the root zone. The effects of water stress can be described by multiplying the crop
evapotranspiration with a water stress coefficient Ks (Allen et al., 1998). If the root zone
depletion is smaller than RAM, Ks =1. For root zone depletions larger than RAW, Ks is
given by:
TAM  Di

i
K s ,i
TAM i
 RAM i

where,
- Total available water in the root zone on day i, TAMi = 10(fc - wp)Zr,i mm
- Readily available water in the root zone on day i, RAMi = p TAWi mm
- Root zone depletion in the active layer on day i, Di = 10fcZi - Wi mm
- Soil water content in the root zone on day i, Wi =10Ziθi mm
- Ks,i is water stress coefficient on day i -
- wp is the water content at wilting point vol%
- p is soil water depletion factor for no water stress -

Actual evapotranspiration in the root zone is given as:


ETai = Ks,iETci

If D1,i  RAW then Ks,i = 1


TAW  D 1,i
If D1,i  RAW then K s ,i

i

TAW i
 RAW i

where, ETc,i is crop evapotranspiration on day i [mm] and the other terms have already been
explained.

59
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

The soil water depletion factor depends on the type of crop and ETc (Doorenbos and Pruitt,
1975). The fraction p is function of ETc and a numerical approximation for adjusting p for
ETc rate is: 0.1  p = pTable 6.4 +0.04(5-ETc)  0.8

6.7 Soil water balance

To generate or evaluate irrigation schedules, information of the moisture content in the


rooting zone is required. This moisture content is simulated by means of a water balance
model. Such model keeps track of all inputs of water through rainfall, irrigation and capillary
rise and all withdrawal of water through run-off, soil evaporation, crop transpiration and deep
percolation. These processes affect the moisture content of the rooting zone and consequently
the water balance can be written as:
W i  1  W i  P  I  Ge  ETa  RO  Drain

Wi+1 = the soil water content in the root zone at time i+1 (mm)
Wi = the soil water content in the root zone at time i (mm)
P = rainfall (mm)
I = irrigation (mm)
Ge = capillary rise from ground water (mm)
ETa = actual crop evapotranspiration (mm)
RO = runoff from rainfall or irrigation (mm)
Drain = deep percolation of rainfall or irrigation water (mm)

P, I, Ge, ETa, RO and Drain indicate or represent rainfall, irrigation, capillary rise, actual
evapotranspiration, surface runoff and deep percolation occurring between time i and time
i+1.

If the soil water content at the start of the irrigation season (W0) is known (by measurement)
and the changes in all factors are continually measured or calculated, then the water balance
equation can be used to calculate the soil water content at the subsequent days (intervals): i =
1, 2, 3, …, N = last. Depending on the required accuracy, the interval i may be one day, one
week, one decade or even interval between two irrigations. An example of the calculation of
water balance is given in Table 6.7.

Schematically (Figure 6.5) the root zone can be presented by means of a container in which
the moisture content may fluctuate. After irrigation (I) or heavy rainfall (P), which
thoroughly wets the rooting zone, the moisture content may be close to saturation. The
moisture above field capacity however cannot be held against the forces of gravity and will
drain out of the rooting zone (Drain). After some time, which depends on the drainage ability
of the soil, field capacity will be reached. Crop evapotranspiration will further deplete water
from root zone. As discussed before, between field capacity and critical moisture content the
water extraction by plant roots is at the potential rate as dictated by climatological demand
(ETa=ETc). If the moisture content in the root zone drops below the critical moisture content,
the actual evapotranspiration rate will be smaller than the potential rate. Furthermore, the less
water in the root zone, the more the actual evapotranspiration rate will deviate from the
potential rate. In the end the rate becomes zero when wilting point is reached. The moisture
content below wilting point can not be extracted by the plant roots.

60
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Table 6.7: Soil water balance

Calculate the daily water balance of the root zone of potato (Z =0.6 m, p = 0.5) cultivated on loamy soil
(θFC=31 vol% θWP =14 vol%) for 10-day period.
- at the start of the 10-day period, the root zone is at field capacity
- at day 2, a shower of 50 mm is observed (10% is lost by surface run-off)
- the mean potential crop evapotranspiration is 6 mm/day
- at the end of 10-day period, an irrigation refill the root zone up to field capacity
W0 = WFC =10ZθFC = 10(0.6)31= 186 mm
As long as the soil moisture content in the root zone is larger than θ critical, the actual evapotranspiration is equal
to Etcrop = 6 mm/day
θcritical = 31 –0.5(31-14) = 22.5 vol%
or the water content of the root zone may not drop below:
Wcritical =10Z θcritical =10(0.6)22.5 = 135 mm
Day Wi (mm) P I Ge ETa RO Drain Wi+1
(mm) (mm) (mm) (mm) (mm) (mm) (mm)
1 186 - - - -6 - - 180
2 180 50 - - -6 -5 - 219
Since the water content is above field capacity (219>W FC, water will drain out of the root zone till field
capacity is reached
-33 186
3 186 - - - -6 - - 180
4 180 - - - -6 - - 174
5 174 - - - -6 - - 168
6 168 - - - -6 - - 162
7 162 - - - -6 - - 156
8 156 - - - -6 - - 150
9 150 - - - -6 - - 144
10 144 - - - -6 - - 138
To refill the root zone up to field capacity, (186-138) = 48 mm is required
48 180

In stead of calculating water contents, one can also express the results as water shortage versus field capacity,
i.e. root zone depletion. At field capacity, the root zone depletion is zero. If water is consumed by
evapotranspiration, the depletion increases. If water is added by rainfall or irrigation, the depletion decreases.
RAM = pZTAM = 0.5(0.6)(170mm/m) = 51 mm
As long as the root zone depletion (D) is smaller than RAM the actual evapotranspiration, is equal to ETc = 6
mm/day
Day Di (mm) P I Ge ETa RO Drain Di+1
(mm) (mm) (mm) (mm) (mm) (mm) (mm)
1 0 - - - 6 - - 6
2 6 -50 - - 6 5 - -33
The depletion becomes negative (i.e. surplus of water above field capacity). Since the water content is above
field capacity, water will drain out of the root zone till field capacity is reached
33 0
3 0 - - - 6 - - 6
4 6 - - - 6 - - 12
5 12 - - - 6 - - 18
6 18 - - - 6 - - 24
7 24 - - - 6 - - 30
8 30 - - - 6 - - 36
9 36 - - - 6 - - 42
10 42 - - - 6 - - 48
To refill the root zone up to field capacity, 48 mm is required
-48 0

61
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

To avoid any water stress and to maintain full production, irrigation should be considered
before the actual crop evapotranspiration (ETa) drops below the potential level of ETc. This
answers the question when to irrigate.

The amount of water to apply is given by the root zone depletion, i.e. the amount of water
required to refill the root zone up to field capacity. Any surplus water will be lost by deep
percolation, since this water is not strongly be held to the soil matrix and will drain freely
under gravity. The planning of the timing and depth of irrigation is discussed in the next
chapter.

Figure 6.5: Schematic presentation of the water balance of the root zone

62
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 7- Irrigation scheduling


Irrigation scheduling means the planning of timing and depth of future irrigations. The
primary objective is to apply irrigation water at the right period and in the right amount. If
water deliveries are untimely or not in the appropriate amount, irrigation efficiency decreases.
Limited supply results in yield reduction due to water stress. Too much water may not only
result in deep percolation losses, which may leach relevant nutrients out of the rooting zone
but might decrease the yield as well.

Starting with a soil at field capacity, water is extracted by the crop at a rate equal to ETcrop.
As the water uptake progresses the readily available moisture is depleted and actual crop
evapotranspiration ETact starts to fall below the optimal level of ETcrop. At this point one
should irrigate and refill the root zone profile up to field capacity.

This gives the irrigation interval and amount of water to apply. Any delay results in restricted
water supply situation (ETact < ETcrop) and increasing water stress, until at permanent wilting
point the crop will no longer recover resulting in total crop failure.

7.1 Definitions

7.1.1 Root zone depletion


Root zone depletion expresses the shortage of water in the root zone with respect to its
maximum water holding capacity i.e. the field capacity. Any surplus of water will be lost
through deep percolation. At field capacity the root zone depletion is 0 mm.

7.1.2 Maximum depletion


The maximum allowable root zone depletion is the maximum amount of water that a crop can
extract from the soil without suffering water stress, i.e. the readily available soil moisture
(RAM). If the root zone depletion is larger than RAM, the evapotranspiration drops below its
potential level (ETact <ETcrop) and water stress occurs.

7.1.3 Water application depth (DA)


The practice in on-farm irrigation is to express the amount of irrigation water applied in
equivalent water depth (mm water). The depth is called the water application depth and
denoted here as DA.

Soil water holding characteristics determines how much water can be applied. If the actual
soil moisture content at time i is equal to θi, then the maximum water application depth is
equal to the root zone depletion. Indeed, irrigation should at maximum replenish soil
moisture up to field capacity, as surplus water will be lost through deep percolation below the
root zone. For rooting depth Z:

D A
 10  FC
 i
Z [mm]

where θFC and θi represent average values over the rooting depth (Z) of respectively the
moisture content (vol%) at field capacity and the actual moisture content (vol%).

63
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

For soil moisture content at time i equal to maximum depletion level, i.e. RAM, the net and
gross application depths are respectively:

RAM p  Z  Sa
D A ,n
 RAM  p  Z  Sa (net) D A,g
  (gross)
Ea Ea

where, DA,n = the net application depth (mm)


DA,g = the gross application depth (mm)
p = depletion factor (-)
Z = depth of root zone (m)
Sa = total available soil moisture (mm(water)/m(soil depth))
Ea = irrigation application efficiency. The subscript “a” indicates field level, but
could also be taken at block level or even scheme or project level.

Application depths are normally adapted to the irrigation method. Indicative values for
different irrigation methods are given in Table 7.1.

Table 7.1: Typical application depths for different irrigation methods


Irrigation method Application depth
Surface irrigation : basin irrigation 50 – 150 mm
furrow irrigation 30 – 60 mm
border irrigation 40 – 80 mm
Sprinkler irrigation 30 – 80 mm
Drip irrigation 10 – 30 mm

7.1.4 Water application duration (WAD)


For the scheme management the application depth is useful, but what is of more interest to
them is the time required to create this particular application depth i.e. the water
application duration (WAD). Normally the application duration will be given per hectare per
crop, but in the irrigation schedule this will be finally be converted to the duration per field
and/or farm and/or block. The application duration is basically expressed in seconds [sec] but
may ultimately be expressed in minutes or in hours or in combination of both.

For a water application depth of DA mm i.e. 10DA m³/ha or 10,000 DA l/ha, an area of Au ha
and a discharge (flow) of qu, l/sec, the volume of water (V) to be supplied and the water
application duration (WAD) are respectively:

V  10 D A A u [m³]
10 , 000 D Au

A
WAD [sec]
qu

7.1.5 Unit flow (qu)


The factor relating application duration and application depth is the unit flow (qu) i.e. the
flow received at the inlet of the unit, whether this is a field, a farm or a block of fields, having
the same crops. Unit flow is expressed in litres per second [l/sec].

64
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

7.1.6 Water application time (ti) and Irrigation interval (INT)


The scheme management will be interested to know when to irrigate (ti). The water
application timing depends on the choice one makes; water application can be related directly
to soil moisture content, which results in different intervals; but very often water application
is done at fixed intervals and soil moisture content will then vary.
The time between successive water applications is called the irrigation interval (INT) and is
measured from the start of one water application to the start of the next water application and
therefore includes the water application duration (WAD). The irrigation interval is basically
expressed in seconds [sec], but ultimately (in the schedule) will be expressed in hours or in
days or in both.

For optimum water supply, the soil moisture should not be depleted below its maximum
depletion i.e. RAM, or the soil moisture content at the end of the interval i (θi) should not
drop below θcritical. Thus, the water application time here is determined by soil moisture. In
reality there are several other ways for timing water applications and these will be discussed
in more detail hereafter.

The net irrigation requirement when combined with soil water characteristics enables to
specify when to irrigate. The maximum interval between successive water applications can be
calculated as follows:
- if, after the previous water application (at time (i-1)) soil moisture content was equal to
θi-1 (which not necessarily was equal to field capacity θFC) and
- For In being the average net irrigation requirement over interval i, the maximum duration
in days is:
10  i  1   i  Z
INT i

In

If θi = θcritical = θFC – p(θFC –θWP) i.e. maximum depletion, and

θi-1 = θFC i.e. maximum water application depth:

10 p    WP  Z pS a Z RAM
  
FC
INT i
In In In

Example
Given the actual soil moisture content θi =17.5 vol%, calculate the net application depth
which replenish the soil up to field capacity. What is the maximum interval between
successive water applications. The net irrigation requirement: In = 65 mm/decade. Sandy
loam soil (θFC =21 vol%, θWP =9 vol%). Maize (Z = 0.8m, p = 0.50).

Solution
DA = 10(21-17.5)0.8 = 28 mm
RAM = 0.50(21-9)10(0.8) = 48 mm
INT = 48/65 = 0.74 decade = 7.4 days.

The net application depth is 28 mm. For the given net application requirement and soil type,
the maximum irrigation interval is 7 days.

65
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

7.2 Real time scheduling

7.2.1 Plant observation


This is the probably the oldest method of irrigation scheduling. The crop will be irrigated,
when individual plants start showing visible signs indicating that they are beginning to
experience water stress. The signs range from colour changes and leaf curling to wilting
during the afternoon.

The method is easy and requires no data gathering or computations, but has the disadvantage
that by the time plants show symptoms of stress they have already suffered some growth and
yield reduction.

More modern methods involve monitoring certain plant physiological states, such as stomatal
closure using porometers, leaf water status using pressure bombs and leaf temperature using
infrared radiation. These methods are not used on a routine basis.

7.2.2 Soil moisture meters


Notable among these instruments are tensiometers, resistance blocks and neutron meters. The
principle of these instruments is that they measure certain physical characteristics of the soil,
which depend on its moisture contents. A calibration curve is then used to determine the
corresponding soil moisture contents. When the reading indicates critical soil moisture
content the crop should be irrigated. The disadvantages of the method include sensor
breakdowns or malfunction and the fact that the method does not give the amount of water to
apply at each irrigation.

Tensiometers
Tensiometers consist of porous cup connected by tube to a pressure gauge. Both cup and tube
contain air-free boiled water. The cup is embedded to a required depth within the soil and
maintained in water and temperature equilibrium with the soil. This allows the suction, with
which water is being held in the soil, to be registered by pressure gauge. The disadvantage of
tensiometers is that they are reliable only up to about 0.7 bars. They are also very difficult to
maintain in constant working state as any disruption of equilibrium with the soil causes errors
in the readings.

Electric resistance blocks


Electric resistance blocks are usually made of gypsum (plaster of Paris) with two electrodes
embedded. The block is buried at a required depth in the soil, where its water content
equilibrates with the soil. Resistance to electricity flow between the electrodes is a function
of the water content of the block and surrounding soil. Resistance blocks are sensitive to
wider range of moisture contents than tensiometers, but tend to deteriorate with time,
especially in saline soils.

Neutron moisture probes


Neutron moisture probes uses a neutron scattering system to monitor soil moisture contents.
It gives an easy method of measuring soil moisture. However, the equipment is expensive and
can be hazardous if not handled with the recommended caution. The method is more widely
used in agricultural research studies and cannot be recommended for individual small
farmers, both for its cost and skill required in handling the equipment.

66
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

7.2.3 Cumulative pan evaporation


This method is based on the cumulated pan evaporation (Epan) starting from the last irrigation.
It is based on the fact that Epan is an estimate of ETcrop.

In a sense it is a version of the soil water budget method, because the value of cumulative
Epan, at which water should be applied, is related to available soil moisture. When the value of
cumulative Epan is reached, the available water has been depleted to a level where irrigation is
necessary.

The disadvantage of the method is that additions to soil moisture storage from precipitation or
shallow ground water table are not considered. It is therefore most appropriate at locations,
where these additions do not occur during the irrigation season.

7.2.4 Soil water budget


This is the most complex, but in most cases, the most accurate of all the methods. Irrigation
timing and depth of irrigation water are determined on the basis of the soil water balance.
Information on the weather (ET0 and rainfall), the crop and the soil are required. The
disadvantage is the large amount of data, which must be processed.

Water is added to the soil moisture in the profile by rainfall, irrigation or capillary rise from
ground water table. Water is extracted from the soil moisture in the profile by
evapotranspiration, run-off and deep percolation. If the soil water content drops below a
critical value, the crop should be irrigated. The (net) volume of water to be given should be
maximal equal to the amount required to bring the rooting zone back to field capacity.

7.3 Practical irrigation schedules

Irrigation schedules where the application depth and the irrigation interval varies constantly
over the season may create real problems in practice. Not only it requires considerable skill
from the irrigators and from the farmers, but it also creates peak water demands at certain
times while at other times water requirements will be below average.

Many systems cannot deal with such a situation, the more so as water availability may be
constant because of restriction. Also water supply may not be under the direct control of the
scheme management. For gravity irrigation under restricted water supply, rotational supply is
practised, indicating not only fixed intervals, but also fixed volumes. Under less restricted
water supply and different irrigation methods other options are possible and are discussed
below.

7.3.1 Variable interval(s), fixed amount(s)


In this case a fixed application depth (DA) is selected and the interval length has to be
adjusted to the (net) irrigation requirements. The flexibility in the timing of the successive
applications then will determine in how far over – or under irrigation will occur.

7.3.2 Fixed interval(s), variable amount(s)


The application depth can be determined to conform to the deficit at the end of each fixed
interval (θFC – θi). In as far as soil moisture depletion exceed readily available moisture (θi >
θcritical), water stress will occur, resulting in yield reduction.

67
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

7.3.3 Fixed intervals and fixed amounts


This is the most frequently used distribution method: gravity irrigation with rotational supply.
Both the interval(s) and the amount(s) may not correspond to the soil moisture conditions, i.e.
the amount may exceed or fall short of soil moisture depletion or the water requirements for
the fixed interval may not correspond with readily available soil moisture.

7.4 Adjustments of irrigation schedules to environment

7.4.1 Various weather conditions


When planning a practical irrigation schedule in climates with highly variable rainfall,
alternative schedules might have to be developed for the different weather conditions. A
schedule should not result in (excessive) deep percolation losses in rainy years or in crop
stress in dry years.

7.4.2 Different soil types


An irrigation schedule designed for a given crop on one soil type might not always be valid
on the other soil types. Since the readily and total available soil moisture varies from one soil
to another, an irrigation schedule valid for one soil type, might result in deep percolation
losses and crop stress on another soil type.

7.5 Deficit irrigation

Often water availability – actual water supply – is limited and below irrigation requirements.
There is no enough water to replenish soil moisture up to field capacity and to irrigate before
all the readily available soil moisture (RAM) is depleted. Consequently the crop will
experience water stress, to a degree depending on its growth stage, its sensitivity to water
stress and the seriousness of the soil moisture deficit.

The following are strategies to reduce water stress and yield reductions:
- Fill the soil profile up to field capacity over the maximum root depth + some 20 – 30 cm
for capillary rise. This should be done prior to sowing or at the initial stage when water
availability is still relatively high and crop water requirements relatively low.
- Make the irrigation interval as long as possible, even inducing slight water stress as this
enhances root development by the young plant, looking for water (“growing after water”).
To enhance rapid and deep root growth a water deficit during the early growth periods can
be advantageous for some crops (maize).
- For some crops the sensitivity to water stress during a sensitive period is less pronounced,
when water deficit has been experienced during a preceding period (For instant maize,
which is less sensitive to water stress during flowering when water stress has been
experienced during the vegetative period).

Methods of dealing with short water supply at block or scheme level, such as crop selection,
change of planting date, selection of drought resistant varieties, staggered sowing/planting
and fixation of water allowances per individual block will not be discussed here.

Assignment 3: Deficit Irrigation scheduling


Objective: The student will learn to evaluate the impact of supplementary irrigation on crop
yield and water productivity using AquaCrop model using data for a given area.

68
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 8.0 - Irrigation methods


8.1 Surface irrigation
Surface irrigation refers to large group of irrigation methods in which water is distributed by
gravity over the surface of the field. Water is typically introduced at the high point or along
the edge of a field and allowed to cover the field by overland flow. The efficiency and
uniformity of irrigation is dependent upon soil uniformity, quality of land grading, field
topography and control of the relationship between stream size, soil infiltration rate and
duration of application.

The defining features of surface irrigation methods is that the soil is used as the transporting
medium (as opposed to pipelines or through the air, as with sprinklers). The soil also controls
the depth of infiltrated over time (as opposed to the application rate being controlled by
sprinklers or emitters). Furthermore, the infiltration and advance characteristics of surface
irrigated fields change with time, making it impossible to pre-determine many management
recommendations. Irrigation control by field management is more important with surface
methods than for mechanical systems where design and equipment replaces much of the need
for intensive management.

Surface irrigation methods have two basic categories: ponded and moving water. The
moving water methods require some runoff or ponding to ensure adequate infiltration at the
lower end of the field. Tail water return flow systems are often required by law to prevent
runoff from farms. They also provide valuable tools for labour reduction and improved
uniformity is designed properly.

8.1.1 Types of surface irrigation methods

Basin
Basin irrigation is a ponded water irrigation method used to apply water to a level area of
land bounded by dikes (Figure 8.1). The soil surface is not kept flooded. Because water is
ponded until it infiltrates, there is no runoff. In rainfall areas consideration must be given to
provide for draining unneeded water. This method of irrigation can be used for both field and
row crops, often interchangeably, with or without interior ridges or with wider flat beds. It is
also used on trees and vines. The soil intake rate should be the same within any one basin as
infiltration uniformity is very sensitive to variation of intake rate. The basins need not be
rectangular nor straight sided and the dikes need not be permanent. Under good management,
a pre-determined volume of water is rapidly discharged into the basin.

Figure 8.1 A level basin

69
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Crop damage due to inundation is reduced by use of precisely levelled basins. Rain water and
excess irrigation is uniformly distributed over the entire area rather than being concentrated
in low spots or at lower end of the field, where crop damage typically occurs. However,
since there is no runoff, excessive over application by irrigation or excess rainfall can cause
crop damage. In areas with high intensity rainfalls and low intake rate soils, surface drainage
systems must be considered.

Basins may have some advantages over other manually controlled irrigation methods because
of ease of operation, simple equipment requirements, potentially low labour requirements and
the capability to utilise large fixed rates streams. The method is most efficient with uniform
soils, precise levelling, and large stream sizes relative to basin area. None of these are
absolute requirements, but they tend to make basins more effective.

Some means of erosion control at the inlet to the basin is required if the stream is large.
However, standard erosion control structures are available. When the supply canal is below
the field level, no erosion will occur since, to irrigate, the water level in the canal is simply
checked up until water begins to flow into the channels or basin.

Basins sizes range from a few square meters to 10 – 15 ha. Basin size and length are often
limited by topography, soil texture changes, infiltration, depth of topsoil (when extensive
levelling is involved), available stream size, degree of land levelling, and farmer equipment
capabilities. From a design standpoint, basin length is limited primarily by soil uniformity,
intake rate, and available stream size as well as by potential for erosion. Large basins must
consider earth curvature in land grading, as a horizontal plane does not follow the earth’s
curvature.

Border strip
Border strip (contoured, graded, and guided) irrigation consists of a sloping strip of land
essentially level across the strip and bounded by borders (dikes, levees, ridges, etc) to prevent
lateral spread of water (Figure 8.2).

Figure 8.2 A straight, graded border

The borders are usually parallel for convenience of cultivation, but this is not necessary for
irrigation, particularly for contour border strips. A relative large stream of water is turned in
at the upper end, spreads across the strip, and advances down the strip. When all pertinent

70
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

items (unit stream size, soil moisture deficiency, intake rate, and length) are in proper
relations, the stream is cut off when the advancing water is between about 0.6 of the length
for slow intake soils to 0.9 of the length for fast intake soils if little or no tail-water is desired.
The lower end of the strip is irrigated by water supplied from the temporary surface storage
on the upper end. It is possible to have only a small amount of water run off (perhaps 10%)
and the upper and lower ends are generally under irrigated relative to the middle portion. This
scenario requires one specific combination of soil moisture depletion and flow rate.

All variations of border strip irrigation flood the soil surface and will cause some soils to
form a crust. Such a crust may inhibit sprouting of seeds. For some crops, wetting of the plant
may be detrimental due to fungi and diseases that may find a suitable moist environment.
This tends to be more a problem on the finer texture soils that remain moist for a long period.
All crops that are not limited by the above cultural consideration can be adapted to the border
strip method. Field crops planted in rows: grain, hay, pasture, crops planted flat: orchards and
vineyards sometimes planted on the border or in the middle or edge of the border strip, are
easily irrigated this way. Ponding at the lower end of a blocked end border strip frequently is
detrimental to crops, especially if ponding continues for longer than a few hours duration, as
may occur on fine textured soils.

Border strip irrigation is the most complicated of all irrigation methods. While the method
has a very high potential application efficiency, high efficiency is seldom obtained. This is
due to inadequate knowledge of the need for practical compromises of the management
allowed soil moisture deficit (design control) to the actual soil moisture deficit on the date of
irrigation, and the optimum stream size and distance and time of cutoff.

A large stream size allows a large area covering many strips to be set at one time, increasing
labour efficiency. It will require a larger distribution system with a higher capital cost. The
labour needed for a conventional small flow border strip system siphons from a ditch is
appreciable. Gates from the ditch greatly reduce the work but the irrigator still has to go to
the field for each set. If a large stream is available to allow large sets and the water control is
easily accomplished because of adequately designed controls, very little labour is required.
Piped lines may be desirable and economical. It is practical for one man to control irrigation
on 250 to 400 ha, opening and closing a few gates or valves.

For rapid intake rates, strips may be only 100 m long. For slow intake conditions, strip length
may need to be 800 to 1,000 m to provide a flow duration that allows time for an adequate
irrigation to infiltrate. Also, stream sizes need to be varied. Lengths and stream sizes other
than the ideal are practical but at reduced efficiency. Changing the slope or direction may be
practical. Uniform intake rate and water holding capacity in the root zone are very desirable
within any one unit that requires similar management, especially in the down slope direction.

Furrows
Furrows are sloping channels formed in the soil (Figure 8.3). Water moves down the furrow
and infiltrates for a longer time at the upper end (time of application, Ta) than at the lower
end (time of infiltration at the lower end, Tl) by the duration of the time of advance, Tadv, less
the time of recession, Tr. The time of recession is relatively small on sloping fields, but can
be significant on low gradients and soils with low intake rates. In order to assure adequate
water and time at the lower end, runoff must occur, so return flow systems may be essential.

71
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Infiltration is often a slow process that occurs laterally through the wetted perimeter of
furrow.

Figure 8.3: Graded furrows

Assuming that the soil is uniform and that good land grading has occurred, the three most
important hardware items which facilitate simple and efficient furrow irrigation are: (i) a tail-
water return flow system which incorporates a reservoir, (ii) short furrows for an acceptable
advance ratio, (iii) a large variable water supply stream.

There are many furrow shapes; typically “Vee”, trapezoidal, parabolic, and broad-flat shapes
with wetted widths varying from 150 to 750 mm or more.

Furrows are well adapted to row crops and orchards or vineyards. They are less well adapted
to field crops if cultural practices require tractor travel transverse to the furrows. Corrugations
and shallow broad furrows can be used with flat field crops to alleviate some cultural
problems. Furrows can even be adapted for wide-spaced vine crops such as melons or crops
that should not lie on the wet soil. Vines may be “trained” out of the furrows, furrows
adjacent to the plant may be relocated out or abandoned as plants extend, or alternate or
widely spaced furrows away from the plant may be used as the root system expand (part area
wet management).

Small, shallow seeded crop irrigations require carefull land grading of the surface as
overtopping furrows may cause crusting of the soil and high spots that do not become wet so
that seeds do not germinate.

Furrows can be shaped and spaced to allow appreciable dry soil between them, to adapt to a
great range of row spacings, or to create flat beds between them. Furrow spacing and shape is
often determined by the spacing of equipment used for cultural activities and by ideal plant
spacing.

The use of alternate side irrigation, applying water in alternated furrows on either side of a
crop row at each irrigation, can be a desirable practice for row crops where higher frequency
and smaller application depths are desirable. In this way, an individual crop row receives
water at half frequency on one side or the other. Twice as large a set can be made with the
same size stream. Labour needs can be minimised by adequate use of proper equipment and
semi-automation of the supply. Alternate side irrigation are impractical if soils crack, as the
water will move laterally into adjacent dry furrows.

72
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

The set time or application duration should be the time needed to infiltrate the desired depth
of water at the low end of the furrow, plus the advance time, less the recession time if any.
The spacing between the wetted edges of adjacent furrows should not exceed twice the
distance across which water will move laterally by capillary movement within the duration of
irrigation at the lower end. The entire surface between furrows may not appear wet but may
be adequately wet below the surface. The distance between the wet edges can be adjusted by
changing the shape as well as furrow spacing centre-to-centre. Such changes can greatly
affect the intake rate and required irrigation duration.

Furrow lengths usually range between 60 and 600 m, but this range can be exceeded in very
low intake rate soils unless root zone saturation occurs. Lengths are dominantly controlled by
intake rates and stream size as well as field length. Furrow lengths may be shortened by use
of supplemental cross ditches or grated pipe on continuous furrows. Use of large furrow
streams relative to the furrow length reduces the excess time water is at the upper end and
will result in more uniform infiltrated depth but more need for cutback streams or return flow
systems to reduce runoff. The maximum stream size must not be erosive.

Intake rates in furrows are sometimes quite variable, even presuming that soils are uniform
for the entire length. The variations are caused by cultural practices. New furrows have open
soil conditions and high intake rates. Reuse of the same furrow after the soil has settled or
slicked over will reduce the intake rate and assist in increasing the rate of advance. Driving a
tractor wheel and/or dragging an object down a furrow will reduce the intake rate, and can
make adjacent furrow infiltration more consistent, especially for new furrows. At the upper
end of V-furrows, the stream size, wetter width, and hence, the intake rate, are greater relative
to the lower end. In the broad furrow, there is very little difference between ends as the
furrows are level across.

Because of the numerous management controllable items, furrows can be adapted and
modified for many conditions within the limits of soil uniformity and amenable topography.
If all conditions are just right it is possible to have high uniformity, and with a return flow
system, a high efficiency is possible. However, because of the uniformity and efficiency are
highly dependent upon management, management can severely affect performance. The
method can be used on short or long fields and with all but the most extreme intake rates,
providing the rates are quite uniform. In general, the soil moisture depletion needs to be
greater than what is optimum for sprinkler and drip systems.

8.1.2 Capabilities and limitations

The primary advantages of surface irrigation methods are:


- They typically use simple irrigation equipment, and if equipment fails, water can often be
applied if enough hand labour shovel work is employed;
- On many soils and topographies, these methods have the lowest initial capital investment;
- Silty and dirty water can be used, where filtration of that water might be very expensive
for sprinkler or drip;
- High efficiencies and uniformities are possible given the correct combination of
conditions such as medium-heavy soil types, excellent land grading, uniform soils,
excellent management, a large variable flow rate supply, and tail-water (runoff) return
systems;

73
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

- In many cases, surface irrigation systems require no pumps. This advantage, however, can
be considered a negative because tail-water return systems require pumps, and tail-water
return systems are essential for the effective management of most furrow and border strip
systems; and
- If the system have flexible and large flow rate supplies and tail-water return systems (for
sloping methods), they can have low labour requirement.

The primary disadvantages of surface irrigation methods are:


- They require the most “art” of all the irrigation methods, both to obtain a high distribution
uniformity and high application efficiency. In general, people have not learned the art;
- The distribution uniformity of surface irrigation methods is extremely sensitive to soil
differences within a field;
- It is difficult to apply small depths of water evenly with a high application efficiency;
- Irrigation scheduling, on a scientific basis, is difficult and requires excellent historical
records on each field;
- Excellent land grading is required for some of the surface irrigation methods. It is
difficult to obtain excellent land grading on small fields; and
- The efficient furrow systems of the western U.S. that utilise tail-water return systems and
have excellent land grading are difficult to economically duplicate in the very small fields
found in other areas of the world.

Crops
Virtually all crops have been grown using surface irrigation methods. Surface irrigation
methods are best adapted to soil moisture depletion of 50 mm or greater if a good distribution
uniformity and high application efficiency are desired; some crops grow best with lower soil
moisture depletions and ideally would be irrigated more efficiently with other methods.

Soils
Surface irrigation methods are most successfully used on medium-low intake rate soils. Many
clay loam and silt loam soils have medium-low intake rates at the end of the irrigation season,
but they may have very high intake rates early in the season. For this reason, farmers may
irrigate with sprinklers early in the season and then switch to furrows later when the intake
rates are low.

Topography
In general, surface irrigation methods are suitable for slopes of less than 1%. Steeper slopes
are irrigable but generally some degree of water control is lost and erosion can be a serious
problem.

Water supply
Previous sections have pointed out the importance of having a flexible and reliable water
supply. If the irrigation project or water well/borehole does not provide a flexible supply, on-
farm reservoirs can be constructed to provide flexibility and to also facilitate simple and
efficient use of tail water return flows.

Salinity and water quality


For surface irrigation methods with beds, salinity damage can be minimised by using special
bed shapes such as creating a ridge in the centre of the bed to which the salt will “wick up”.
Planting is done on the edges of the bed.

74
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Basin systems provide excellent opportunities for salinity and leaching control. Ponding of
water on the surface can leach salts out of the root zone. All the surface methods can leach
the majority of the root zone with sufficient time. However, furrow beds and border strip
ridges cannot be leached effectively. In many areas with salty water, sprinklers are used for
the initial germination and pre-irrigation events, and later furrow irrigation is used.

An advantage of all the surface irrigation methods is that the plant leaves are not contacted by
salty water, as is the case with most sprinkler irrigation methods.

Climate
Surface irrigation methods are generally incapable of applying small depths of water
uniformly across a field, except during the latter part of the growing season when the intake
rates of the soil may have significantly decreased. Therefore it is difficult to manage surface
irrigation methods with precise small water applications that would allow a farmer to leave
part of the root zone dry enough to store rain. An option used by some farmers is to use the
alternate side technique in the events that when it rains at least half of the field will have
ability to store some rainfall.

Evaporation from wet soil surfaces before the crop maturity is greater in the full coverage
methods (e.g. basin, border strip) than those with only partial wetting (furrow). Surface
irrigation methods have an advantage over many sprinkler methods in that they are
insensitive to wind conditions.

Efficiency
Surface irrigation methods have broad range of potential application efficiencies that are
affected by topography, flexibility of water supply, soil conditions, boundary constraints and
management. Under ideal conditions, most methods, when well designed and well operated,
can have high potential application efficiencies. However, there are nearly always constraints
that cause conditions to be less ideal. Therefore, the practical potential application
efficiencies will be somewhat less.

Irrigation scheduling
Scientific irrigation scheduling is more complicated and difficult with surface irrigation
methods than with drip and sprinkler. The soil (rather than sprinkler or emitters) controls the
rate of water infiltration set, making it important that the soil moisture depletion match the
amount of water that one expects to infiltrate at low quarter point during an irrigation event.
Because both the target soil moisture depletion and the intake rate will vary from one
irrigation event to another, efficient irrigation scheduling requires care-full observation and
documentation of irrigation practices and infiltrated depths on each soil and for each
irrigation event. After a history of is built up, one can estimate how fast water will advance
and recede, and how much water will infiltrate at the lower quarter point for a typical
irrigation at a particular time of the year.

In short, surface irrigation scheduling is much more “artful” than for sprinkler or drip
irrigation, but with careful attention to historical records, the art factor can be reduced.

75
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

8.1.3 A comparison

Table 8.1 summarises the different features and aspects of suitability of main methods of
surface irrigation, as discussed in the previous sections. For more detailed information on
determining the appropriate method for a particular situation, the reader is referred to
standard irrigation textbooks.

Table 8.1 A comparison of the main methods of surface irrigation


Characteristic Basins Borders Furrows
Flow section Wide, rectangular Wide, rectangular Various shapes
Infiltration Vertical only Vertical only Vertical and sideways
Slope Longitudinal slopes Slope length usually Slope length
and cross-slopes <0.5%, but can go up to usually<1%, but can go
zero or nearly zero 4% in sod crops; avoid up to 3%.
cross-slopes.
Method of Constant flow Fixed flow, cutback, and Fixed flow, cutback, and
operation re-use. re-use.
Topography Flat, or uniform, Relatively flat, uniform Rather wide range of
gentle slopes slopes. topographic conditions.
Crops Most crops that Most dry-foot crops Most row crops
tolerate some
inundation;
Narrowly-spaced
crops; orchards.
Soils Medium to fine Moderately low to High and low intake
textured moderately high intake rates, not on soils prone
rates to puddling.

8.2 Drip Irrigation

Drip irrigation refers to a variety of irrigation methods in which water is delivered directly to
small areas adjacent to individual plants through emitters or applicators placed along a water
delivery line. In an orchard or vineyard there will typically be one or more emission devices
per tree. For row crops and field crops the emission devices are spaced closely enough so that
the capillary action of the soil provides water to each plant root zone. It is unusual to use drip
on broadcast field crops because of difficulties in wetting all of the plants and the low prices
of such crops.

Drip irrigation systems are almost always “solid set”, meaning that equipment such as hoses
and emission devices remain in one place during the growing season. Systems may be
permanently installed or may be portable and moved to a different field after an irrigation
season is completed. Layout of drip irrigation system with various types of emitters is
presented in Figure 8.4.

Drip irrigation systems require very clean water to avoid plugging of emission devices.
Filtration components represents a major portion of the cost and maintenance of drip
irrigation. In addition, chemigation is generally required to avoid plugging due to bacterial
growth and/or chemical precipitation in the laterals and emission devices. Flow rates of
individual emission devices are typically very small.

76
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Figure 8.4: Layout of drip irrigation system

Because drip irrigation systems are “solid set”, many of them are easily automated. They are
ideal systems for irrigation managers who are interested in fine tuning the applications of
water and fertiliser (fertigation) through the irrigation system. Irrigation water is generally
applied through emission devices daily or several times per week.

8.2.1 Capabilities and limitations

Drip irrigation systems have the following typical advantages over some other irrigation
methods:
- They can be used effectively on extremely steep ground;
- They require minimal land grading. Land grading is necessary, however, to prevent
surface drainage problems which occur with rain, and to accommodate any special
equipment used;
- It is more difficult to have gross over-irrigation during months of peak ETcrop. This is
generally because many drip systems are not designed with a large system pump capacity;
- The Distribution Uniformity (DU) of new systems can be very high (0.93 or greater) in
reasonable terrain and with an excellent design;
- Systems can be installed on virtually any size or shape of parcel;
- Generally, there are no runoff problems to contend with;
- The systems are capable of high frequency irrigation. High frequency irrigation allows
the maintenance of optimum soil moisture content in the root zone, which is especially
important for salty water, or for shallow rooted crops. It should be noted that very high
frequencies are not necessarily optimum for some crops;
- Fertiliser can be directly applied uniformly to the root zone at any stage of growth on any
day with any dosage, without wetting plant foliage; and
- The upper portion of the root zone can be maintained moist, which enhances the uptake of
nutrients, such as phosphorus and ammonium that are typically concentrated near the soil
surface.

The above advantages are “typical”, meaning that they are not universal. As with any other
irrigation method, good operation depends upon good design, good equipment, and good
maintenance. Similarly, drip irrigation systems can have the following disadvantages:
- The DU can degrade quickly with time due to insufficient water filtration, lateral
flushing, and/or chemical injection. In other cases, the circumstance such as fresh water

77
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

clams growing inside the hoses, or an unusual insect that prefers to nest inside a certain
type of emitter. Rodent damage can be devastating in some areas;
- These systems are susceptible to damage by vandalism, and vandalism repairs may be
complicated, time consuming, and costly;
- Evaporation losses can be high with some drip designs that frequently wet large areas of
bare soil;
- Although the potential for excellent results (water savings, fertiliser efficiency,
optimisation of yield) often exist, they can only be achieved with excellent design and
excellent management. Often it takes several years for irrigators and farmers to develop
even average management skills, and catastrophic failure can result before those skills
are gained;
- Water must be available to the system on a very frequent and dependable basis. Drip
cannot be used in irrigation projects that deliver water on a rotation schedule (which
includes most of the world’s acreage) unless the fields are supplied with ground water;
- Energy costs for installation and operation of irrigation systems itself are usually higher
than for surface irrigation methods (assuming similar efficiencies) on flat ground;
- There are dozens of different types of essential parts (fittings, valves, etc.). The systems
must be supported by an excellent re-supply infrastructure;
- The initial cost of some forms of drip, in particular some permanent row/truck crops
systems, is among the highest of any irrigation method; and
- In very arid areas, a sprinkler system may be needed once every few years (and in some
cases more often, such as with some buried row crop drip) to leach salts which have built
up near the soil surface or for germination and establishing the crop.

Crops
Drip irrigation of any crop almost always requires a period of learning for a farmer who may
have grown that crop for many years under sprinkler or surface irrigation. Melon, for
example, seems to respond much quicker to drip irrigation than they do to furrow irrigation.
Care must be taken to not over-irrigate some such fruit crops, because they can quickly turn
vegetative.

Soils
Drip irrigation has been used on virtually all soils – from coral sands to heavy cracking clays.
The soil type will impact the number of emitters used per plant, as well as the decision to use
micro vs. drip. On trees in very sandy soils, micro would be the choice over drip because of
the limited lateral movement of water under drip (Figure 8.5).

Figure 8.5: Infiltration profiles from drip system for different soil types.

78
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Topography
Drip is found on virtually any topography. It is the most adaptive of irrigation methods in this
regard.

Water supply
There are two primary considerations regarding water supply for drip in addition to the
universal need for enough quality during the season- flexibility of water delivery and water
quality. Drip irrigation systems require the availability of almost continuous flow rates to a
field during the periods of peak ETcrop. As with sprinklers, the supply must be flexible
enough to adjust the hours of operation per day or per week at all times of the season to
match the ETcrop rate. Furthermore, if automation is contemplated, the water supply must be
available on true “demand”.

Reservoirs are often used to buffer the irrigation supply, whether it be an irrigation scheme
supply or from wells. The flow rate from well pumps may change during a year as ground
water levels fluctuate, yet a drip system needs a constant flow rate into a block. By
discharging the wells into a reservoir and then boosting the water out of that buffer reservoir,
the problem can be solved.

Salinity/Water quality
Drip systems can typically use saltier water than other irrigation methods because drip can
maintain the soil moisture at a high optimum water content, thereby reducing osmotic stress.
However, the emitters are very sensitive to solids concentrations in the water. Excessive and
expensive filtration is needed for dirty source water. In some cases, the water is so dirty that
reservoirs are needed to settle out sand and silt, or to oxidise iron in well water. Those
reservoirs serve as pre-filtration, prior to the regular filters. Drips system also are sensitive to
fairly (greater than 0.2 ppm) concentrations iron and manganese, as these can cause plugging
problems unless chemically treated.

Climate
One of the advantages of drip irrigation is that water is applied directly to the soil. Therefore,
wind has no effect on water distribution. Evaporation losses may be less than, equal to, or
greater than those found in other methods. They depend upon the frequency of irrigation, the
percentage of soil wet, the type of soil, and the location of the wet soil compared to shade.

Efficiency
Drip systems generally have application efficiencies that are higher than those for other
methods. This is partly due to the ability to schedule irrigations at almost any duration, and
partly due to the built-in flow rate limitation of most drip pump designs.

Irrigation scheduling
Drip irrigation systems are among the most simple to schedule accurately. They are capable
of applying both small and large amounts of water with the same uniformity, as compared to
surface irrigation methods and tend to have very poor DU with low application amounts.
Furthermore, the scheduling is very simple because the hours of irrigation are easy to adjust,
and the infiltrated amounts for most cases depend only upon the hours of application, not
upon the soil intake characteristics. The result is that the simplicity of scheduling typically
leads to savings in applied water amounts as compared to prior irrigation methods, even
though the DU may be the same.

79
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Some farmers unknowingly under irrigate because they expect large water savings, and the
annual irrigation efficiencies on drip irrigation fields are above 90% because of the small
deep percolation amounts and gross under irrigation on parts of the fields. A good water
management plan would recommend that these farmers improve the system DU and reduce
the under irrigation, and the desired annual irrigation efficiency would approach 80-85% or
so rather than some value greater than 90%.

8.3 Sprinkler Irrigation

In sprinkler irrigation system, water is applied to the soil using a pressurised piping system
with nozzles, jets, or perforated pipe that spray the water into the air. These sprinkler devices
or perforations are spaced to give a relatively uniform application of water over the field
being irrigated using a series of sets or a continuous move system. A variety of sprinkler
devices are available. These devices may be peculiar to a specific type of system but often are
adaptable to a number of systems.

8.3.1 Wetting pattern


The wetting pattern from a single rotary sprinkler is not very uniform (Figure 8.6). Normally
the area wetted is circular (Figure 8.6 b). The heaviest wetting is close to the sprinkler (Figure
8.6 a). For good uniformity several sprinklers must be operated close together so that their
patterns overlap. For good uniformity the overlap should be at least 65% of the wetted
diameter. This determines the maximum spacing between sprinklers.

Figure 8.6: Wetting pattern for a single sprinkler

8.3.2 Types of sprinkler systems

Sprinkler irrigation systems use one or more of the sprinkler devices (rotating head, low-
pressure spray nozzles, perforated pipe). Water is delivered to the sprinkler by a piping
system which may be hand portable, power moved, self-propelled or permanently installed.
Numerous ways to classify sprinkler systems have been developed. One method is to group
the system according to the method used to provide coverage to the entire field.

Hand move portable/Lateral move portable


A hand move portable system consists of one or more laterals (Figure 8.7). Laterals are
sections of pipeline with sprinkler heads installed at regular intervals along the pipe.

80
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Figure 8.7: Hand move laterals for sprinkler irrigation system

Perforated sprinkler lines may function as both lateral and sprinkler. The lateral pipe is
generally of aluminium with quick coupling connections at each pipe joint. Pipe section
lengths are generally 6, 9, or 12 m. The sprinkler lateral is “set” in one location until the
desired amount of water has been applied, the lateral line is then disassembled and carried to
the next “set” (Figure 8.6). When the lateral line set has been moved completely across the
field, it is disassembled and moved back to the starting location. When supply main is placed
in the middle of the field, one or more half laterals may be set on each side and rotated
around the field. Since only half of the water supply goes through each section of the main
line, locating the laterals at the opposite ends of the field in this layout makes it possible to
reduce the size of the main delivery line. Several half laterals may be used on long fields.

Rotating head impact sprinklers are generally used for agricultural crop irrigation, but newer
non-impact plastic sprinklers of the rotator design are gaining popularity, especially when the
pipe is configured in the “solid set” design for high value produce crops. The sprinkler head
is installed on the pipe riser so that it operates above the crop being grown. The risers may be
installed in the pipe coupling, but for ease in moving, are usually in the centre of a pipe
length. The length of the pipe joint is selected to correspond with the desired sprinkler
spacing. This system has a low initial cost but has a high labour requirement for carrying out
the irrigation. This type of system can be used on all types of topography and on most crops.
With some crops, such as maize, the difficulty of moving the lateral increases to an
impossibility as the crop reaches maturity. On bare sticky soils, moving becomes very
difficult because the irrigator must walk through the wet soil.

Centre pivot systems


The centre pivot system is a self-propelled system with the lateral supported from wheeled
towers spaced from 30 to 50 m apart (Figure 8.8). The towers are self-propelled so that the
sprinkler lateral rotates around an anchor, or pivot point, in the centre of the irrigated area.
Water is supplied to the lateral through the pivot point. Alignment controls cause the other to
move or stop to maintain proper system alignment. The towers have been propelled by water-

81
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

operated cylinders, revolving jets, electric motors, compressed air, or hydraulic motors.
Electric motors are by far most common, with some usage of hydraulic motors.

Figure 8.8: Centre pivot irrigation system

Generally, the speed of rotation can be varied from 12 to 120 hours per revolution for typical
400 m long lateral. The rate of water supply is constant depending on the pump rate and
design of the system, but the application rate increases with an increase in the length of the
lateral. The longer the lateral, the faster the end travels and the larger area the end irrigates
with each rotation. Consequently, the water application rate required at the end to cover the
area during one rotation is higher for longer laterals. If the system is not properly designed
and operated, the high application rate at the outer end of the lateral often causes runoff to
occur. The depth of water applied during one irrigation event depends on the speed of
rotation of the lateral. The most common operation is to apply 12 to 25 mm with 1 ½ to 3 day
rotation cycle. Since the lateral moves in a circle, the corners of the field are left un-irrigated.
Solid set sprinkler system, or drip irrigation can be used to irrigate the corners.

These systems have a moderate initial cost but have a low operating labour requirement.
They do require periodic maintenance and good water filtration, but beyond that they can be
very easy to operate and automate.

Solid set system


A solid set system is a system with a main line and laterals that remain in place during all or
part of the growing season of the crop (Figure 8.9). It requires enough mains and lateral pipes
to cover the entire irrigated area. If the main and laterals are buried and left in place from one
season to the next, the system is referred to as permanent solid set system. Where the system
is installed on the surface, the system is referred to as a portable solid set system. Portable
solid set systems are usually installed after the crop is planted and then removed just before
harvest.

Portable solid set systems are also used to germinate seed and establish a crop stand with
other methods of irrigation being used after the crop is established. After germination is
complete, the portable system may be moved to another field and the field is irrigated by
surface irrigation or by drip irrigation.

82
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Permanently installing the system, by burying the main and lateral lines underground,
facilitates farming operations while protecting the system from damage from equipment.

Figure 8.9: Solid set sprinkler irrigation system

8.3.3 Capabilities and limitations

There are so many different types of sprinklers systems that it is difficult to generalise about
advantages and disadvantages. However, some major advantages are:
- Irrigation scheduling is relatively simple, as the system dictates the application rate,
which is predictable;
- Because the sprinkler spacing, nozzle size, and pressure are typically fixed by the
design, there are very few management options. This means that the management
emphasis is more on maintenance and scheduling than on strategies for obtaining good
DU (such as with surface irrigation);
- Land grading requirements are minimal. This is extremely important in many
developing countries where the fields are small and good land grading is almost
impossible to achieve; and
- Labour requirement with some of the sprinkler irrigation systems, in particular centre
pivots, is very small.

Some key disadvantages of sprinklers are:


- They require more pumping energy than surface and drip irrigation methods;
- Sprinkler methods require better source water filtration than do the surface irrigation
methods, but less than drip;
- Some of the sprinkler methods (especially hand move) are labour intensive in terms of
physical exertion required. For this reason, it is sometimes difficult to find irrigators; and
- Sprinkler methods that apply water to leaves are unsuitable for irrigating if the water is
salty.

Crops
Nearly all crops can be irrigated with sprinklers. The characteristics of the crop, especially
the height, must be considered in selecting the type of system. There are some crops for
which sprinklers are not desirable, for example, onions for dehydration may discolour if their
tops are wet with sprinklers, and rot may increase on tomatoes.

83
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Sprinklers are sometimes used to germinate seed and establish ground cover for such crops as
grass and alfalfa. They are commonly used for pre-irrigations on soils with high initial intake
rates. Light, frequent applications may not be economical and are more difficult to effect with
surface irrigation methods, but can be easily managed under sprinkler systems. Crop damage
due to weather extremes, such as high temperature or frost, can be reduced. Special designs to
facilitate continuous application of water are required.

Soils
Most soils can be irrigated by some sprinkler method. However, sprinklers are not easily
suited to soils having less than 3 mm/hour final intake rate unless special measures are used
to increase intake or provide uniform surface ponding to control runoff.

Topography
Topography can vary from flat to fairly steep and rolling and still be suitable for sprinkler
irrigation by one of the several types of systems that are available. Land levelling is not
normally required. Some surface grading or smoothing may be required where surface
drainage of rainfall is needed.

In general, sprinkler irrigation can be used on any topography that is suitable for farming but
may be restricted by cultural practices, gullies, poorly drained spots, etc. Sprinklers are
applicable to soils on rolling topography or that are too shallow to permit surface shaping or
too variable to permit efficient use of surface irrigation method.

Water supply
Sprinkler irrigation systems require an essentially constant rate of water supply that is
available 24 hours per day for the least costly, most energy efficient designs. Water supplies
that are available on a rotation basis require excess system capacity or on-farm storage, both
of which add to the cost of the system. Although the water supply should be available 24
hours per day, it must be flexible in duration so that set times can be adjusted to the soil
moisture depletion.

Salinity/water quality
Salinity can be a factor in selecting the irrigation method. Salts can be leached from the soil
by sprinklers with high uniformity. Less water is required than with flooding methods as
water moves through smaller soil pores in an unsaturated condition. This is important in areas
with a high water table. The method used to leach salts is influenced by the soil and
topography.

Many crops are sensitive to foliar absorption of salts dissolved in the irrigation water.
Defoliation may take place where the water strikes the leaves. Under low humidity
conditions, water on the leaves and salts become more concentrated due to drying between
sprinkler rotations. Irrigating at night can help alleviate this problem.

Surface water supplies typically require some filtration to remove debris that could plug
orifices. The lower the sprinkler discharge, the finer the filtration requirement. To reduce
nozzle wear and lateral sedimentation, sediment-laden water require settling before entering
the system.

84
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Climate
Climate is a factor in selection of the method of irrigation. In the higher rainfall, higher
humidity areas, sprinklers are well suited to the supplemental application of water for crop
use. While very low humidity has an effect on sprinklers due to direct evaporation of the
spray before the water hits the soil, the most significant climate factor in selecting the method
of irrigation and the type of sprinkler system to be used is wind. Under strong winds severe
distortion of the application pattern will occur.

Efficiency
Efficiencies that can be attained with sprinklers depend on a number of factors. In practice,
the efficiencies range from very low to very high depending on the system selected the design
of the system and its operation. It is fairly easy to schedule irrigations with almost any type of
sprinkler irrigation method, with scheduling with centre pivots and solid sets being the
easiest. Evaporation losses vary, but can be in the 1- 20 % range, depending upon the nozzle
type, height of trajectory, climate conditions, extent of wetting of the crop canopy, and
numerous other factors. A well designed and properly operated sprinkler irrigation system
will have little or no runoff.

Irrigation scheduling
Irrigation scheduling is relatively simple with sprinklers, but it has fundamental differences
when contrasted to drip and surface irrigation methods. Even within the various sprinkler
methods there are fundamental differences.

Hand move sprinklers are typically irrigated for a known duration of 12 to 24 hours (less the
move time). Therefore, the application depth is determined in advance by the set duration.
Good scheduling involves waiting until the soil moisture depletion matches the actual
application depth (adjusted for application efficiency). This contrast with drip, for which the
application depth and time are adjusted to match the desired soil moisture depletion.

8.4 Buried clay pot irrigation

The buried clay pot is one of the most efficient systems of irrigation known and is ideal for
many small farmers. Buried clay pot irrigation uses a buried, unglazed, porous clay pot filled
with water to provide controlled irrigation to plants planted near it (Figure 8.10). Buried clay
pots can either be filled by hand if labour is inexpensive or connected to a pipe network or
reservoir.

Figure 8.10 Buried clay pot

The water seeps out through the clay wall of the buried clay pot at a rate that is influenced by
plant’s water use. This auto-regulation leads to very high efficiency, considerably better than

85
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

most drip irrigation systems, and up to ten times more efficient than conventional surface
irrigation. Buried clay pots are better than high technology drip systems in several respects:
- They are not sensitive to clogging as drip emitters, although they may clog over time
(three to four seasons) and require renewal by reheating the pots;
- Buried clay pots can be used without pressurised or filtered water systems;
- Buried clay pots can be made with locally available materials and skills;
- Buried clay pots are less likely to be damaged by animals or clogged by insects than drip
systems; and
- While even a brief interruption of water supply to a drip irrigation system due to a pump
or filter failure can lead to serious problems and costly damage to crops, the buried clay
pot systems may require water only once a week.

Buried clay pot irrigation should be considered wherever water conservation is important. It
will probably continue to prove most valuable for producing high value crops in dry lands.
Buried clay pot irrigation is also valuable for food production and re-vegetation of areas
affected by salinity or where only saline water is available for irrigation.

The size of the buried clay pot, the porosity of clay (a factor of clay firing), pot wall
thickness, water quality and the water demand of the plants determine the refill interval, but
for larger buried clay pots this can extend to a week or more. This is particularly helpful
where water must be carried from a distant river, well, or reservoir, and for remote fields. The
interval between refilling makes buried clay pot irrigation a time saver, and although small
farmers are often thought to have free time, they are usually time limited and any strategy
that can save time is desirable. It is also helpful when the farmer must be away for several
days at another job or task.

The drawbacks of buried clay pots include the cost of the clay pots, the energy required to
fire them, the time to install them, and less flexibility once they are installed. Buried clay pots
may have to be removed for tillage between crops. They may also clog over time and require
scrubbing and soaking or re-firing to clean out the pores. Although the clay is brittle,
breakage has not been a problem.

The most common mistake is placing the plant too far from the clay pot, outside the wetted
soil. Problems can also be caused if the clay has been over-fired and is not porous enough. If
a pot with bottom hole is not sealed well much of the benefits of buried clay pot will be lost.

Clay pot irrigation will prove most useful in helping farmers grow crops successfully in areas
with salinity problems (about 50% of the world’s irrigated land). The stable soil moisture
maintained by buried clay pot irrigation enables crops to be grown in very basic or saline soil
or with saline water under conditions in which conventional irrigation would fail. Very low-
fired pots may break up in very saline soils as a result of chemical reactions with the salts.

86
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 9- Irrigation water quality


Not all water is suitable for irrigation use. Unsatisfactory water may contain:
- chemicals toxic to plants or persons using the plant as food;
- chemicals which react with the soil to produce unsatisfactory moisture characteristics;
and
- bacteria injurious to persons or animals eating plants irrigated with the water.

Actually it is the concentration of a compound in the soil solution which determines the
hazard, and soil solutions are 2 to 100 times as concentrated as the irrigation water. Hence,
criteria based on salinity of the irrigation water can only be approximate. At the beginning of
irrigation with undesirable water no harm may be evident, but with the passage of time the
salt concentration in the soil may increase as the soil solution is concentrated by evaporation.
Free drainage of soil allows the downward movement of salts and helps to prevent serious
accumulations. Artificial drainage of soil may be necessary for this reason if natural drainage
is inadequate.

High salt concentration may sometimes be avoided by mixing the salty water with better
quality water from another source so that the final concentration is within safe limits. Rainfall
during the non-growing season will help to leach salts from the soil. It may, however, become
necessary to apply an excess of irrigation water so that deep percolation will prevent
undesirable salt accumulation in the soil.

A large number of elements may be toxic to plants or animals. Traces of boron are essential
to plant growth, but concentrations above 0.5 mg/L are considered deleterious to citrus, nuts,
and deciduous fruits. Even for the most tolerant crops a concentration of boron exceeding 4
mg/L is considered unsafe. Boron is present in many soaps and thus may become a critical
factor in the use of wastewater for irrigation. Selenium, even in low concentration, is toxic to
livestock and must be avoided.

Salts of calcium, magnesium, sodium, and potassium may also prove injurious in irrigation
water. In excessive quantities these salts reduce the osmotic activity of plants, preventing the
absorption of nutrients from the soil. In addition they may have indirect chemical effects on
the metabolism of the plant and may reduce soil permeability, preventing adequate drainage
or aeration. The effect of salts on the osmotic activity of plants depends largely on the total
salts in the soil solution. The critical concentration in the irrigation water depends upon many
factors; amounts in excess of 700 mg/L are harmful to some plants, and more than 2000 mg/L
of dissolved salts is injurious to almost all crops.

Most normal soils of arid regions have calcium and magnesium as the principal cations, with
sodium representing generally less than 5 % of exchangeable cations. If the sodium
percentage in the soil is increased to 10 % or more, the aggregation of soil grains break down
and the soil becomes less permeable, crust when dry, and its pH increases towards the
alkaline soils. Since calcium and magnesium will replace sodium more readily than vice
versa, irrigation water with a low sodium-adsorption ratio (SAR) is desirable.

Bacterial contamination of water is normally not serious from the irrigation viewpoint unless
severely contaminated water is used on crops which are eaten uncooked. Raw wastewater is

87
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

used for irrigation in many countries with limited water resources. Table 9.1 gives the
existing standards governing the use of treated waste water in agriculture in some countries.

Table 9.1: Existing standards governing the use of wastewater in agriculture


Crop type Country
California Israel South Africa Germany
Orchards and Primary effluent; Secondary Tertiary effluent; heavily No sprinkler
vineyards no sprinkler effluent chlorinated where Irrigation in the
irrigation possible; no sprinkler Vicinity
irrigation
Fodder, fibre Primary effluent; Secondary Tertiary effluent Pre-treatment with
crops, and Surface or effluent, screening and
seed crops sprinkler but irrigation of settling tanks; for
irrigation Seed crops for sprinkler irrigation,
Producing edible biological treatment
vegetables not and chlorination
permitted
Crops for For surface Vegetables for Tertiary effluent Irrigation up to 4
human irrigation, primary Human Weeks before
consumption effluent. consumption not Harvesting only
that will be For sprinkler to be irrigated
processed to kill irrigation, with treated
pathogens disinfected wastewater
secondary effluent unless it has been
(no more than 23 properly
coliforms disinfected
organisms 100ml) (<1000 coliform
organisms/100ml
in 80% of
samples)
Crops for For surface Not to be irrigated Potatoes and cereals
human irrigation, no with treated – irrigation through
consumption in more than 2.2 wastewater unless flowering stage
a raw state coliform they consists of only
organisms/100 ml fruits that are
For sprinkler peeled before
irrigation, eating
disinfected,
filtered
wastewater with
turbidity of 10
units permitted,
providing it has
been treated by
coagulation
Source: California State Department of Public Health (1968); Indian Standards Institute (1965); Israel, Ministry
of Agriculture, Water Commission (1969); Muller (1969); Peru, Ministry of Health, Department of
Environmental Sanitation (1970); Shuval (1976).

The Kenyan situation is given in the Environmental Management and Co-ordination (Water Quality)
Regulations, 2006. The microbiological quality guidelines for wastewater use in irrigation are given in the
EIGHTH SCHEDULE and standards for irrigation water in the NINTH SCHEDULE.

88
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Environmental Management and Co-ordination (Water Quality) Regulations, 2006


EIGHTH SCHEDULE

Microbiological quality guidelines for wastewater use in irrigation


Reuse conditions Exposed group Intestinal nematodes Coliforms (MPN/100 ml)
(MPN/L)*
Unrestricted irrigation Workers, consumers,
(crops likely to be eaten Public <1 <1000**
uncooked, sports fields,
public parks)
Restricted irrigation Workers
(cereal crops, industrial <1 No standard recommended
crops, fodder crops,
pasture and trees***

* Ascaris lumbricoides, Trichuris trichiura and human hookworms.


** A more stringent guideline (<200 coliform group of bacteria per 100 ml) is appropriate for public
lawns, such as hotel lawns, with which the public may come into direct contact.
*** In the case of fruit trees, irrigation should cease two weeks before fruit is picked and fruit should be
picked off the ground. Overhead irrigation should not be used.

Environmental Management and Co-ordination (Water Quality) Regulations, 2006


NINTH SCHEDULE

Standards for Irrigation Water


Parameter Permissible Level
pH 6.5-8.5
Aluminium 5 (mg/L)
Arsenic 0.1 (mg/L)
Boron 0.1 (mg/L)
Cadmium 0.5 (mg/L)
Chloride 0.01 (mg/L)
Chromium 1.5 (mg/L)
Cobalt 0.1 (mg/L)
Copper 0.05 (mg/L)
E.coli Nil/100 ml
Fluoride 1.0 (mg/L)
Iron 1 (mg/L)
Lead 5 (mg/L)
Selenium 0.19 (mg/L)
Sodium Absorption Ratio (SAR) 6 (mg/L)
Total Dissolved Solids 1200 (mg/L)
Zinc 2 (mg/L)

And any other parameters as may be prescribed by the Authority from time to time

89
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 10- Water Sources and Water Availability


10.1 Introduction
In each irrigation area, there is a demand for water and a supply of water. The demand for
water varies over time and depends on the types of crops, crop growth stages and on the
climate. While transporting irrigation water from the water source and applying it to the plant
roots, a portion of the water is lost through evaporation, leakage from the canals and
percolation below the roots of the crop. The scheme's gross irrigation need, SINgross includes
these water losses. To express the percentage of irrigation water that is used efficiently and
what percentage is lost, the term irrigation efficiency is used.

The supply of water to the irrigation system, SWS, should be almost if not equal to the
demand for water, SINgross, which is the total of the irrigation need for all fields plus the
water loss during transport and distribution.

When the supply of water exceeds the demand, plants may receive too much water which has
a negative effect on their growth. Or, on the other hand, costly water may be spilled and
disappear into the drainage system. When the supply is less than the demand, the irrigation
area may suffer from drought and plant production will decrease.

Extension agents and field technicians need to be able to advise farmers on how to handle
existing as well as anticipate future problems of water shortage. Problems may arise when an
irrigation area is extended, or when new crops are introduced. The technician therefore needs
to have an understanding of water shortage problems together with knowledge of the correct
approach, albeit simplified, to matching the supply with the demand for water. This chapter
deals with the supply aspects of irrigation. Various types of water sources are presented
which are largely what determine water availability. River and groundwater hydrology are
beyond the scope of this course, so that only a conceptual example is provided.

10.1.1 Water sources


The water needed to supply an irrigation scheme is taken from a water source. The most
common sources of water for irrigation include rivers, reservoirs and lakes, and groundwater.

10.1.2 Water availability


The possibility of supplying as much water to the irrigation area as is needed during each
period of the irrigation season depends primarily on the availability of the water at its source.
Availability may vary a lot over the year, or even between one year and another. Secondly,
the supply depends on the capacity of the facility installed to withdraw the water from the
water source. Further, technicians should be aware that water must be available during each
week or month of the growing season. This is illustrated with an example.

It is thus important to know how much water one can tap from a given water source over the
course of a season or a year, when:
- developing a new irrigation scheme;
- extending an existing scheme;
- changing the cropping season;
- adding a second or third irrigation season; and
- switching from a low water demand crop to a crop such as rice with high water
requirements.

90
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

EXAMPLE: During a five-month irrigation season a small river has the following discharges and the irrigation
area has the following INgross:
Qriver INgross the potential irrigable area
Month 1: 2000 l/s 5 l/s/ha 2000/5 = 400 ha
Month 2: 1000 l/s 2 l/s/ha 1000/2 = 500 ha
Month 3: 300 l/s 2.5 l/s/ha 300/2.5 = 120 ha
Month 4: 200 l/s 2.5 l/s/ha 200/2.5 = 80 ha
Month 5: 400 l/s 1.5 l/s/ha 400/1.5 = 267 ha
This example shows that in month 4 only 80 ha can be irrigated adequately, although there is much more water
available during the rest of the season. The month during which the ratio between the amount of water available and
the irrigation need is lowest (the smallest potential irrigable area) is called the critical month. In the example above
the critical month is month 4.

The scheme irrigation water supply depends, in technical terms, on two factors:
- whether the water source itself has a limited flow or a limited volume; and
- whether the facilities that control the tapping (gate, well, pump) and conveyance
(canal, pipe) have a limited capacity.

There may also exist water regulations which prohibit an unlimited withdrawal of water from
the source to ensure environmental flows downstream of the irrigation project intake.

10.1.3 Methods of tapping water from water sources


Water for irrigation has to be carried from the river, the reservoir, the lake or the groundwater
to the field. This can be done in two different ways:
- by making use of gravity. A simple gravity system can only be used if the water
level of the river or reservoir is higher than the level of the fields in the irrigation
scheme; and
- by using a pump to lift the water above ground level and then let it flow to the
fields.

The tapping of water from a river or reservoir by gravity, and the pumping of water from a
river, lake or groundwater are explained briefly.

10.2 Rivers

10.2.1 Introduction
Rivers are used all over the world as sources of irrigation water. The most typical quality
defining a river is that it flows, it is not a reservoir which contains a fixed amount of water.
At each moment a new amount of water is passing any given location along the river. The
flow of river fluctuates over time. The flows of some rivers fluctuate greatly over relatively
short periods of time; these are mainly small local rivers which respond quickly to rainfall in
their catchment area. A catchment area is the area from which a particular river or lake
receives both surface flow and drainage water originating from precipitation.

Other rivers show little fluctuation or vary only over a long period of time. These are mainly
rivers with a large catchment area, where the rains are spread over a greater area and for a
longer period of time. River flows vary considerably, not only within a given year, but also
from one year to the next. In a year with little rain during the rainy season, the river flow will

91
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

be small; sometimes the river flow will cease altogether during the dry season. The river flow
will be far more in years with heavy rainfall during the wet season.

The discharge of rivers is commonly expressed in cubic metres per second (m3/s). A simple
method of measuring the discharge in canals is described in Hydrometry Manuals. Although
the method is applied to canals, it can also be used for small rivers.

10.2.2 Tapping water from rivers

Direct river diversion


In some irrigation schemes water is supplied directly without any structures to divert water
from the river, dams or other devices. The offtake canal is directly excavated through the
river bank and no gate is provided. Consequently, there is no control of the discharge into the
canal. The diverted discharge depends on the water level in the river. That means the
discharge will be high during periods of high water and low during periods with a low water
level.

Usually, to avoid changes in the discharge, a control gate can be installed. The gate can then
be opened completely when the river level is low, and opened only partly when the water
level is high.

The advantage of this system of water abstraction is it’s relatively low cost. However, the
method can only be used when the river water level, even at low flows, is higher than the land
which is to be irrigated. In other words, the point of diversion must be located at some
distance upstream of the irrigation area, where the water level in the river is still higher than
the level of the irrigation area. If the area is very flat, the canal distance may be too long to
justify this simple river diversion system.

River diversion using a weir


To avoid the problems caused by fluctuating water levels in a river, a weir can be built across
the river. During periods of high river discharges, water will flow over the weir. The water
level upstream of the weir will show little variation during the year, and it will remain higher
during the dry season than it would without the weir. Since the water level upstream of the
weir is higher than it would normally be, the offtake structure can be closer to the irrigation
area. Consequently, the main irrigation canal can be shorter.

Pumping from a river


Some schemes are supplied with water which is pumped directly from the river. The cost of
pumping water increases with the size of the scheme. The cost of a weir, however, depends
largely on the size of the river and only partly on the size of the scheme. Therefore, pumping
is often used to irrigate small areas whose sources of water are larger rivers. Also, this may
be a much less expensive solution than building a weir. Pumping is essential if irrigation by
gravity is not possible. This will be the case when rivers lie well below field level.

10.2.3 Availability of river flow


It is important to make sure that there will be no lack of irrigation water. If water is in short
supply during some part of the irrigation season, crop production will suffer, returns will
decline and part of the scheme's investment will lay idle. Irrigation schemes depend on the
supply of water from the river throughout the irrigation season. The reliability of that supply

92
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

is an essential factor in the success of the irrigation scheme. Thus, once the SINgross of a
scheme has been estimated, the flow of the river and the reliability of that flow during the
irrigation season must be ascertained.

The availability of river water will fall into one of the three scenarios described below.
- The river flow is sufficient to meet the SINgross only when the flow is abundant.
Because the river flow fluctuates during the year, and over the years, water
shortages do occur when the flow is low during dry months of the year and during
dry years. It may be ascertained during which months of the year the river has its
lowest discharge. It is clear that under such circumstances any change in irrigation
practices resulting in higher irrigation needs during the dry months should be
discouraged;
- The river flow is abundant as compared to the SINgross even during dry months of
the irrigation season, but water shortages do occur in dry years. A change in
irrigation practices which requires more water, or small scheme extensions, may
result in more severe water shortages in the future; and
- The river flow is always much greater than the SINgross over the irrigation season,
and this remains true over the years. There has thus never been a problem for the
irrigation scheme as regards water availability. This can be checked with local
farmers. In this case, there will be no problem with small scheme extensions, or
with the introduction of crops, such as rice, which consume a lot of water.

The best solution is to consult the hydrological service or commission an hydrological study.
This service has information on water resources and river flows as they collect hydrological
data, such as rainfall and discharges of rivers and streams. They can estimate the minimum
river flow during the irrigation season based on historical records and predict the minimum
river flow for a dry year. If, however, no measurements of river discharge are available, then
data collection must start as soon as possible so that future plans may be made based on
reliable data.

It may happen that the river flow during one part of the irrigation season is limited, while at
other times during the season a large surplus of water is available. Part of this surplus can be
stored by constructing a dam in the river. The stored water can then be used in addition to the
available river flow. The cost of building a dam however is high, and expert help should be
sought to study alternatives and produce appropriate designs.

10.3 Reservoirs and lakes

10.3.1 Introduction
Lakes are natural depressions of the land which are filled up with water. Fresh water lakes
have a natural outlet through which the lake discharges superfluous water.

Lakes are supplied with water by rainfall that falls directly on the surface of the lake, by
water run-off from the adjacent land and small streams, or by groundwater that seeps through
the soil to the lowest point which is the lake. Lakes lose water via evaporation from the lake's
surface, via the lake's natural outlet (overflow), or through percolation from the bottom of the
lake to the groundwater.

93
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

A reservoir is an artificial lake. It can be formed by building a dam across a valley, by


excavating the land or by surrounding a piece of land with dykes. The water is stored in the
reservoir and can be used for irrigation.

Huge reservoirs are built by damming major rivers, which can supply water to large irrigation
areas measuring thousands of hectares. Small and medium sized reservoirs have a far more
modest capacity, with enough water to irrigate 10 to 100 ha in one season. This course
focuses on the latter type of reservoir.

10.3.2 Water availability in lakes and reservoirs

Natural lakes
What has already been explained for rivers, applies also to lakes. The amount of irrigation
water required must be less than the quantity of water which is available in the lake.

When lakes are supplied with water from rivers, their level may vary from one year to
another depending on the river. When lakes are supplied from groundwater, their water levels
fluctuate more slowly, following the level of the groundwater.

The amount of lake water available for irrigation may be calculated as the difference between
the water available in the lake and the amount lost due to evaporation and seepage. For large
irrigation schemes, local or national authorities will be responsible for evaluating the
available quantity of water. They will have data on river flows and groundwater supplies.
They can also estimate the evaporation and percolation losses and will have information on
other uses for the water.

Reservoirs
The amount of water that can be stored in a reservoir depends on:
- river discharge (which in turn depends on rainfall);
- height of the dam (when closing the river in a valley); and
- area of the reservoir (in the case of an artificially excavated storage area).

A dam built in a valley will block the river or stream and water can then be stored behind it.
The deeper and wider the valley, the more water that can be stored. This is called on-stream
storage.

Water can also be stored off-stream by diverting the flow into a natural depression or dead
river branch alongside the river, or by constructing a storage reservoir. This is similar to the
lake situation.

Evaporation from open water can easily reach 7 mm per day in arid or semi-arid countries.
This equals 5 cm per week and 20 cm per month. The amount of water lost by evaporation
can be considerable, particularly in reservoirs which are large and shallow. Therefore,
irrigation from shallow lakes and reservoirs should be started as soon as possible after the
rainy season. Wait too long, and the reservoir may well have dried up.

Because the cost of building and maintaining a reservoir can be significant, expert help
should be sought in planning and constructing a dam. Specialists can calculate the volume of
runoff water, anticipate water losses from the reservoir and estimate the quantity of water

94
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

available for irrigation. They should also be able to provide correct engineering designs and
sound construction techniques to ensure the stability of the dam, including spillway
arrangements to pass excess water safely downstream without harming the dam.

10.3.3 Water tapping from a reservoir


Water stored in a valley usually has a higher level than the valley bottom downstream of the
dam. Because of this difference in level, the valley can be irrigated by a gravity system.
Water can be taken from the reservoir via a concrete or steel pipe. This pipe connects the
reservoir to an irrigation canal downstream. A valve is usually located on the upstream end of
the pipe to control the discharge water into the canal.

Pumping from a lake or reservoir


The fields located around the reservoir upstream of a dam or surrounding a natural lake are
higher than the reservoir or lake's water table. Irrigation here will only be possible with the
help of pumps. The water level in the reservoir will usually be highest at the end of the rainy
season, and lowest at the end of the dry season or the irrigation season. Pumps installed at
reservoirs and lakes must be able to handle these fluctuations, which are not only vertical, but
which are even more pronounced horizontally, because the water recedes back to the lowest
parts of the reservoir.

A dead branch of a river can also be made to function as a reservoir. The branch is filled with
water during the wet season and closed off during the dry season so that the stored water may
be used. Due to the low water level, pumps will normally be needed to irrigate fields from
such a reservoir.

10.4 Groundwater

10.4.1 Introduction
Groundwater is an important source of irrigation water, especially for small-scale irrigation
projects. Because groundwater is only available below ground level, it must be lifted, or
pumped before it can be used. Pumping groundwater from wells is a well-known method of
utilizing groundwater the world over.

To understand how groundwater functions, think of it as a series of lakes below the surface of
the earth. The earth is built up of different layers - sand, gravel, clay, rock, etc. The layers of
rock or compact clay cannot store water as they are solid, rather than porous. The layers of
coarse sand and gravel, on the other hand, contain many pores and cracks, which allow
rainfall to enter the soil and percolate from the surface. These porous layers filled with water
are called aquifers. Groundwater flows, in most cases, slowly to the lower parts. Where the
aquifer meets the surface, the groundwater flows out of the soil into, for instance, a river or a
spring.

Groundwater may be found close to the surface or at profound depths. In coastal plains the
groundwater is often brackish or saline due to the proximity of the sea. Inland groundwater
may also be brackish in places where the soil contains many soluble salts. Such water cannot
be used for irrigation.

95
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

10.4.2 Groundwater availability


The availability of groundwater is less irregular than that of small rivers. Aquifers at shallow
depths, however, are likely to be very thin with a limited storage capacity for groundwater.
One may start pumping the water during the rainy season, after the aquifer has been
replenished, and find a dry well after a few months. Generally, local farmers know the wells
in their areas. When digging a new well, it is always wise to consult local farmers and benefit
from their experience of existing wells in the neighbourhood.

The total water availability from shallow groundwater is determined by the number of wells
and the capacity of these wells, or by the capacity of the pumps installed, whichever is
smaller.

When a borehole is driven to access deep groundwater, the engineer always carries out pump
tests to measure the capacity of the well. The volume of water stored in deeper aquifers is
quite large, so that the only limitation to water availability is the capacity of the well and the
pump to be installed. Excessive exploitation of the groundwater by many users, however, will
bring about a decline in the water table. To ensure a stable supply of groundwater, the rate of
use should not exceed the rate of recharge.

When pumps and other equipment break down, the resulting lack of water can be critical for
plant growth. Ideally, a spare pump should be kept on hand in case of breakdowns in the
main pumping system, as it can take a long time to repair or replace equipment.

10.4.3 Pumping from wells


To tap groundwater, a well must be dug deeper than the groundwater level. The groundwater
then seeps through the pores of the surrounding soil or porous rock into the well until the
level of water in the well is the same as the level of the groundwater.

Shallow groundwater
When the groundwater lies within a few metres of the surface, exploitation is possible with
shallow wells which are mostly dug by hand. These wells normally have a diameter of 1
metre or more and are dug below the groundwater table. Water is pumped from these wells,
often using human or animal power but, increasingly, with small diesel-powered pumps. The
amount of water that can be abstracted from shallow wells is limited, and, as a result, the
areas which are irrigated from these water sources will also be limited.

Deep groundwater
When the groundwater level is very deep, constructing a hand-dug well becomes impossible.
Deep wells must be drilled into the ground. Generally, submersible pumps are installed below
the groundwater table to lift water to the surface. They may be driven by an electric motor or
a diesel-powered engine on the surface with a long vertical shaft, or by a submerged electric
motor inside a waterproof casing.

96
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 11- Design of surface irrigation systems


11.1 General
In surface irrigation, water is released from a channel directly onto a farmer’s field. Drawn by
the pull of gravity, the inflow advances towards the downstream end of the field, slowing
down as the water infiltrates into the soil. When the inflow reaches the downstream end of
the field, it either stops (if the flow path is blocked), or it leaves the field as surface runoff (if
the flow channel is free draining). When the field has received enough water inflow is cut off,
which can happen either before or after the water front reaches the end of the field. Any water
remaining on the surface of the field can either infiltrate into the soil or flow away as surface
runoff. If the inflow travels along sloping flow channels, it starts to disappear from the
upstream end of the field while the drying front continues to travel downstream until all water
has disappeared. In theory, on a level field all the surface water disappears at the same
moment in time. Part of the water that infiltrates into the soil profile may percolate beyond
the root zone. Primary factors that affect the uniformity of the inflow distribution over the
length of a given field are flow rate and infiltration rate. Secondary factors are flow resistance
of the soil surface, field slope and geometry of the flow cross-section.

Surface irrigation is the oldest and most widespread way of providing agricultural crops with
water. There are several methods of surface irrigation, the most common being level basins,
borders and furrows for inflow to farmers’ fields. All these methods can be automated. All of
them can be adapted for use with a wide range of soils, topographies, crops and water
sources. Most of the farmers have developed the experience necessary for laying out and
using surface irrigation systems. Structures or devices at field level are generally
uncomplicated and easy to make from locally available material. Operating and maintaining
basins, borders and furrows, however, can require substantial investment of time and labour.
This is especially true of maintaining proper land levelling, which is necessary for achieving
high irrigation efficiencies and uniformities of irrigation application. Inadequate levelling
reduces the quality of irrigation dramatically by causing local over or under irrigation.

Table 11.1 summarises the different features and aspects of suitability of the main methods of
surface irrigation, as discussed in chapter 8. For more detailed information on determining
the appropriate method for a particular situation, the student is referred to standard irrigation
textbooks.

Table 11.1: A comparison of the main methods of surface irrigation


Factor Basin Borders Furrows
Flow section Wide, rectangular Wide, rectangular Various shapes
Infiltration Vertical only Vertical only Vertical and sideways
Slope Longitudinal slopes and cross- Slope length usually <0.5%, Slope length usually <1%, but
slopes zero or nearly zero but can go up to 4% in sod can go up to 3%
crops; avoid cross-slopes
Method of Constant flow Fixed flow, cutback, and Fixed flow, cutback, and reuse
operation reuse
Topography Flat, or uniform, gentle slopes Relatively flat, uniform Rather wide range of
slopes topographic conditions
Crops Most crops that tolerate some Most dry-foot crops Most row crops
inundation; narrowly-spaced
crops; orchards; rice
Soils Medium to fine textured Moderately low to High and low intake rates; not
moderately high intake rates on soils prone to puddling

97
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

11.2 Principles of surface irrigation

11.2.1 Hydraulic phases

To irrigate land is to supply it with water at a certain flow rate and for certain duration. As a
rule, surface irrigation has four hydraulic phases (Figure 11.1):

Figure 11.1 Surface irrigation process

The advance phase, which lasts from the start of the inflow until the time that the advancing
waterfront, reaches the downstream end of the field. The latter is known as the advance time.
As water advances over the soil surface some of it infiltrates into it. Infiltration causes the
flow rate and flow velocity to decrease, so the closer the water front gets to the end of the
field, the longer it takes to cover a unit of distance. This slowdown is reflected in the shape of
the advance curve from which the time that it will take the water front to reach a certain point
on the field can be read.

The ponding phase or storage phase, which lasts from the advance time until the time the
inflow is stopped because sufficient water has been supplied to the field. The latter is known
as the cutoff time. Other terms for cutoff time are “application time” and irrigation time”.
During the ponding phase, the water depth on the soil surface increases and, in open-ended
systems, surface runoff at the end of the field starts.

98
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

The depletion phase, which lasts from the cutoff time until that the upstream end of the field
falls dry. The latter is known as depletion time or the lag time. Surface water storage
decreases during the depletion phase.

The recession phase, which lasts from the depletion time until the time that the downstream
end of the field falls dry. The latter is called the recession time. The surface water disappears
gradually in a downstream direction. The recession curve shows when the water will
disappear from the surface at every point along the length of the field.

There may be differences in the duration of phases depending on the surface irrigation
method used. In level basins, for example, there is no surface runoff – an ideally – water
disappears from the entire field at the same moment purely as a result of infiltration. When
this happens, the depletion and recession phases coincide and the depletion (or recession)
curve is horizontal. In graded borders and furrows, the depletion phase is often short because
the water level at the field inlet is fairly shallow and drops due to a combination of infiltration
and continued surface flow. During the recession phase of graded borders and furrows,
infiltration and surface flow causes the water front to retreat further. In open-ended systems
this flow includes surface includes surface runoff at the lower end, whereas in blocked
systems the water level at the far end drops by infiltration only towards the end of the
recession phase.

Other differences in phase duration can occur when inflow at the upper end of the field is cut
off before the water front has reached the lower end of the field. In this case the cutoff time is
shorter than the advance time which means there is no clear ponding phase, so there may be
an overlap between the advance phase and the depletion and/or recession phases.

11.2.2 Process and purpose


The surface irrigation process can be summarised as follows:
- Water is put onto the field at certain flow rate and for a certain duration;
- Water leaving the inlet is divided between overland flow and infiltration;
- The spreading of water over the field depends on the combination of flow rate and
field characteristics (e.g. Field dimensions, flow resistance, slope);
- The flow rate reduces down the field because water infiltrates into the soil at a rate
depending on the soil infiltration characteristics;
- The shape of the advance curve is the result of the two processes of overland flow
and infiltration;
- Some of the flow may or may not run off after it reaches the end of the field,
depending on the surface irrigation method (basin, borders, furrows);
- The surface runoff and the infiltration together determines the shape of the
recession curve;
- The difference between the advance curve and the recession curve is the
infiltration opportunity time, which determines the final infiltration profile; and
- The final infiltration profile and the surface runoff together determine the level of
surface irrigation performance provided that it has reached the required depth at
the end of the field (Figure 11.1).

The purpose of surface irrigation is to get the water to infiltrate to certain target depth, the
required depth, uniformly distributed over the length of the field. The time needed for water
to reach the required depth is called the “required time” or “net opportunity time”. The actual

99
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

infiltrated depth at certain point along the field depends on the infiltration opportunity time,
which is the time when the water front arrives until it dries up. In Figure 11.1 it is assumed
that the actual infiltrated depth at the downstream end of the field equals the required depth,
consequently, the net opportunity time equals the recession time minus the advance time.
Therefore, to achieve infiltration to the required depth over the length of the field, designers
and operators should follow the guidelines below:
- Make the infiltration opportunity time and the net opportunity time as equal as
possible to ensure infiltration to the required depth;
- Make the advance curve and the recession curve as parallel as possible to ensure
equal infiltration (uniformity) over the entire length of the field; and
- Minimise the runoff at the end of the field.

These guidelines can be followed fairly closely by adjusting the amount and duration of the
inflow and by changing the field dimensions, assuming that other field characteristics cannot
be changed. The success of operation is usually measured in terms of adequacy, efficiency,
and uniformity of the irrigation application.

11.2.3 Method of operation


So far, it has been assumed that the flow rate until the cut-off time (Tco) is constant.
Applying a constant flow rate is a simple method of operation that requires neither extra
attention from the farmer nor automatic control of the flow rate. This method can, however,
make it difficult to realise good irrigation efficiency because it causes high surface runoff,
particularly in open ended borders and furrows. Somewhat more complicated methods of
operation to improve irrigation efficiency are cutback, reuse, and surge flow.

The cutback method


Runoff losses from open-ended furrows and borders can be reduced either by blocking the
lower ends of the furrows and borders or by reducing the inflow (cutback) at a pre-
determined moment. With the cutback method, irrigation starts at a selected flow rate and
continues at this rate for part of the application period. At a certain moment, the flow rate is
cut back to a level that is sufficient to complete the irrigation application adequately. In
furrow and border irrigation systems, it is usually assumed that this moment will come at the
end of advance phase. This is a fairly easy way to use the cutback method.

The purpose of cutback method is to reduce surface runoff and create more uniform
infiltration. Operating a surface irrigation system successfully with this method requires
either more involvement of the farmer or automatic control of the inflow, plus a thorough
understanding of when and how much to cut back the flow rate. A drawback is that the
cutback method is practical only when the remaining flow is used to irrigate other furrows
and borders.

The tail-water reuse method


The tail-water reuse (or recovery) method can be used only in open-ended borders and
furrows. Closed surface irrigation systems do not have surface runoff, and so there is no tail-
water to use for an extra irrigation application. The reuse method, like the cutback method,
aims to reduce total runoff losses. Surface runoff can either be collected in a ditch or sump
and then pumped back to the head end of the same field, or it can be used to irrigate borders
and furrows downstream. If it is intended to pump water back onto the same field, it should

100
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

be realised that considerable extra costs for the necessary equipment to do this will be
incurred.

The surge flow method


The surge flow method can be used if inflow at the head of the field is regulated
automatically, e.g., through gated pipes. With the surge flow method, cutback is not initiated
by reducing the flow size at certain moment, but rather by closing and opening the valves in
the gated supply pipes several times in succession. The advantage of this method is that if
you carefully assess the required application depth and then time the surges of inflow
accordingly, higher irrigation efficiency can often be achieved. One reason for this seems to
be the formation of a surface seal during the first cycle of surge and cutback, which reduces
the infiltration rate and allows the irrigation water front to advance towards the end of the
field more quickly. The surge flow method requires computer modelling or extensive field
experience to achieve optimum inflow rates and cycle times. This method is not widely used
because it is fairly complex and automatic regulation of the inflow system is costly.

11.3 Computer applications in surface irrigation

11.3.1 Mathematical modelling


The many variables and interactions involved in surface irrigation makes it a complex
process that is difficult to simulate quantitatively. For many years the only design aids for
irrigation engineers were tables and simple formulae, which gave only very rough guidelines.
The most prominent and most widely used manuals during the 1970s and 1980s were those
published by the United States Department of Agriculture, in particular the manuals on basins
and borders (USDA, 1974), followed by the manual on furrows (USDA, 1983). These
included many formulae and diagrams that were the accepted design standard in the United
States for a number of years.

This changed with the advances in computer programming in the 1970s and the spread of
personal computers in the 1980s. Mathematical modelling of surface irrigation became more
sophisticated, giving a more complete picture of the processes involved. So, irrigation
engineers started making computer programs of mathematical models that a wide variety of
users could apply without practical problems. As computer programs became more widely
available the previous approaches will became outdated. Indeed, some of them have already
lost their practical relevance.

11.3.2 Computer programs


The literature contains information on many mathematical models for surface irrigation, but
very few of these have led to workable, user-friendly computer programs. Those that have
and are currently available are listed below.
- BRDRFLW (Strelkoff 1985)- simulation of border and basin irrigation
- SRFR (Strelkoff 1999)- replaces BRDRFLW and gives the user the choice of
several calculations options
- SURDEV (Jurriens, M. et al., 2001)- for design, operation, and evaluation of
basin, border and furrow irrigation

101
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

11.3.3 Design, operation and evaluation


All of the computer programs mentioned in the previous section are useful in solving
problems of surface irrigation. The many possible issues that could be addressed ultimately
all have to do with design, operation or evaluation.

Design
Although design deals with a yet-to-be-built irrigation system, field data will depend on the
actual physical situation, giving the designer little room for manoeuvre. The site will be
known, so there will be some information available on the soil infiltration characteristics, the
crop that will probably be grown and the surface flow resistance. It is the combination of soil
and crop data with climatic data that will determine the most reasonable irrigation application
depth. If the data is insufficient, the designer must either collect more in the field or make
professional estimates. The same conditions, but with more choice for the designer, apply to
the parameters for field slope, furrow spacing and geometry. In other words, the field
parameters are given, assumed or estimated.

Design can then be focused on the determination of one or more of the decision variables:
flow rate, field length, field width and/or cutoff time. Depending on the situation, either the
flow rate is fixed and with the most of the available computer programs one can calculate
field dimension and cutoff time, or field dimensions are fixed and the computer programs can
calculate the flow rate and cutoff time. Performance indicators are always obtained as a
result.

Operation
Operation takes place in the existing situation, which means that all the relevant field
parameters including the field dimensions are known. Most of the published computer
programs can be used to calculate the remaining decision variables of flow rate and cutoff
time. Performance indicators are always obtained as a result.

The degree of freedom a farmer has in operating a surface irrigation system depends on the
local situation, which is dictated by the design of the system and the method of operation. If,
for example, the system is operated according to a strict rotational distribution, the full flow
rate and the total duration will be fixed and the farmer will only be able to operate within
these limits. He could decide to split up the flow for various units within his own farm and
juggle with the cut-off time per unit. In addition, the cutback and reuse methods of operation
can offer the farmer an extra measure of flexibility.

Evaluation
In this hand out, the term “evaluation” stands for assessment of irrigation performance, the
performance indicators being adequacy, efficiency and uniformity. In this case, evaluation
can be extended to cover quality of design and method of operation. The term is also used to
denote analysis of an existing situation for all known data: field parameters, field dimensions,
flow rate and cutoff time. Then, evaluation indicates the measure of the appropriateness of
that situation and the modifications that could improve irrigation performance.

11.4 Parameters and variables involved

In this section the various parameters and variables involved in the surface irrigation process
are discussed, focusing somewhat more on simulation aspects. The many variables involved

102
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

in surface irrigation are categorised according to whether they are field parameters, decision
variables or evaluation variables. Field parameters are situational data (i.e. data that describe
the field situation) and are not variables in the true sense of the word, because a design
engineer or farmer cannot assign them another value. Decision variables are those parameters
or variables that an irrigation designer or operator can adapt to find the best irrigation
performance for given or selected field parameters. Evaluation variables are basically
yardsticks for determining which combination of decision variables and field parameters will
produce the best irrigation performance.

11.4.1 Field parameters


Field parameters always include the infiltration characteristics, the surface roughness or flow
resistance and the required depth. In border and furrow irrigation, the gradient in the
downstream direction is also a field parameter, because the existing natural slope of the land
may suggest a field slope, which could be too low for water to advance quickly enough or too
steep to avoid erosion. A field choice may be restricted by the cost of land grading or the
orientation of the fields, for instance. Finally, in furrow irrigation, the shape and spacing of
the furrow are limited-choice field parameters. Furrow spacing is dictated mainly by
agronomic requirements; the furrow shape mainly depends on the available farm equipment
and local practice.

Soil infiltration characteristics


The objective of surface irrigation is to allow as much water as possible to infiltrate into the
soil to the required depth. Because infiltration largely controls the advance and recession of
the surface flow, the infiltration characteristics of the soil constitute a fundamental and
independent input parameter for many calculation methods and for all computer simulations.

Soil infiltration properties can be described in different ways. Over the years, the major
approaches were based on one-dimensional porous-media flow equations, or on physical
considerations, or on empirical relationships. For reasons of simplicity and minimum data
requirement, the last-mentioned category has been the most widely used, also in recent
irrigation models.

Here, only some of the most commonly used infiltration approaches are discussed, namely
the SCS and modified SCS intake families, and the time-rated intake families.

SCS intake families


To impose a limited number of standard classes on the widely varying infiltration coefficients
and exponents, the US Soil Conservation Service developed several “intake families” of soil
types according to soil texture and infiltration rate (USDA, 1974). These intake families can
be expressed numerically using the following equations:

D i  aT  c
b
(11.1)

b 1
I  abT (11.2)

Table 11.1 gives the SCS family numbers (0.05, 0.1, 0.15, and so on) and their corresponding
values for a and b according to Equation 11.1. The SCS family numbers approximate the
long-term infiltration rate in inches per hour according to Equation 11.2.

103
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

In Table 11.1, the data for basins and borders come from the SCS National Engineering
Handbook that was published in 1974 and the data for furrows come from later version
(USDA, 1983). The two sets of data show slight differences, the latter does not explain why.
The only clue in the chapter on furrow irrigation- which is the source of the data – is in the
statement that: “The intake curves developed for furrow or corrugated irrigation have the
same general shape as curves developed for the border and contour-ditch methods. The
furrow curves on a specific site are not necessarily the same as the border curves, but should
generally be parallel.”

Table 11.1 Values of a and b in Equation 11.1 for SCS intake families
Intake Basin/Borders Furrows
Family a (in/minb) a (cm/minb) b a (in/minb) a (cm/minb) B
0.05 0.2010 0.0681
0.10 0.0244 0.0620 0.661 0.0244 0.0620 0.6610
0.15 0.0276 0.6834
0.20 0.0306 0.0777 0.6988
0.25 0.0336 0.7107
0.30 0.0368 0.0935 0.721 0.0364 0.0925 0.7204
0.35 0.0392 0.7285
0.40 0.0419 0.1064 0.7356
0.45 0.0445 0.7419
0.50 0.0467 0.1186 0.756 0.0471 0.1196 0.7475
0.60 0.0520 0.1321 0.7572
0.70 0.0568 0.1443 0.7656
0.80 0.0614 0.1560 0.7728
0.90 0.0659 0.1674 0.7792
1.00 0.0701 0.1781 0.785 0.0703 0.1786 0.7850
1.50 0.0899 0.2283 0.799 0.0899 0.2283 0.7990
2.00 0.1084 0.2753 0.808 0.1084 0.2753 0.8080
3.00 0.1437 0.3650 0.816
4.00 0.1750 0.4445 0.823

Modified SCS intake families


The Kostiakov equation 11.3 shows the relationship between the cumulative infiltration depth
(Di) and the elapsed time (T) since the start of infiltration.

D i  kT
A
(11.3)

in which we see an empirical coefficient (k) and an empirical exponent (A). The derivative of
this equation (Equation 11.4) shows the relation between the instantaneous infiltration rate (I)
indicating the infiltration depth per unit time period at time T.
A 1
I  kAT (11.4)

It is possible to modify the SCS intake families for use with the Kostiakov equation 11.3, by
linearizing the original curves between two particular depths. Fangmeier and Strelkoff (1979)
made such a linearization based on infiltration depths of 50 and 100 mm. Table 11.2 shows
the results of these operations.

104
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Table 11.2 Modified SCS intake families for the Kostiakov equation for basins
Intake family Coefficient k A Soil texture
(in/hrA) (mm/minA) (mm/hrA)
0.1 0.49 1.10 12.57 0.595 Clay, silt clay
0.3 0.90 1.59 22.76 0.650 Silty clay, clay loam
0.5 1.25 1.93 31.76 0.684 Clay loam, loam
1.0 2.01 2.84 51.13 0.706 Loam, silt loam
1.5 2.65 3.56 67.32 0.718 Silt, sandy loam
2.0 3.25 4.22 82.46 0.726 Sandy loam, fine sand
3.0 4.32 5.41 109.68 0.735 Fine-to-medium sand
4.0 5.37 6.59 136.37 0.740 Medium-to-coarse sand

According to Walker (1989), it is also possible to modify the SCS intake families to fit the
Kostiakov-Lewis equation (Equation 11.3) for furrows. This has been done by determining a
different value of fo for each intake family and then calculating the values of A and k to equal
the values of the original SCS intake families (Garbi 1984, as mentioned in Walker 1989).
Table 11.3 shows the results of these operations.

Time-rated intake families


According to Merriam and Clemmens (1985), the SCS intake families in Table 11.1 do not
correspond well with field data on irrigated borders and furrows. They suggest using “time-
rated” intake families instead, which are based on the time it takes irrigation water to
infiltrate 100mm: a typical target irrigation depth. From their field data, they approximated
the value of the exponent A in the Kostiakov equation (Equation 11.3) as:

A  0 . 675  0 . 2125 log T i ( 100 )


(11.5)

where Ti(100) is the time (hrs) that it takes to infiltrate a depth of 100mm. You can get the set
of time-rated intake families of Table 11.4 by calculating A for different times T i(100) with
equation 11.5 and then inserting this A into the Kostiakov equation for Di=100 mm.

Table 11.3 Kostiakov-Lewis infiltration parameters for modified SCS intake families for
furrows and borders (after Garbi, 1984)
Intake family K (m/minA) A fo (m/min) Soil type
0.05 0.00426 0.258 0.000022
0.10 0.00383 0.317 0.000035 Clay
0.15 0.00360 0.357 0.000046
0.20 0.00346 0.388 0.000057
0.25 0.00337 0.415 0.000068
0.30 0.00330 0.437 0.000078 Clay loam
0.35 0.00326 0.457 0.000088
0.40 0.00323 0.474 0.000098
0.45 0.00321 0.490 0.000107
0.50 0.00320 0.504 0.000117
0.60 0.00320 0.529 0.000136 Silty loam
0.70 0.00321 0.550 0.000155
0.80 0.00324 0.568 0.000174
0.90 0.00328 0.584 0.000193 Sandy loam
1.00 0.00332 0.598 0.000212
1.50 0.00361 0.642 0.000280 Sandy
2.00 0.00393 0.672 0.000339

105
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Table 11.4 Time-rated intake families with k and A fitting the Kostiakov equation (Marriam
and Clemmens 1985)
Ti(100) Coefficient k Exponent A
Hrs Min mm/hrA mm/minA
0.5 30 167.0 8.10 0.739
1 60 100.0 6.31 0.675
2 120 65.5 5.37 0.611
4 240 46.8 4.98 0.547
8 480 36.6 5.07 0.483
16 960 31.3 5.63 0.419
32 1920 29.2 6.83 0.355

Obtaining field measurements of infiltration


Field measurements of infiltration can be helpful in determining the correct infiltration
characteristics of the soil, which is often a major problem in surface irrigation. The most
common method is to collect point data with a ring or double-ring infiltrometer, as described
in detail in various standard field-irrigation handbooks. This infiltrometer method enables
field data to be collected that give the values of the soil infiltration parameters. Using the
Kostiakov equation, data on infiltrated depths and time can be plotted on logarithmic scales,
making it possible to determine the values of A and k by regression.

The major disadvantage of an infiltrometer method is that it yields point measurements,


whereas one needs to know average values for the entire field. To obtain such values, data
can be collected from many points on the field and then determine average infiltration
parameters. Data can also be collected using the inflow-outflow method, which is suitable for
all irrigation methods, or by using the blocked-furrow or recycling-furrow-infiltrometer
method, which is specifically for furrows.

A more recent method and perhaps the most practical of all, is to collect advance data from
the entire field and then insert them into inverse use of a simulation model (including the
governing infiltration equations) to calculate the “average” field-infiltration parameters. The
type of data to collect and the processing thereof will depend on the irrigation method and
simulation model used.

Flow resistance
Flow resistance (n) is a basic input parameter in simulations of surface irrigation, which has a
direct effect on flow velocity and, consequently, on advance time, infiltration pattern and
total irrigation performance. The higher the flow resistance the longer the advance time. The
longer the advance time the more non-uniform the infiltrated-depth distribution.

Manning’s roughness coefficient n meant for steady uniform flow in canals is also commonly
used for surface irrigation. The 1974 edition of the SCS National Engineering Handbook
recommended values of Manning’s roughness coefficient for basins and borders (Table 11.5).
These have appeared in nearly all work on surface irrigation ever since.

Required depth
Required depth (Dreq) is another basic input parameter in simulations for surface irrigation.
The methods for determining this parameter is given in chapters 6 and 7 of this lecture notes,
the student is referred to the chapters for more information.

106
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Table 11.5 Recommended values of n (Manning’s coefficient) for basins and borders
n value Conditions
0.04 Smooth, bare soil surface; row crops
0.10 Drilled, small-grain crops, drill rows in flow direction
0.15 Alfalfa, mint, broadcast small grains
0.20 Dense alfalfa or alfalfa on long fields without secondary ditches
0.25 Dense sod crops and small grains, drilled perpendicular to flow directtion

Field slope and non-erosive velocity


For graded borders and furrows, the field slope (So) should not be too high (to prevent
erosion) or too low (to prevent slow advance). The engineer may have problems ensuring the
best field slope if the cost of land grading is high or if the orientation of the fields is
unfavourable. A relative flat slope may pose drainage problems in areas of high rainfall
intensity. For borders, the most suitable slopes are usually less than 0.5 %. But, if planted
with sod crops, slopes up to 4 % can also work well. For furrows, suitable slopes vary
between 0.05 and 1 %. Slopes up to 2 % can work for small furrows and corrugations.

Furrow spacing
In furrow irrigation, the spacing of furrows is limited-choice field parameter. It is usually not
the irrigation engineer who dictates spacing, but rather the agronomic constraints like the
required distance between crop rows and the width of farm implements. Furrow spacing
(Ws), the distance from centre to centre of two adjacent furrows, is in fact a dual-purpose
parameter. It is clearly a field dimension used primarily to convert volumes to depths
(D=Q/[LWs]), but it is also an input that assists in the modelling of infiltration process.

In simulations of flow in furrows, it is assumed that infiltration from the furrow spreads out
over the width of the furrow spacing and then is entirely vertical below a certain depth. W s is
used to convert the A and k values corresponding to the modified SCS intake families.

Furrow geometry
To simulate surface flow, there must be a means to relate the flow rate to the flow cross-
section. For basins and borders the situation is fairly simple, flow is assumes to be uniform
over the field width, therefore it is a two-dimensional process: only the head-to-tail direction
and the flow depth are involved. Infiltration happens only in a vertical direction and can be
calculated per unit width.

Furrows are a bit more complicated. Furrows are miniature channels in which the relation
between the top width of the flow section and the flow depth varies with the flow size and
with the various shapes the furrow channel can have. Moreover, infiltration takes place along
the wetted perimeter, which varies with the flow depth. Consequently, the geometry (or
cross-sectional) parameters of furrows require further scrutiny.

In simulations of flow in furrows, it is common practice to use Manning’s equation (Equation


11.6) to describe surface flow.
1 2 1
Q  AR 3
S0 2
(11.6)
n

in which Q is the furrow flow rate, So is the furrow slope, n is the flow resistance, A is the
wet cross-section, and R is the hydraulic radius.

107
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

11.4.2 Decision variables

Decision variables are those parameters or variables that a design engineer can manipulate to
find the best irrigation performance for given or selected field parameters. The decision
variables in surface irrigation are normally the field dimensions (length and width), the flow
rate and cut-off time. Of course, there are also limits to the values of these variables: the field
length has to fit in the existing infrastructure; the flow rate must be such that a farmer can
handle it; and the cut-off time can be limited to particular period (e.g., to the daylight hours).
The main design consideration in surface irrigation is usually the choice of the appropriate
combination of field dimensions, flow rate, and cut-off time.

Field dimensions
For basins and borders, the field dimensions are width (W) and length (L). For furrows, there
is only one field dimension: the furrow length. Furrow spacing is important only in the
context of field parameters. It should be realised that, in practice, it is not always possible to
choose an optimum field length. Field length is subject to limitations from the local
infrastructure (e.g., road, canals, slopes and so on). For basins, there is no best size or length.
If there is a given flow rate, obviously increasing the length of the basin will lower the
irrigation performance because the advance time will be longer. A higher flow rate, however,
will permit the designer to increase the length of the basin. Nevertheless, at a certain point,
increasing the flow rate will no longer enable the designer to extend the length of the basin.

The criteria for best field length are different for borders and furrows, at least when these are
open-ended. If a field with open-ended borders or furrows is too long, irrigation performance
will be low because the advance time will be long, causing uneven infiltration and
considerable losses to deep percolation at the upstream end of the field. If a field is too short,
surface runoff will be too high. So, there will be an optimum length for a field with open-
ended borders and furrows. It is the length for which the sum of deep percolation and surface
runoff losses is minimum, making the irrigation application efficiency maximum.

In simulations, field dimensions can be either the input to determine the required flow rate, or
the output of such a simulation, to assist a designer/evaluator who needs to know the best
field dimensions for a given flow rate.

Flow rate
For basins and borders, there is a total flow rate (Qo) for the field. The flow rate is divided by
the field width to obtain a unit flow rate (qo) per metre of width, and the result is used in the
theoretical analysis. For furrows, qo is the rate of inflow into one furrow, which is the unit
flow rate per width of one furrow spacing.

Flow rate is a key variable that affects the outcome of an irrigation event because it
influences the advance time of the inflow and, consequently, the irrigation uniformity,
efficiency, and adequacy. The flow rate should not be too high. For basins and borders, there
is a practical upper limit above which performance will not improve. For furrows, the flow
rate should not be too high to prevent scouring. There is an optimal flow rate (just as there is
an optimal field length) for open-ended borders and furrows. This is the rate at which the sum
of deep percolation losses and surface runoff losses is minimum.

108
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Inflow can enter borders and basins through a point inlet (one inlet structure), or there can be
a more uniform application of flow over the head end of the field by digging a head ditch
from which the field receives the water directly, either from overflow or from inlets or
siphons. If there is a point inlet, energy-dissipating measures may be necessary to avoid
erosion from large flows.

In existing situations, the flow rate is usually relatively easy to measure. Common flow sizes
(less than 50 l/s) can be measured using portable RBC flumes. Measuring larger flows in
basins and borders can be done with a custom-made portable flume. Siphons used in furrows,
borders or basins can be used to measure the flow.

In simulations, flow rate can be either an input (chosen by the designer/evaluator or dictated
by the existing supply conditions) or the outcome of one or more simulations that the
designer/evaluator makes while searching for the best flow rate.

Cut-off time
Cut-off time (Tco) is the amount of time that elapses from the start of irrigation to the cut-off
of the inflow. In simulations, cut-off time can be either input or output, as with the other
decision variables.

Cut-off for all three irrigation methods occurs usually sometime after the end of the advance
time so as to obtain infiltration to the required depth at the downstream end of the field. If
cut-off is substantially later than advance time, this will have a clear effect on the deep
percolation and surface runoff losses. If cut-off occurs too early, infiltration to the required
depth will often not happen at the end of the field. So, clearly, there are limits to the value
that can be chosen for the cut-off time, to achieve good irrigation performance.

Furthermore, cut-off time can be subject to practical restrictions. If, for example, there is a
strict rotational distribution within the tertiary unit, the cut-off time is fixed by the rotational
roster.

11.4.3 Evaluation variables

Evaluation variables are basically yardsticks for determining the combination of decision
variables and field parameters that will produce the best irrigation performance. In most
cases, the quality of an irrigation application is judged in terms of adequacy (i.e. whether
sufficient water was supplied to the field), efficiency (i.e. a relative measure of how much
water is “lost” during irrigation), and uniformity (i.e. the distribution of infiltrated water
depths over the length of the field). The primary irrigation performance indicators are:
storage efficiency, application efficiency and distribution uniformity. All performance
indicators are directly related to the pattern of infiltrated depths over the length of the field.

Water layer depths


As earlier mentioned, all the amounts of water is expressed in terms of water layer depths
(D). These is calculated as volumes divided by the field area (A), which for basins and
borders is length (L) times width (W) and for furrows is length (L) times furrow spacing
(Ws).

109
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

An important variable is the required depth (Dreq) or target depth, which is the depth of water
we want to store in the soil. This is not usually the same as the applied depth (Da),

Q o T co
Da  (11.7)
A

in which QoTco equals the applied volume (flow rate times cut-off time). The applied depth
infiltrates the field area either partly or entirely. This infiltrated part is expressed as average
infiltrated depth (Dav). Dav equals Da if there is no surface runoff, otherwise Dav < Da, and the
difference is the surface runoff depth (Dsr).

Even when Dav is more than Dreq, there may still be a part of the field (usually the lower end)
that is not properly wetted. The average depth of water that is actually stored in the target
zone Dreq is the storage depth (Ds). When the target zone is entirely filled, Ds will equal Dreq.
If Ds < Dreq, then there is under irrigation. On the other hand, some of the infiltrated water
may exceed Dreq. Then, the depth of the water is called the “deep percolation depth” (Ddp).
Infiltrated depths are expressed in terms of an imaginary water layer; they do not refer to
physical depth below the field’s surface. The real physical depth to which irrigation water
infiltrates is dependent on soil porosity and the initial soil moisture content as earlier
explained in chapter 6.

Adequacy
Adequacy of an irrigation turn is expressed in terms of storage efficiency (Es), which is
defined as the ratio between the storage depth and the required depth
Ds
Es  (11.8)
D req

The purpose of an irrigation turn is to meet at least Dreq over the entire length of the field. If
this purpose is achieved, then Ds=Dreq and Es =1. This means that the minimum infiltrated
depth (Dmin) will be equal to or larger Dreq. Figure 11.2 shows a situation in which the
purpose has not been achieved because Dmin < Dreq. The area with the cross-hatching
indicates the depth of water that was added to achieve Dreq (the numerator). Converting this
depth to an average depth over the entire length of the field, it is seen that it is equal to Ds. As
this is less than Dreq, so it follows that Es<1.

Figure 11.2 Storage efficiency

110
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Efficiency
The application efficiency (Ea) is a common yardstick of relative irrigation losses. It is
defined as the depth added to the target zone divided by the applied depth

Ds
Ea  (11.9)
Da

This definition is valid for all situations and all irrigation methods. Figure 11.3 shows the
application of this definition of Ea for various possible infiltration profiles. The differences
are related to the relative values of Dmin, Ds, Dav and Dreq.

Figure 11.3 Possible infiltration profiles

In Figure 11.3a (the design case), Dmin=Dreq so Ds=Dreq, consequently Ea=Dreq/Da, and Es=1.
In Figure 11.3b, Dmin>Dreq, meaning that there is over irrigation, while Es=1 and Ea= Dreq/Da.
For basins, Da=Dav and Ea=Dreq/Dav. For under irrigation, however, the situation is more
complicated because there are three possibilities. In all three cases, Dmin<Dreq and Ds<Dreq
(thus Es<1), but the relative values of Ds, Dav, and Dreq are different. In Figure 11.3c there is
so much deep percolation that Dav>Dreq, but in Figure 11.3d there is so little deep percolation
that Dav<Dreq. In Figure 11.3e there is under irrigation along the entire length of the field, and
the maximum infiltrated depth (Dmax) is less than Dreq, and Ds=Dav.

In fact, (1-Ea) indicates the fraction of the applied water that is “lost” (i.e., the fraction that is
not actually stored in the target zone). Such losses can be due to surface runoff and deep
percolation, for which the indicators below are used.

D sr
SRR  (11.10)
Da

111
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

The fraction of water that is lost to deep percolation is expressed by the deep percolation ratio
(DPR), which is defined as the deep percolation depth divided by the applied depth
D dp
DPR  (11.11)
Da

Because both of these indicators have the same Da in the denominator, it follows that (1-Ea)
=DPR +SRR or Ea = 1-DPR-SRR.

If there is runoff, the infiltrated fraction of the applied depth is (1-SRR). The literature gives
no indicator for the part of the infiltrated water that goes to deep percolation. DPR above
refers to Da and not to a lower Dav or Dreq.

Uniformity
To get a complete picture of an irrigation performance you need to know more than just the
indicators above, because these are averages taken over the entire length of the field.
Although different cases might produce the same results for Es and Ea their distribution
patterns could differ. One indicator used to represent the pattern of the infiltrated depths
along the field length is the distribution uniformity (DU), which is defined as the minimum
infiltrated depth divided by the average infiltrated depth
D m in
DU  (11.12)
D av

Because the distribution uniformity involves neither Dreq nor Da, it cannot give an indication
of irrigation adequacy and losses. Nevertheless, in level basins Dav=Da, and when
Dmin=Dreq=Ds, then DU=Ea.

Another frequently mentioned indicator for uniformity, namely, Christiansen’s uniformity


coefficient (UC), is defined as

n
 D i  D av 
UC  1   
 nD av


(11.13)
i 1  

For n points along the length of the field, UC gives the average of all differences between the
infiltrated depths and the average depth (Figure 11.4).

Figure 11.4 Christiansen’s uniformity coefficient

112
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

11.5 The SURDEV package

The SURDEV program package used in this course comprises of three programs: BASDEV
(for level basins), BORDEV (for sloping borders) and FURDEV (for graded furrows).
Together, they enable the user to design, operate and evaluate surface irrigation systems. All
three of three programs are published as ILRI Publication 59.

113
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 12 - Sprinkler Irrigation Design


12.1 Farm systems vs Field systems
The authors of textbooks only devote a few paragraphs to this topic, but it is one of great
importance. A complete understanding of the distinctions between farm and field systems
comes only through years of experience. Variability in design, operation and management
conditions is limitless.

“A poorly designed system that is well managed can often perform better than a well
designed system that is poorly managed”

- Farm systems may have many fields systems;


- Planning considerations should include the possibility of future expansions and extra
capacity;
- Permanent buried mainlines should generally be oversized to allow for future needs-it is
much cheaper to put a large pipe in at the beginning than to install a secondary or larger
line later;
- Consider the possibility of future automation;
- Consider the needs for land levelling before burying pipes;
- How will the system be co-ordinated over many fields?
- What if the cropping patterns change? (tolerance to salinity, tolerance to foliar wetting,
peak ET rate, root depth, need for crop cooling or frost protection, temporal shifting of
peak ET period);
- What if energy costs change?
- What if labour availability and or cost change?
- What if the water supply is changed (e.g. from river to groundwater, or from old well to
new well)?
- What if new areas will be brought into production?

12.2 Outline of sprinkler design procedure


1. Make an inventory of resources – visit the field site personally if at all possible, and talk
with the farmer, get data on soil, topography, water supply, crops, farm schedules,
climate, energy, etc. be suspicious of parameter values and check whether they are within
reasonable ranges;
2. Calculate a preliminary value for the maximum net irrigation depth, dx;
3. Obtain values for peak ETc rate, Inet (average crop water requirement during the peak-use
period) and cumulative seasonal ETc,season In,season;
4. Calculate maximum irrigation frequency, (fx,=dx/Inet) and nominal frequency, f’ (is value
of fx rounded down to the nearest whole number of days)-this step is unnecessary for
automated fixed systems and center pivots;
5. Calculate the required system capacity, Qs- first, calculate gross application depth, d;
6. Determine the “optimum” (or maximum) water application rate - application rate is a
function of soil type and ground slope;
7. Consider different types of feasible sprinkle systems;
8. For periodic-move and fixed (solid-set) systems:
- Determine nozzle and lateral spacing, nozzle discharge (qa), nozzle size, and
pressure (P) for optimum application rate,
- Determine number of sprinklers (N= Qs/qa) to operate simultaneously to meet Qs,
- Decide upon the best layout of laterals and mainline,

114
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

- Adjust f, d, and Qs to meet layout conditions,


- Size the lateral pipes,
- Calculate the maximum pressure required for individual laterals,
9. Calculate the main line pipe size(s);
10. Adjust mainline pipe sizes according to the “economic pipe selection method”;
11. Determine extreme operating pressure and discharge conditions;
12. Select the pump and power unit; and
13. Draw up system plans and make a list of items with suggestions for operation.

12.3 Summary
- Note that p (depletion factor) is not a precise value; actual precision is less than two
significant digits- this justifies some imprecision in other values (don’t try to obtain very
precise values for some parameters when others are only rough estimates);
- When determining the seasonal water requirements we subtract effective dependable
rainfall (Peff) from crop water requirement (ETc). To be safe, the value of Peff must be
reliable and consistent from year to year, otherwise a small (or zero) value should be
used;
- Note that lateral and sprinkler spacing are not infinitely adjustable: they come in standard
dimensions from which designers must choose. The same goes for pipe diameters and
lengths; and
- Note that design for peak (ETc- Peff)peak may not be appropriate if sprinklers are used only
to germinate seeds (when later irrigation is by a surface method).

12.4 Design Example


Given: Alfalfa crop. Top soil is 1.0 m of silt loam, and subsoil is 1.8 m of clay loam. Field
area is 35 ha. Depletion factor is 0.5 and ECw is 2.0 dS/m. Application efficiency is estimated
at 78%, and the soil intake rate is 15 mm/hr. Lateral spacing is 15 m and lateral length is 400
m. Assume it takes 30 minutes to change sets. Seasonal effective rainfall is 190 mm and
climate is hot. Inet=7.6 mm/day, and ETcseason=914 mm/season. Top soil Sa=167mm/m and
sub soil Sa=183mm/m and root depth Zr=1.8m. Salinity for 10% yield reduction gives
ECe=3.4 dS/m.

Solution:
1. Average water holding capacity in root zone: top soil is 1.0m deep; and root zone is 1.8m
deep. Interpolate - Sa=(1.0(167)+(1.8-1.0)(183))/1.8 = 174 mm/m

2. Max net application dx=pSaZ =0.5(174)(1.8) = 157 mm

3. Maximum irrigation interval fx = dx/Inet = 157mm/7.6 mm/day =20.6 days

4. Nominal irrigation interval (round down) f’ =INT(fx) =20 days

5. Net application depth: dn=f’.Inet=(20 days)(7.6mm/day) =152 mm

6. Operating time for an irrigation: f =f’- days off: 20 days is almost three weeks, and
depending on which day is off, there could be 3 off days in this period (for breakdowns
and rest). So, with one day of per week, we will design the system capacity to finish in
20-3 =17 days. Thus, the actual days of operation, f=17 days. But, remember that we still

115
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

have to apply 20 days worth of water in these 17 days (we irrigate 6 days/week but crop
transpires 7 days/week)

7. Leaching requirement LR= ECw/(5ECe-ECw)=2.0/(5(3.4)-2.0) = 0.13

8. Gross application depth d = dn/((1-LR)Ea) = 152/((1-0.13)0.78) = 225mm

9. Nominal set operating time: with 225 mm to apply and a soil intake rate of 15 mm/hr, this
gives 15 hours minimum set time (so as not to exceed soil intake rate). Then we can make
the nominal set time equal to 16 hours for convenience. With 0.5 hrs to move each set,
there are a total of 15.5 hrs/set of water application, and the farmer can change at 0600
hrs and 2200 hrs on day 1 and at 1400 hrs on day 2 (for example). The actual application
rate will be 225/15.5 =14.5 mm/hr (<15 mm/hr, therefore OK). At this point we could
take the lateral spacing, Sl, sprinkler spacing, Se, and actual application rate to determine
the flow rate required per sprinkler.

10. Sets per day: From the above, we can see that there would be three sets each two days

11. Number of sets per irrigation cycle: (17 days/irrigation)(1.5 sets/day)=25 sets per
irrigation cycle

12. Area per lateral per irrigation: Lateral spacing on mainline is Sl=15m. Lateral length is
400m. Then, the area per lateral is: (15m/set)(25 sets)(400 m/lateral) = 15ha/lateral

13. Number of laterals needed: 35 ha/(15ha/lateral) = 2.3 laterals. Normally we would round
up to the nearest integer, since 2 laterals would not quite cover the field. Three (3) laterals
will provide increased safety and flexibility.

14. Number of irrigations per season:


(ETcseason-Peff/dn)=(914 mm-190 mm)/152 mm/irrigation =5 irrigations events;
Thus, there would be approximately five (5) irrigations in a season. Note that this is a very
general value; since the value of ETcseason is very general.

15. System flow capacity- with 15.5 hours operating time per set and 1.5 sets per day, the
system runs about 23 hrs/day.
Qs=2.78((A×d)/(f×T))=2.78((35 ha)(225mm))/((17 days)(23 hrs/day))=56.0 lit/sec. This
is assuming no rainfall during peak ET period.

16. Given the critical head, Table 15.1 can be used to size the laterals and the main line and to
determine the actual head that the pump is to work against.

12.4 Exercise
Given:
Mean ETo (mm/decade) and historic rainfall data (mm/decade) (1972-2003) for Eldoret
Meteorological Station. Crop to be grown is cabbage under intensive management and start
of irrigation season is early October. The soil in the project area is silty clay (θfc =44.7 vol%
and θwp =25.7 vol%). Water source is from a stream with reliable flow and low turbidity. The
net field dimension is 100 m ×60 m, the field length is parallel to the stream. The field level

116
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

is 1.5 m above the water level in the stream, the distance from the edge of the stream to the
field is 3 m and field has very minimal slope.

Asked: Design an appropriate sprinkler irrigation system including the conveyance system
and appropriate pump. Use information given in Table 12.1 and Figure 12.1.

Table 12.1: PVC and GI Friction loss tables


Head loss in metres per 100m for different classes of PVC and GI pipes
Flow 3/4 1 1 1/4 1 1/2 2 2 1/2 3
(m³/h)
PVC GI PVC GI PVC GI PVC GI PVC GI PVC GI PVC GI
D E D E D E C D E C D E C D E B C D E
1.5 2.4 3.2 8.0 1.9 1.5

2 5.0 6.8 17.8 1.6 2.2 4.2 2.6

2.5 8.6 11.5 31.9 2.8 3.8 7.5 1.3 1.8 3.9 1.5

3 12.9 17.4 49.8 4.2 5.7 11.6 2.0 2.7 5.7 2.3

3.5 19.0 25.8 71.7 6.2 8.4 17.7 2.7 3.7 8.0 1.2 1.4 1.9 3.3

4 8.5 11.6 25.2 3.5 4.7 10.3 1.5 1.8 2.4 4.3 1.0

5 10.8 14.8 32.7 4.9 6.7 15.5 2.2 2.6 3.5 6.3 1.5

6 15.5 21.3 50.5 6.9 9.3 21.7 3.1 3.6 4.9 9.0 1.0 1.2 1.6 2.1

7 21.6 29.6 72.7 9.2 12.5 29.3 4.1 4.7 6.5 12.3 1.4 1.6 2.2 2.9

8 28.8 41.8 98.9 10.5 15.6 38.8 5.1 5.9 8.1 15.6 1.7 2.0 2.7 3.7 1.2

9 14.4 19.6 49.1 6.4 7.5 10.2 20.0 2.1 2.5 3.4 4.7 1.6

10 17.5 23.8 60.7 7.8 9.0 12.4 24.6 2.6 3.0 4.1 5.8 1.9

12 33.3 87.4 10.8 12.6 17.3 35.3 3.6 4.2 5.8 8.4 1.2 1.4 1.9 2.7 1.1

14 14.3 13.9 22.9 48.3 4.8 5.6 7.6 11.5 1.6 1.8 2.5 3.7 1.5

16 18.3 25.9 29.3 63.0 6.1 7.2 9.8 16.2 2.0 2.3 3.2 5.0 2.0

18 7.6 8.9 14.4 20.9 2.5 2.9 4.0 6.3 1.0 1.0 1.3 1.8 2.5

20 9.2 10.8 18.7 25.6 3.0 3.5 4.9 7.6 1.2 1.4 1.6 2.2 3.1

2 2½ 3 4 6
PVC GI PVC GI PVC GI PVC GI PVC GI
C D E C D E B C D E B C D E B C D
20 9.2 10.8 18.7 25.6 3.0 3.5 4.9 7.6 1.2 1.4 1.6 2.2 3.1

25 13.9 16.2 28.7 37.3 4.5 5.3 7.4 12.0 1.8 2.1 2.4 3.3 4.8 1.1

30 6.4 7.5 10.4 17.2 2.5 2.9 3.4 4.7 6.9 1.7

35 8.5 9.9 13.8 23.2 3.4 3.8 4.5 6.2 9.3 1.0 1.1 1.3 1.8 2.3

40 10.9 12.5 17.7 30.3 4.3 4.9 5.8 8.0 12.2 1.2 1.4 1.7 2.3 2.9

45 13.6 16.3 22.0 38.2 5.4 6.1 7.2 9.9 15.4 1.5 1.8 2.1 2.9 3.7

50 7.5 9.2 12.0 19.0 1.9 2.2 2.5 3.5 4.6

60 10.7 13.1 17.2 27.6 2.7 3.1 3.6 5.0 6.5

70 14.3 17.8 23.0 37.3 3.6 4.1 4.8 6.7 8.8 1.1

80 17.8 23.3 28.8 48.7 4.5 5.2 6.1 8.4 11.5 1.5

90 5.6 6.5 7.6 10.9 14.6 1.9

100 6.8 7.9 9.2 13.5 18.0 1.0 1.2 1.4 2.4

120 9.5 11.0 13.6 19.5 25.9 1.4 1.7 1.9 3.4

140 1.9 2.2 2.6 4.7

160 2.4 2.8 3.3 6.3

180 3.0 3.5 4.1 7.9

200 3.6 4.2 5.0 9.5

225 4.5 5.3 6.2 11.4

250 5.5 6.4 7.8 15.0

117
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Figure 12.1- Pump performance curve

118
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 13: Drip Irrigation System Design

13.1 Basic Guidelines

13.1.1 Application uniformity


A fundamental requirement in design of an irrigation system is to achieve uniform water
application in concurrently irrigation blocks having the same crop and its age, phonological
phase and spacing. Since 100% uniformity in flow rate of emitters is never attainable, by
convention, a deviation of 10% between maximum and minimum of emitters’ flow rate is
acceptable.

13.1.2 Peak water consumption


The system capacity should match with the crop water requirement. The pipe network should
enable the matching of water application to seasonal changes in consumption.

13.1.3 Durability
The selected system components as well as their installation process should guarantee their
reasonable durability.

13.1.4 Economic considerations


Economic considerations should be taken in to account in choosing the equipment and the
layouts. Both, the annual return on the initial investment and the long term current annual
expenses have to be taken in to account.

13.2 The Design Procedure


The preliminary design process comprises following steps.

13.2.1 Selection of emitter type and layout


In the preliminary planning phase, peak daily and hourly water demand, irrigation intervals
and irrigation depth are calculated and fertigation regime are determined. Following which
the design process starts. In design phase, the first step is to choose an emitter and an optional
system layout. The spacing between laterals and between the emitters on a lateral should
commensurate with the crop spacing and water conductivity attributes of the soil.

13.2.2 Checking alternative layouts


When designing an irrigation system, it is imperative to analyse several alternatives like –
comparing initial investment cost, labour and energy expenses. A pipe network (main, sub-
mains, manifolds, and laterals) can be laid out in six configurations viz. (i) comb type, (ii)
splitted comb type, (iii) central fish bone type, (iv) asymmetric fish bone type, (v) splittes fish
bone type, and (vi) dual fish bone type, as shown in Figure 13.1.

Selection of an optional layout depends on diverse and occasionally contradicting


considerations. Longer laterals may enable shorter mainlines and savings of accessories but
oblige the use of wider diameters. The comb layout saves outlets but necessitates the use of
wider diameter of the distribution line. In many cases, it is economically favourable to lay
manifolds of narrower diameters that reduce the cost of accessories (Figure 13.2). Manifolds
help simplify operation and automation of the system and splitting of the field for separate
water applications.

119
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

120
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

In the primary design phase of the system, the head losses created by the water flow in the
pipes and fittings are calculated. There are diverse procedures for calculating head losses. In
the past, designers used tables, slide rulers and nomograms. Dedicated software for irrigation
design and on-line calculators replaced those old fashion procedures.

Emitter manufacturers recommend the maximum allowable length of drip and micro-emitter
laterals on flat and slopping lands, keeping emitters’ flow rate variance within 10% (± 5% of
the average). The design process is divided into two phases:
(i) head losses are calculated from the distal end to the head of the plot, using nominal
values of head and discharge; and
(ii) the design is checked and adjusted, going from the head to the distal end.

At this stage, the calculation relate to accurate data. The plot is divided into sectors and
detailed calculations are performed for each pipe segment. The data is recorded in the design
form.

Head losses in accessories can be calculated using the concept of equivalent length. The data
manipulated with this concept presents head losses in a virtual pipe of the same diameter as
that of the accessory. Most manufacturers provide tables and nomograms of head losses in
their products. Local head loss in an accessory can also be calculated, using its flow factor
(Kv), if available.

13.2.3 Water flow velocity


Water flow velocity determines the head losses in the system. As flow velocity increases, the
head losses are amplified. The flow velocity over 2.5 m/s leads to higher head losses – which
in turn cause higher energy losses, consequently reducing the economic viability of the
system. In mainlines high velocities may trigger water hammer that may result in the pipe
bursting. Hence in the preliminary phases of design, the expected velocity in manifolds is
kept within the range of 2 – 2.5 m/s and in mainlines below 1.5 m/s.

13.2.4 Spacing of laterals and emitters


Spacing between laterals and between emitters on the lateral is dictated by the spacing
between rows and between plants in the row, soil depth, its texture and the characteristics of
root system.

13.2.5 Choosing emitters and laterals


Selection of emitter (inline, online, splitted, etc.) and lateral type is related to the cultivation
technology. The desired durability of the system determines the choice between thick-walled
and thin-walled laterals. The topography plays a role in choosing of pressure compensating,
semi compensating or conventional emitters.

The selected emitters and their flow rate will correspond with spacing, planned irrigation
regime, soil permeability and the capacity of water supply. Additional considerations are –
crop response to water distribution patterns and climate control requirements. The actual
design is separated into two phases:
(i) Layout of laterals and manifolds; and
(ii) Layout of the mainline and the sub-mains.

121
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

13.3 Design of drip irrigation system for row crops


In row crops, the layout should correspond with crop spacing, soil texture and topography.
Example: Design of a retrieval drip system for irrigation of maize on flat ground.

(a) Basic data


Crop Maize,
Length of growing season – 135 days,
Root system depth – 1 m.
Plot dimensions 420 × 200 m.
For harvesting convenience, plot is splitted in the
middle and row length is 100 m on each side.
The plot is partitioned to 7 blocks, 60 m wide each.
Soil Sandy loam
Available water – 12 vol%
Allowable depletion in peak season – 60%
Topography Flat ground – 0% slope
Spacing 100 cm between rows,
30 cm within a row
Peak season daily gross water 7 mm
requirement
Water source Reservoir, pumping required.

Figure 13.3 shows a retrievable drip irrigation system layout for example maize field.

In sandy loam, 30 cm distance between drippers on lateral (corresponding to plant spacing),


is the favoured choice. In this case it guarantees one emitter per each plant.

Firstly, choice of drippers and laterals has to be made. Lying laterals of 100 m long on both
sides from the middle, represents 333 drippers per lateral. In retrievable on-surface drip
irrigation the favoured emitter is the internal integral dripper which is least prone to damage
during deployment and retrieval of laterals. Four types of integral drippers can be considered:
(i) Pressure compensating dripper in thick-wall lateral;
(ii) Pressure compensating dripper in thin-wall lateral;
(iii)Non-pressure compensating dripper in thick-wall lateral; and
(iv) Non-pressure compensating dripper in thin-wall lateral.

The technical data of applicable drippers is given in Table 13.1.

Table 13.1: Compensating dripper (compensating pressure threshold – 4 m) data


Model OD (mm) Wall thickness (mm) ID (mm) PN (bars) Kd
16012 16.10 1.2 13.70 4.0 1.6
16010 16.10 1.0 14.10 3.5 1.3
16009 16.10 0.9 14.20 3.0 1.2
OD = outer diameter, ID = inner diameter, PN= working pressure, Kd = coefficient of disturbance to flow

Table 13.1 demonstrates the difference in Kd between model 16012 (wall thickness 1.2 mm)
and model 16009 (wall thickness 0.9). The greater is the inner diameter, the lower is the Kd,
enabling use of a longer lateral. On the other hand, the thicker wall latter allows for higher
working pressure and longer durability.

122
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Table 13.2 and Table 13.3 give the maximum allowable lateral lengths for two IDs (13.7 mm
and 14.2 mm), different dripper spacing and nominal flow rates.

Table 13.2: Maximum allowable lateral length using compensating drippers (mm)
Model 16012, ID = 13.70 mm, Inlet pressure = 3.0 bars
Nominal Spacing between drippers (m)
flow rate 0.20 0.30 0.40 0.45 0.50 0.60 0.75 0.90 1.00 1.25 1.50
(lit/hr) Maximum lateral length (m)
1.2 104 151 195 216 237 277 333 386 420 500 575
1.6 86 125 162 179 196 229 276 320 348 415 477
2.3 68 98 127 141 155 181 218 253 275 328 378
3.5 51 75 96 107 118 137 166 198 209 250 288

123
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Table 13.3: Maximum allowable lateral length using compensating drippers (mm)
Model 16009, ID = 14.20 mm, Inlet pressure = 3.0 bars
Nominal Spacing between drippers (m)
flow rate 0.20 0.30 0.40 0.45 0.50 0.60 0.75 0.90 1.00 1.25 1.50
(lit/hr) Maximum lateral length (m)
1.2 118 170 219 242 264 307 368 426 462 548 627
1.6 98 141 181 200 219 255 305 353 383 455 521
2.3 77 111 143 158 173 201 241 279 303 360 413
3.5 58 84 108 120 131 153 184 212 231 274 314

In compensating drippers, the allowable lateral length depends on four factors:


(i) The pressure at the lateral inlet: Higher the working pressure (in the limits of
maximum), the longer the lateral, ensuring the regulating pressure at its distal end;
(ii) Dripper flow rate: Higher flow rates impose the use of shorter laterals;
(iii)Distance between drippers: Smaller the spacing, shorter the lateral; and
(iv) Inner diameter: Smaller the inner diameter, shorter the lateral length, considering the
combination of higher friction head loss and enhanced dripper’s Kd.

The nominal and actual flow rates in case of non-compensating drippers installed on thick
walled laterals at different pressures is given in Table 13.4.

Table 13.4: Nominal and actual flow rates of non-compensating dripper installed on thick-walled laterals
Nominal flow Pressure (bars)
(lit/hr) 1.00 1.50 2.00 2.50 3.00
Actual flow rate (lit/hr)
1.05 1.05 1.27 1.44 1.60 1.74
1.60 1.60 1.93 2.20 2.44 2.65
2.10 2.10 2.53 2.89 3.20 3.48
4.20 4.20 5.06 5.78 6.40 6.96
8.40 8.40 10.12 11.56 12.81 13.93

Lateral length of non-compensating drippers is affected by topography. Comparing the same


nominal flow rate of 1.6 lit/hr for the same lateral diameter – it can be seen that the allowable
lateral in case of the non-compensating dripper lateral is significantly shorter than that of
compensating ones (Table 13.5 and Table 13.6). Longer lateral lengths for compensating
drippers are allowable since uniform flow rate is kept as far as the head is above the
regulating pressure. It can be noticed in tables 13.5 and 13.6 that topography affects the
allowable length using non compensating drippers, while its impact on lateral length having
compensating drippers is less pronounced and hence ignored in the data (Table 13.2 and
Table 13.3).

The nominal and actual flow rates in case of non-compensating drippers installed on a thin
walled lateral of different pressures are given in Table 13.7.

Table 13.7: Nominal and actual flow rates of non-compensating dripper installed on thin-walled lateral
Nominal flow Pressure (bars)
(lit/hr) 1.10 1.40 1.70 2.00 2.30
Actual flow rate (lit/hr)
1.20 1.25 1.40 1.52 1.64 1.75
2.00 2.09 2.35 2.58 2.79 2.98
2.90 3.03 3.41 3.74 4.04 4.32

124
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Table 13.5: Maximum allowable lengths of thick walled lateral using non-compensating drippers
Model 16012, OD=16.10 mm. Wall thickness =1.20 mm, ID=13.70 mm, PN = 4.0 bars, Kd=0.45
Inet pressure =1.4 bars, Nominal flow rate = 1.05 lit/hr
Spacing between drippers (m)
Slope (%) 0.20 0.25 0.30 0.40 0.45 0.50 0.60 0.75 0.90 1.00
Maximum lateral length at 10% flow variation (m)
Uphill -2 62 72 80 94 100 105 113 124 131 135
-1 68 80 91 110 118 126 140 158 175 183
0 74 88 102 127 139 150 171 201 229 246
Downhill 1 78 95 111 141 155 170 196 235 271 295
2 83 101 119 152 168 185 215 260 302 330
Inet pressure =1.4 bars, Nominal flow rate = 1.60 lit/hr
Spacing between drippers (m)
Slope (%) 0.20 0.25 0.30 0.40 0.45 0.50 0.60 0.75 0.90 1.00
Maximum lateral length at 10% flow variation (m)
Uphill -2 49 58 65 77 82 87 95 106 114 119
-1 52 62 71 86 94 100 112 128 142 150
0 56 67 77 96 105 114 130 153 174 187
Downhill 1 59 71 83 105 116 126 145 173 200 217
2 61 75 87 112 124 135 157 189 220 240

Table 13.6: Maximum allowable lengths of non-compensating thin walled dripper lateral
Model 16320, OD=17.02 mm. Wall thickness = 0.81 mm, ID=15.40 mm, PN = 2.3 bars, Kd=0.1
Inet pressure =2.0 bars, Nominal flow rate = 1.20 lit/hr
Spacing between drippers (m)
Slope (%) 0.20 0.30 0.40 0.50 0.60 0.75
Maximum lateral length at 10% flow variation (m)
Uphill -2 80 101 116 128 139 151
-1 87 113 134 153 169 189
0 95 126 154 178 202 234
Downhill 1 101 137 170 201 229 273
2 107 148 184 219 251 299
Inet pressure =1.4 bars, Nominal flow rate = 2.00 lit/hr
Spacing between drippers (m)
Slope (%) 0.20 0.30 0.40 0.50 0.60 0.75
Maximum lateral length at 10% flow variation (m)
Uphill -2 63 80 92 103 115 125
-1 67 87 104 118 133 151
0 72 96 116 136 154 177
Downhill 1 75 103 126 151 170 202
2 79 109 135 160 186 220

In thin walled laterals, the increased inner diameter allows for somewhat longer laterals but
the restriction on the level of working pressure has the opposite effect. Using the data from
the given tables, for 100 m lateral length and 30 cm spacing between drippers on a lateral,
optimum combination of nominal flow rates, working pressures, and the maximum allowable
lateral lengths for different dripper models are shown in Table 13.8.

125
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Table 13.8: Optimum combination of compatible drippers and allowable lateral lengths
Model Pressure Nominal flow rate Working pressure Allowable
compensation (lit/hr) (bar) length (m)
RAM 16012 Yes 1.2 4 151
RAM 16012 Yes 1.6 4 125
RAM 16012 Yes 2.3 4 98 (marginal)
RAM 16009 Yes 1.2 3 170
RAM 16009 Yes 1.6 3 141
RAM 16009 Yes 2.3 3 111
Tiran 16012 No 1.05 4 102
Typhoon 16320 No 1.2 2.3 126
Typhoon 16320 No 2.0 2.3 96 (marginal)

The final choice of the dripper model depends on preferences, cost and designer’s past
experience. The use of thicker-walled laterals guarantees longer durability and better pressure
surge withstanding. Lower flow-rate requires longer irrigation duration.

In calculation of head losses, the critical water path needs to be considered, namely, from the
control head to the most distant and/or topographically highest located irrigation block. A
final layout showing alignment of laterals, manifolds, main pipe for the example maize field
using retrievable drip irrigation system is shown in Figure 13.3. A summary of total dynamic
head requirement at the pump using compensating RAM dripper 16012 having 1.6 lit/hr at
nominal flow rate and 30 m as the inlet pressure is shown in Table 13.9.

Table 13.9: Calculation of total dynamic head-design summary


Compensated RAM dripper 16012, 1.6 lit/hr, Inlet pressure = 30 m
a. Flow rate
Item Unit Quantity Unit flow rate (lit/hr) Total flow rate (lit/hr)
Dripper 1 1.6 1.6
Lateral Dripper Nos. 333 1.6 533
Block Laterals Nos. 60 533 31,980
Plot Block Nos. 14 31,980 447,720
b. Head losses on the way to the critical path of the pipe network
Segment Length (m) Diameter/class Head loss (m) Cumulative head required
E-end point Minimum required head 10 m*
D-E Lateral 100 m LDPE** OD 16 mm, 5.8 15.8 m
ID 13.6 mm
C-D manifold 30 m HDPE*** 50/4 mm 1.6 17.4 m
Hydraulic valve Kv = 95 2 1.2 m + 0.4 m riser 19.0 m
A-C main 490 m PVC**** 140/6 mm and Tee 6.0 25 m
Hydrometer Kv = 135 Globe 4” 2.2 27.2 m
Filter Pair-3” 5.0***** 32.2 m
Spin-kleen filters
Riser, bend and connectors 4” 2.0 34.2 m
Water lift from a reservoir (suction lift) 7.0 41.2 m
Dynamic head required at the pump 41.2 m
Comments:
* Although the compensating (regulating) pressure threshold is 4 m, it is advised to design somewhat
higher pressure to guarantee drippers’ performance in case of pressure drop.
** Low Density Polyethylene
*** High Density Polyethylene
**** Polyvinyl Chloride
***** The head-loss related to filters is the upper limit of head loss allowed before back flushing.

126
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

(b) Pump selection


The required pump horse power (N) can be calculated as follows:

N=(Q×H)/(270 ×ƞ)

Where;
N = required Horse Power (HP)
Q = pump discharge (m3/hr)
H = total dynamic head (m)
Ƞ = pump efficiency (expressed as decimal fraction)

By replacing the respective values in the above formula we get:


N = (64 m3/hr × 41.2 m)/(270 × 0.75***)
= 13 HP

*** Comment: Generally pump efficiency is higher than 75%. However, gradually the pump
efficiency goes on decreasing. The 75% value is therefore regarded as the threshold
efficiency for refurbishing the pump. This lower value is adopted in design to guarantee the
appropriate performance of the drip system in long run.

(c) Equipment list


The equipment list includes a brief description of all the system components, the price of
each item and total cost. It will be used for comparing alternative layouts. In retrievable drip
system, the cost of the device for deployment and retrieval of laterals should be included. A
sample proforma to prepare the equipment list is shown below.

Item Unit Quantity Unit cost Total

(d) Operation schedule


The operation schedule of the system is based on the data collected and elaborated in the
planning phase. The schedule is based on the peak season water requirement.

Using the basic data given in section 13.3, following design parameters are worked out.
1. Allowable depletion: Available water × % of allowable depletion =12%×60%=7.2 vol%
2. Effective soil reservoir per Ha: 10,000 m2 × 1m (root depth) × 40% (1 m row spacing) =
4,000 m3.
3. Water deficit in the allowable depletion state: 4,000 m3×7.2% = 288 m3/Ha. This is the
water amount needed for replenishment of the deficit.
4. Daily water requirement: 7 mm/day = 70 (m3/Ha)/day.
5. The derived interval between water applications: Deficit/daily water demand = (288 m3)/
(70 m3/day) = 4.11 days. The integer number will be the scheduling time interval. The dose
has to be adjusted to the actual interval: 70 m3× 4 days = 280 m3/Ha per application.
6. Number of emitters per Ha: 10,000 m2/ (1 m × 0.3 m) = 33,333
7. Application rate: 1.6 lit/hr × 33,333 = 53,333 lit = 53.333 m3/Ha/hr = 5.333 mm/hr
8. Irrigation duration: 280 m3/hr/53,333 m3/hr = 5.25 hr. Since drip irrigation is not sensitive
to wind and evaporation, water application can take place round the clock. It is advised to
leave few reserve hours to secure water application in case of electricity shutdowns or

127
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

maintenance requirements. Four turns per day will leave three reserve hours and three turns –
8 ¼ reserve hours.
9. Time schedule: The field is divided into 14 blocks. If one block is irrigated at a time, four
blocks can be irrigated per day and the irrigation cycle will last for four days, corresponding
with the calculated interval. Since some farming activities have to be performed, although
drip irrigation intervene only slightly in these activities, it is better to finish the irrigation in
two days, by irrigating two blocks at a time. The irrigation will last two days, leaving two
days in the cycle for farming activities. Taking in to account these, two opposite blocks will
be irrigated simultaneously in each application.

Alternative layouts
As can be seen in Table 13.8, three more choices of non-compensating drippers are feasible.
(i) The thick-walled lateral imposes use of drippers having flow rates as 1.05 lit/hr,
which are more prone to clogging. Additionally, higher pressure has to be applied in
order to guarantee acceptable uniformity of distribution.
(ii) Of the thin-wall laterals, drippers with nominal flow rates of 1.2 lit/hr and 2.0 lit/hr
can be used. The system will be cheaper but it will last for relatively shorter time
because of the seasonal deployment and retrieval of laterals.
(iii) Use of thin-wall (tape) laterals is common for short period (1 - 5 years), depending
on the wall thickness. Table 13.10 presents an example of one of the latest tape
claiming high uniformity. This tape has CVm (manufacturing coefficient of variation)
as low as 2.5 % (0.025) and attainable EU of 90% - 94%.

In thin-walled tapes, particularly with short spacing between emitters, it is common to


designate flow rate to length unit of the lateral and not of the single emitter. In the past, the
uniformity requirement from tapes was lower than the thick walled laterals; EU 85% was
accepted, compared with 90% in laterals of discrete drippers. Recently, in the thin-walled
tapes, production improvements have facilitated achieving the same high uniformity as in
thick-walled laterals. Use of thin walled tapes has to be done cautiously, as these tubes have
low working pressure (5 – 10 m) and pressure regulators need to be used.

Table 13.10: A thin-walled tape data


Emitter spacing Flow rate Allowable lateral length (m) for different EU values
(cm) (lit/m/hr) 94% 92% 90%
20 5.0 93 113 126
30 3.0 132 155 175
40 2.4 163 192 217
60 1.5 221 258 292

The shaded row shown in Table 13.10 presents a lateral corresponding to the former design
example. This tape can be used, provided no pressure surges would occur and taking into
account shorter service duration. Application rate is lower than in the former design. Time
length of irrigation in the peak season will be 280 m3 per hour/ 30 m3 = 9.6 hours, which is
approximately twice than the chosen alternative. The low application rate allows for
simultaneous irrigation of four blocks, compared with two blocks in the former design.

In a drip system, it is necessary to install air and vacuum vents and pressure regulators to
prevent pressure surges and water hammer that may burst the laterals. In tapes and other thin
- walled laterals - these devices are particularly essential due to the low working pressure.

128
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Effect of topography
In the given example, topography has no effect on the head loss due to flat nature of the field.
However, in case of sloping topography, hydraulic design should take into account its effect.
If feasible, mains and sub-mains should deliver water from the higher points to the lower
ones. Manifolds should be laid on sloping land adjacent to the higher side, asymmetrically
splitting the laterals. As far as possible, laterals should be laid from the higher elevation to the
lower.

Use of compensated drippers decreases the difficulties of design in difficult topographic


conditions. Use of pressure regulators may also enable balance the pressure in slopping
fields.

For sake of convenience in farm activities, whenever possible, water conveyance pipes and
manifolds should be installed at sufficient distance from the crop rows.

13.4 Design if irrigation system in greenhouses


In greenhouses irrigation systems are matched with the combination of crop-bed
characteristics. Many growing beds (substrates) have extremely low water retention capacity
leading to frequent watering. In greenhouses crop water consumption is defined per hour and
in many cases water is applied several times a day. Irrigation in shallow detached-beds of low
water retention releases considerable amounts of water and nutrients in drainage. In order to
eliminate environmental contamination and save the water and nutrients, circulation of
drained water is accomplished in many greenhouses. The reuse of recycled water for
irrigation obligates sterilization of the water. Sterilization is achieved either by deep sand
filtration, UV irradiation, heating or ozonation.

Control of temperature and relative humidity of air is essential for sensitive crops grown in
greenhouses. Irrigation can be used to moderate extreme, high or low temperatures by
absorption or release of heat from the water. The relative humidity of air can be increased by
irrigation.

Irrigation for climate control can be achieved by installing misters, foggers or ordinary micro-
jets and micro-sprinklers. The irrigation for climate control is applied intermittently in small
pulses. Operation is regulated by computerized controllers or simple timers. Frequently, a
dual irrigation system is installed – drippers for irrigation and misters/micro-sprinklers for
climate control.

129
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 14 – Design of distribution network


14.1 Introduction

So far we have discussed the design of on-farm application surface irrigation systems such as
furrows, basin and border; sprinkler and drip systems, and their variants. Once the application
system is designed, it is equally important to design the water distribution system to provide
the required quantity of water at the required elevation, and for the required duration to the
irrigation unit. Lack of a well-designed water distribution system would make it more
difficult to achieve good water management at the farm level, and would nullify the benefits
of on-farm application system design.

One of the outputs from the design of application systems is the required inflow rate from the
on-farm distribution network to the field unit. To accomplish the above, water must be
withdrawn from a tertiary canal, conveyed through and controlled in a channel, and released
into the irrigation unit. The on-farm distribution system must be designed to provide the
required degree of water control and flexibility to the farmer. Several different types of
structures are needed to provide the required quantity of water to the field unit. This chapter
provides the design aspects of a variety of structures used to convey, control and regulate
water supply to field units.

14.1.1 Conveyance structures

Like on-farm application systems, the design of on-farm distribution network has not
received very much attention in the past, and appears to be the weakest link in the overall
water control at the farm level.

Several different types of on-farm structures – open-channel and closed-conduit (both above
ground and below ground), portable and permanent, manual and automated - are used to
convey, control, and distribute water on the farm. This chapter will deal with the design of
several of these structures.

The requirements of any conveyance system are: that it have sufficient capacity to deliver the
required amount of water to any point whenever needed; that the flow of water can be
accurately controlled; that soil erosion and seepage losses can be kept to a minimum; that
maintenance and weed control work can be easily performed; that it be convenient to operate
so that lobor requirements are not excessive; and that installation costs can be economically
justified by the returns expected from the crops to be grown. The distribution system must be
located with care so that it will adequately meet the above requirements.

14.1.2 Control structures

Control structures are necessary to regulate velocity and flow rate in a conveyance system.
Several different types of structures are used depending upon whether it is an open-channel
system or a closed-conduit (pipeline) system. The number and the type of structures required
depend upon the type of ditch, the slope, and the obstacles to be encountered in conveying the
water to its ultimate delivery point, and the degree of control required.

130
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

14.1.3 Outlets
These are devices used to release irrigation water from a ditch or pipeline to the field. Their
purpose is to control the amount of water being diverted into each basin, border-strip, or
furrow. They should be easy for the irrigator to operate, and should provide protection against
erosion of ditch bank or the surface of the field at the point of release. There are several types
of outlets, and the choice depends on the method of irrigation.

14.2 Open-channel structures

Open-channel structures are the most commonly used structures around the world. Though
these are cost effective for a small farmer, they do need regular maintenance and take land
away from cultivation. In sandy soils, lined channels are preferable, particularly where water
and energy are expensive.

14.2.1 Conveyance structures


The gravity flow distribution system consists of lined and unlined channels, and flumes for
conveying water from its source to the high edge of the field, where it can be released into
basins, border-strips or furrows. For open structures, such as ditches or flumes, the alignment
must provide a uniform slope. Where the land has been graded to a uniform plane, the ditch
or flume can usually be placed in a straight line across the upper edge of the field or fields to
be irrigated. Where the land has an irregular surface, the ditch must follow the general
contour of the field in order to obtain a uniform slope. In the design of field channels, a
minimum head difference of 15 cm between the field surface elevation and the water surface
elevation in the field channel must be provided.
1
Q 
2/3
The design of ditches is based upon the Manning equation: AR So
n

in which Q is flow rate through ditch (m³/s); A is cross-sectional area of ditch (m²); R is
hydraulic radius (m); n is Manning roughness coefficient (Table 14.1); and So is slope of
ditch which is usually set equal to the cross-slope in the field. Several different geometries –
triangular, trapezoidal, rectangular, and semi-cicular- are used. The exact relationships for
area and hydraulic radius are presented in Table 14.2. However, the most commonly used
channel is trapezoidal in shape. The recommended side slopes and the maximum permissible
flow velocities for different types of materials are presented in Table 14.3.

14.2.2 Control structure


Control structures are used to regulate the water surface elevation, and the velocity of flow in
the channel in order to minimise erosion. Control structures are also used to proportionately
distribute the available amount of water among several users, and several field units on a
given farm.

Check structures
Check structures are required to regulate water surface elevation in the field ditches. Either a
temporary or permanent type of structure is used. The most commonly used structure is a
portable Canvas Dam. Wooden planks, metallic vertical sliding gates, and thin concrete
blocks are also used to control water level in channels (Figure 14.1). The height of the check
structure depends upon the slope of the channel, and head required to discharge the required
flow rate through the spiles, the siphon tubes, and the concrete outlets, and a free board of 8
to 15 cm.

131
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Table 14.1 Values of Manning roughness coefficient (n) for earthen and lines channels
Type of channel and description Roughness coefficient (n)
Minimum Normal Maximum
A. Excavated earthen channels
a. Straight and uniform
1. Clean, recently completed 0.016 0.018 0.020
2. Clean, after weathering 0.018 0.022 0.025
3. Gravel, uniform section, clean 0.022 0.025 0.030
4. With short grass, few weeds 0.022 0.027 0.033
5. With long grass and weeds 0.030 0.040 0.045
b. Winding and sluggish
1. No vegetation 0.023 0.025 0.030
2. Grass, some weeds 0.025 0.030 0.033
3. Dense weeds or aquatic plants 0.030 0.035 0.040
4. Earth bottom and rubble sides 0.028 0.030 0.035
5. Stony bottom and weedy banks 0.025 0.035 0.040
6. Cobble bottom and clean sides 0.030 0.040 0.050
c. Channels not maintained, weeds and brush uncut
1. Dense weeds, high as flow depth 0.050 0.080 0.120
2. Clean bottom, brush on sides 0.040 0.050 0.080
3. Same, highest state of flow 0.045 0.070 0.110
4. Dense brush, high stage 0.080 0.100 0.140
B. Lined or built-up channels
a. Cement
1. Neat, smooth surface 0.010 0.011 0.013
2. Mortar 0.011 0.013 0.015
b. Concrete
1. Trowel finish 0.011 0.013 0.015
2. Float finish 0.013 0.015 0.016
3. Finished, with gravel on bottom 0.015 0.017 0.020
4. Unfinished 0.014 0.017 0.020
c. Brick
1. Glazed 0.011 0.013 0.015
2. In cement mortar 0.012 0.015 0.018
d. Masonry
1. Cemented rubble 0.017 0.025 0.030
2. Dry rubble 0.023 0.032 0.035

Table 14.2 Channel-section geometric elements


Channel Type Area Hydraulic Radius
Rectangular By by
b  2y
Trapezoidal (b+zy)y b  2 y y

b  2y 1  z
2

Triangular zy² zy

2 1 z
2

Semi-circular r
2
r

2 2

132
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Table14.3 Suggested maximum flow velocities and side slopes for lined and unlined channels
Type of surface Maximum flow velocity (m/s) Side slopes (z)*
Unlined ditches
Sand 0.3-0.7 3
Sandy loam 0.5-0.7 2.0-2.5
Clay loam 0.6-0.9 1.5-2.0**
Clays 0.9-1.5 1.0-2.0**
Gravel 0.9-1.5 1.0-1.5
Rocks 1.2-1.8 0.25-1.0
Lined ditches
Concrete
Cast in place 1.5-2.5 0.75-1.5
Precast 1.5-2.0 0.00-1.5!!
Brick 1.2-1.8 0.00-1.5!!
Asphalt
Concrete 1.2-1.8 1.00-1.5
Exposed membrane 0.9-1.5 1.50-2.0
Buried membrane 0.7-1.0 2.0***
Plastic
Buried membrane 0.6-0.9 2.5***
* z is the horizontal unit to one vertical unit
** side slope of 1:1 for small canals in clay and clay loams are common
!! small precast and brick channels may have vertical walls (z = 0)
*** maximum flow velocities will depend on cover over the membrane

Figure 14.1 Check structure in operation

Drop structures
In places where there is a sudden drop in elevation, drop structures are used to control
erosion. These permit the ditch to be constructed as a series of flat channels, each at a
different elevation. The drops are generally spaced so that the difference in elevation at each
drop structure does not exceed 50 cm. A stilling basin or concrete or rock apron is generally
provided to absorb the energy of the falling water. The drop may also serve as check
structures in which case they are usually built like box gates, except that the floor of the
outlets is at a lower elevation than the inlet section. The flashboards serve as an overflow
crest when the water is being ponded above the structure.

133
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Division boxes
Division boxes are necessary to divide flow between several fields, and irrigation units that
are irrigated simultaneously. These are box like structures with two or more outlets for water.
Wooden as well as concrete division boxes are used in practice. The discharge rate from
division boxes are calculated using either an orifice equation or a weir flow equation with an
appropriate discharge coefficient.

14.2.3 Best Hydraulic Section


Channel cross-sectional shapes can vary, but regardless of the shape selected it is usually
desirable to maximise the hydraulic radius. The hydraulic radius is an index of efficiency of
flow cross section of a channel. The amount of channel surface present to create frictional
resistance to flow increases with increasing wetted perimeter. The larger the flow area is
compared with the wetted perimeter, the easier water moves through a channel. The channel
section having the least wetted perimeter for a given cross-sectional area has the maximum
hydraulic radius, which is known as the best hydraulic section. Table 14.4 lists geometric
elements of best hydraulic sections commonly used in channel design. However, structural
and economic considerations related to the nature of the soil and materials available for lining
channels may impose limitations on the use of best hydraulic sections. The best hydraulic
section should be applied only to the design of lined non-erodible channels.

In irrigation applications, most channels are trapezoidal. The best hydraulic section for a
trapezoidal channel is defined when the channel bottom width is:

b  2 y tan
2
where θ is the angle of side slope. Half of a hexagon, which has θ = 60°, is the best hydraulic
section for a trapezoidal-shape channels. However, the angle of the side slope varies,
depending on the material with which the channel is built; a channel built in sandy soil
requires a milder side slope than one built in clay soil. The best hydraulic section for a
rectangular channel results when channel width is twice flow depth, which is half of a square.

Table 14.4 Dimensions for the best hydraulic sections


Cross section Area Wetted Perimeter Hydraulic Radius Top Width
Trapezoid 1.73y² 3.46y 0.5y 2.31y
Rectangle 2y² 4y 0.5y 2y
Triangle y² 2.83y 0.35y 2y
Parabola 1.88y² 3.77y 0.5y 2.83y
Note: y = depth of flow

14.2.4 Outlets

Concrete pipe outlet


This is the most commonly used device. The pipe is always buried into the ditch bank. In
most cases, rock riprap will be needed at the pipe discharge end to control erosion. The outlet
may be with or without a gate. In the case of gated outlets, the gates are mounted vertically in
a rectangular shaped concrete or masonry lined ditches, and along the side-wall (slanted) for
trapezoidal linings. Gates attached to pipe can be mounted either vertically or at an angle, and
can be used in earthen ditches. The discharge rate through a concrete pipe outlet is calculated
using an orifice equation with an appropriate Cd value.

134
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Circular concrete outlets


The outlet gates are operated fully open or closed, which is particularly adapted to a
rotational irrigation system. These are generally used in pairs-one to check the flow in the
ditch and the other to divert the flow. The discharge through the structure is calculated using
an orifice equation, with an appropriate Cd value.

Siphons and spiles


Siphon tubes are widely used for furrow irrigation and flood irrigation from lined and unlined
head ditches. Aluminium and plastic tubes are quite common, and are sometimes shaped to fit
over the bank of the farm channel. These are set manually over the ditch bank with each
irrigation (Figure 14.2). To use a siphon it must be filled with water to take out all the air, and
then laid over the channel bank. This must be done quickly and a hand placed over the end of
the pipe to stop air from re-entering.

Figure 14.2. Siphon tubes and spiles

Spiles are short lengths of pipe buried in the channel bank (Figure 14.2). They are made from
rigid plastic or metal, but sometimes local materials such as hollowed wood or bamboo are
used. The bottom elevation of spiles should be above the unchecked water surface elevation,
if possible. Otherwise, a plug is needed to close the pipe when it is not in use.

Normally, spiles are installed each season although they can be left in place for longer
periods. Hence they can be classified as permanent or semi-permanent, whereas the siphons
are moved from one set to the other during each irrigation. The discharge rate through the
siphons and spiles is based upon the orifice equation.

14.3 Pipe line system

Sometimes closed conduits (either above or below the ground surface) are used to convey and
distribute water to different points on a farm. In addition to longer life span and improved
control on water, they require less maintenance and provide minimum obstruction to farming
operations. However, because of friction losses, sufficient head must be available for water to
flow through the pipe.

14.3.1 Conveyance structures


Flow in pipelines is based upon gradient in energy. In the estimation of pressure distribution
in a pipe, the following equation, which is called the Bernoulli's equation, is used:

135
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

2 2
V1 P1 V2 P2
Z1    Z 2
   H f ,1  2
2g  2g 

in which Z = elevation(m); P/γ = pressure head (m); γ = specific weight of water (N/m³);
V²/2g=velocity head (m); and Hf(1-2) = friction loss(m) between points 1 and 2. The subscripts
1 and 2 refer to any two points along the length of the pipe.

The velocity is calculated using (V = Q/A), in which Q = discharge rate in pipe (m³/s); and A
= cross-sectional area of pipe (m²). Since the flow rate decreases along the length of the pipe,
the velocity also decreases proportionately.

14.3.2 Control structures

Inlet structure
Inlet structures are required where water from the supply canal enters the on-farm pipe
distribution network. A submerged slide-gate placed in the side of the canal and connected to
the inlet end can be used for controlling the flow of water into the pipeline. The gate should
be mounted on a concrete headwall structure to provide stability and prevent leakage. The
gate should have an opening of nearly the same size as the diameter of the pipeline so that
entrance losses can be kept to a minimum.

Usually a water measurement structure is incorporated into the inlet structure. Screens are
recommended for removing trash or moss that might enter the pipeline from a canal or
reservoir. Sand traps may also be advisable when excessive amounts of sand or silt are
carried in the canal water to be released into the pipeline. These are usually basins, which are
large enough to slow the water to the point that suspended particles settle out.

Stand pipes
Where the land has considerable slope away from the water source, excessive pressures must
be prevented from developing in the pipeline. Open top stands equipped with baffles or weirs
offer one solution; the pressure in the pipeline above a stand is limited by the elevation of the
surface of the water flowing over the baffle in the standpipes.

Valves
These are used to control the flow of water in the mainline and laterals. There are several
types of valves in common use: gate valves, sluice valves, alfalfa valves, and orchard valves.

Slide gates
These are located in open concrete chambers called stands. The gate slides on vertical steel
frame using a hand wheel and screw mechanism. The water level may rise and fall in the
stand as the gates are opened and closed and this is useful measure of water available in the
pipe system. Several gates may be located in one stand from which water is diverted into
different pipelines.

Gate valves
These are sometimes called line gates. They consist of a valve body connected into the
pipelines. A wedge shaped gate slides inside a groove in the valve body and its movement is

136
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

controlled by a hand-wheel and a screw shaft. This type of valve is used when operating
pressures are too high for open stands.

Vents
In any pipe system there is always the problem of forcing air out of the pipe when it is being
filled with water and allowing air to enter again when draining down. To overcome this
problem air vents are used. These are located at each end of a pipeline, at bends and at any
high spot to let air in and out. Sometimes they are also placed at intervals of about 200 m on
straight pipelines with uniform slopes.

One type of vent is a simple open-topped pipe. This must be tall enough to contain the
maximum height to which water will rise under normal operation. A minimum height of 1.25
m above the ground surface is recommended. The diameter of the pipe is at least 15 cm. The
vents should be high enough to provide a free board of at least 60 cm above the maximum
height to which water will rise at that point during normal operating conditions. Open-topped
valves stands will also allow trapped air to escape from a pipeline.

When the operating pressure is too high for an open vent an air valve is used. This contains a
floating ball. When the pipeline is filling with water air escapes round the ball and against the
opening and stops water from escaping. On draining down the pipeline air can re-enter as the
ball moves away from the opening.

14.3.3 Outlets

When using pipeline distribution systems, several different types of outlets are used in
delivering the water from the distribution system to the field. Once the water comes out of the
outlets the water may be let into an open-channel or a gated pipe.

Alfalfa valves
These are used for large flows (10 – 500 l/s) and consist of a steel disc with a screw shaft
mounted in a frame on top of the riser pipe. The size of the valve can vary from 150–500 mm
in diameter depending on the discharge. The riser pipes usually have the same diameter as
the valve. The discharge from the outlet can be easily controlled by raising or lowering the
valve plate. Water is released under low-pressure from the alfalfa valves to minimise erosion
hazard. However, because of friction loss as the water flows through the valve, additional
pressure must be provided to overcome the head loss. The general practice is to limit this
additional pressure requirement to about 0.30 m above the ground surface.

Valves are normally spaced so that there will be one outlet for each border-strip or basin.
When used with furrow irrigation, they may be spaced 20 to 40 m apart. Valves are normally
spaced so that there will be one outlet for each border-strip or basin. When used with furrow
irrigation, they may be spaced 20 to 40 m apart. When irrigating furrows, most commonly,
the alfalfa valve is connected to a gated pipe using a portable hydrant.

Orchard valve
These valves are generally used where the required flow rates are small (1 to 75 lit/sec), for
irrigating a small number of furrows (2 to 4) or for narrow border-strips. Valves are usually
spaced from 2.5 to 6 m apart. The sizes vary from 38-200 mm in diameter. The valve is set
inside the riser and is smaller than the diameter of the pipe. Water flowing from the outlets

137
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

can cause erosion. For this reason they are usually set about 100 mm below the surrounding
ground level. When water is released it forms a pool around the valve which dissipates the
energy in the water.

Hydrant
These are used with alfalfa and orchard valves to improve control over flow from a pipe
outlet. When using gated-pipe a portable hydrant can be used which fits over the pipe outlet.
This is made from aluminium or galvanised steel and it is pushed down into wet soil around
the riser pipe to form a seal. On some hydrants a more effective seal is achieved using clamps
which secure the hydrant to the riser pipe. A screw handle on the hydrant fits over the handle
on the valve so that it can be opened. Another type of hydrant is a more permanent one,
which is often used with furrow irrigation. It consists of a concrete pipe placed vertically over
the riser with small metal slide gates. The slide is generally notched at the bottom so that the
size of the opening can be carefully regulated when only small flows are wanted. The pipe is
usually open at the top so that the orchard or alfalfa valve can be adjusted as needed.

138
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 15 - Water delivery


Water delivery systems transport the lifeblood of an irrigation project. The selection and
management of water delivery systems affect agronomic and social aspects of projects. The
control strategy must be compatible with the flexibility of the ultimate water delivery method,
and the social, political, geographical, and economic conditions under which it will be used.
The strategy of water delivery, the desired delivery method, required interfacing, and the seat
of decision making must be carefully examined before initial selection or modification. Water
deliveries to the fields which are inflexible, unreliable, and/or unpredictable restrict the
attainment of high field irrigation efficiencies in most areas.

15.1 Delivery methods


A primary function of an organisation controlling water in a delivery system is to allocate
water to water users. Main systems are responsible for distributing water to local irrigation
organisations who are responsible for delivering water to farmers. Many means of allocating
water have been devised to adapt to specific environments of individual irrigation systems.
Policies that organisations use to allocate water among users are called delivery methods.

15.1.1 Classification
A common classification of delivery methods is to broadly categorise them into continuous
flow, rotation or demand. When water is delivered as requested by a water user it is called a
demand system. Rotational method is one in which the organisation sets a time and amount of
water which can be taken by a water user. Continuous flow is when water is delivered to each
farm or canal on continuous basis. This classification scheme is useful when broad
characterisations are desired, but does not take into account the many details organisations
use to allocate water among users. In some cases, a modified demand or modified rotation or
combined canal rotation and free demand is used as circumstances require. When equity is
the main objective of water allocation, the delivery method is classified as proportional
supply.

15.1.2 Delivery flow rate, frequency and duration


The actual distribution of irrigation water has different characteristics depending on the water
distribution method utilised. They vary from totally rigid in frequency, rate and duration to
totally flexible in frequency, rate and duration. The rigid delivery methods are mostly easily
managed by the supplier, while the totally flexible methods generally produce the highest
water use efficiency on the farm if the on-farm irrigation system is well managed and the
irrigation is well planned.

The type of delivery system is important in design and management of irrigation projects.
Clemmens(1987) indicated that the following three factors are important in sizing the
delivery system: delivery flow rate, delivery duration, and the peak water requirement or
irrigation frequency.

(i) Delivery flow rate


Clemmens indicates that many systems are designed assuming a normal flow rate called
delivery flow rate. The delivery flow rate is easy to manage because the supplier and irrigator
know the supply rate, which is generally constant. The delivery flow rate is given by:
24 I g , max A t
Q max 
8640 H r

where: Qmax = the delivery flow rate (m3/s)


Ig,max = the average gross peak water use rate (mm/day)
At = total irrigated area of the block (ha)
Hr = the delivery period (hr/day of water delivery)

139
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

At is the area of the block of fields that can be irrigated using a continuous water supply or a
complete rotation system. Clemmens(1987) called At the rotational area.

(ii) Delivery duration


The delivery duration is the time required to irrigate a field (Area Ai) of the total block (Area
At).
A id g
ti 
8640Q

where: ti = the duration of an individual irrigation (days)


Ai = irrigated area (ha)
dg = gross irrigation depth (mm)
Q = system flow rate (m3/s)

The system flow rate is the flow rate used to irrigate the field. It is either equal to Qmax
(rotation of the entire flow rate) or a part of Qmax in case Qmax is split into several streams to
irrigate simultaneously several fields.

(iii)Irrigation frequency
This is irrigation frequency for the peak period.
I g , max
freq 
dg
where: freq = irrigation frequency(days-1)
Ig,max = average gross peak water use rate(mm/day)
dg = the gross irrigation depth(mm)
The minimum irrigation depth that will satisfy crop needs occurs for continuous supply
system where water is supplied for the entire time between irrigations. The frequency (freq)
of irrigation is the reciprocal of the time interval between irrigations (Int). For example, if a
field is irrigated once every 10 days, the frequency is 0.1 day-1.

15.1.3 Operation
The water distribution method is normally linked to the design of the conveyance system
although there are exceptions; therefore once a water distribution method has been selected
there is little possibility to change it. The selection of the water distribution method is thus an
important matter where social, technical and economic characteristics must be taken into
consideration. Each of the methods has its own characteristics which may, or may not, suit
local conditions.

Following the preparation of detailed field and canal layout the criteria on which the canal
system will operate need to be developed. To develop the operation of the supply system the
following indicators can be used: supply requirement factor, supply factor, supply duration
factor and design factor.

Where:
 supply requirement factor fi = Vi/Vmax which for a given period is the ratio between the
average daily supply requirements during the period i (Vi in m3/day) and the average
maximum daily supply requirements during the peak water use period (Vmax in m3/day);
 supply factor fs =Qi/Qmax which for a given period i is the ratio between the required
stream size during the period i (Qi in m3/sec) and the maximum possible stream size or
canal capacity (Qmax in m3/sec);

140
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

 supply duration factor ft = T/Int which is the ratio between supply duration (T in days) and
the supply interval (Int in days); and
 design factor  = 86400 Qmax/Vmax which is the ratio between the maximum possible
stream size or canal capacity (Qmax in m3/sec) and the average maximum daily supply
requirement during the peak water use period (Vmax in m3/day) on which the design is
based.

To function adequately an operation service requires the following personnel:


1. Water guards - They are the main communication channel between the scheme
management and the farmer, so the success of a smooth relationship between the two
parties depends on their capabilities and honesty. Their qualifications are: basic primary
school education, they should have neither relatives nor properties in the areas they serve
and they should have farming and irrigation experience. The number of water guards
needed at the scheme level is greatly influenced by the type of water distribution used and
by the number of intakes to be controlled. Also local circumstances, such as ease of
access, particular configuration of canal layouts, transport facilities, etc., may introduce
considerable variations in the manpower requirements. As far as possible the water guard
should be made responsible for entire ditches or watercourses, because there could be
friction between farmers due to differences in criteria for water distribution.
2. Operators of large structures - Operation of the main canals and large secondary canals
requires a specialised operator to handle the flow regulating hydraulic structures. These
operators have responsibility for the structures in a given stretch of canal. A main intake
may require one or more operators depending on its complexity and hours of work. Main
canal operators are often connected by telephone or radio with the Head Water Master in
the main office. Information from the water guards is passed to the main office, where it is
computed and orders for the control of flows are passed back to the canal operators. Their
qualifications are complete secondary school education and some mechanical knowledge
is desirable.
3. Pump-set operators - Where pump-sets are used, pump operators are needed. They are
particularly liable to abuse their positions because they have monopoly control over
distribution within the areas commanded by their pumps. To keep this situation in hand,
effective management control systems are especially important, linked to incentives, e.g.
through promotion, to induce correct operation. Their qualifications are complete
secondary school education and short course on operation of pump-sets. It is desirable that
they are mechanics, or alternatively have mechanical training. They must be people of
known integrity because undesirable situations can arise when operators accepts bribes
from farmers or are party to other malpractices.
4. Water Masters - They supervise the water guards and canal operators and are the main
channel of communication with the chief of Operation Service. Such an appointment is
only necessary when the group of water guards to be supervised is larger than 12 -15,
otherwise the Chief of Operation Service can do the supervision himself. Their
qualifications are secondary school level of education and ability to manage and direct
people. Preferably they should have a number of years of experience on the same scheme.
5. Chief of the Operation Service - The Chief of the Operation Service - also called
Irrigation Supervisor - is responsible for the operation of the whole scheme or a large
section of it. His main function is to collect the information provided by the water guards,
process it, and issue the operational orders to be executed. The incumbent should be a
qualified irrigation engineer (MSc), i.e. someone with technical understanding of soil-
water-plant relationships as well as engineering matters (hydraulics, construction, etc.). He
should have at least five years of professional experience in the field. An agricultural

141
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

engineer also qualifies, provided that experience in operational matters has been gained.
The ability to direct and co-ordinate people is an essential requirement.
6. Auxiliary Staff - Auxiliary staff (drivers, bookkeepers, clerical staff, etc.) should be
limited to the minimum indispensable to undertake the necessary work. When possible,
these services should be provided by a pool serving the other units (maintenance,
administration, etc.).

15.2 Water delivery methods

15.2.1 Continuous Flow


The continuous flow system is the simplest delivery method. For very rigid continuous
system, a constant flow rate is delivered to the farm turn-out. For a totally continuous system,
the supply rate is delivered at a starting time during the season and is shut off at the end of the
growing season regardless of the on-farm demand. The supplier can easily manage the system
because few decisions are needed and communication between the supplier and the irrigator
is not necessary. The constant delivery system generally leads to poor on-farm efficiency
because water is supplied when it is not needed and is unavailable in enough quantity during
peak use periods.

The continuous supply system is commonly used upto field level in most of the rice growing
areas where paddy fields are submerged continuously throughout the crop growing period. In
irrigated conditions this permanent submergence is normally achieved by providing a
continuous feeble flow that compensates the crop's daily evapotranspiration and percolation.
Any excess water is drained away from one field and provides the supply for the next low-
lying field.

The continuous supply system results in the minimum canal and delivery system capacity
since the design factor  is approximately equal to one. The supply is regulated mainly by
simple diversion structures which are easy to operate. For less rigid continuous system,
accurate adjustments in the supply proportional to actual field requirements are difficult to
handle particularly when stream size becomes small, resulting in low efficiency. However,
except possibly for some rice schemes, the fields need to be irrigated at given intervals, and at
the field level an interrupted or rotational delivery will be required. Continuous supply is in
general limited to canals serving 50 ha or more, and within the 50 ha block the supply is
rotated among the individual fields.

The supply system indicators, for a continuous system supplying acreages of 50 ha, except
possibly for rice (fi = fs) are as follows:
-supply requirement factor (fi =Vi/Vmax) - 0-1
-supply factor (fs =Qi/Qmax ) - 0-1
-supply duration factor(ft = T/I) - 1
-design factor ( =86400 Qmax/Vmax) - 1

The following are some average standards for water guards requirements for each 5000 ha
under continuous flow water distribution system, derived from projects in several parts of the
world (Sagardoy 1982).

-very large farms - 1


-large farms (5-10ha) - 1
-medium farms (2-5ha) - 2
-small farms (2ha) - 3
The computation of the area that can be irrigated by a given continuous flow is as shown in
Example 15.1.

142
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Example 15-1 Continuous delivery system


Given:
A project is to irrigate corn in Marigat using a furrow irrigation system that is 80 percent
efficient. The peak ETc is 5.3 mm/day. The soil is silt loam that has total available water of
160mm per m depth. The root depth during the peak use period (Zr) is 1.2m, and the
management allowable depletion (p) has been determined to be 50 percent. The delivery flow
rate Qmax is 0.116 m3/s, and water is delivered 24 hours a day (Hr =24) .

Find: Compute the rotational area for this system.

Solution:
1. Allowable depletion RAW = p×160×Zr = 0.5×160 mm/m ×1.2 m = 96 mm

5 . 3 mm / day
The average peak water use rate is: I g , max   6 . 6 mm / day
0 .8
2. Using the equation for the delivery flow rate, the rotational area is then:
8 6 4 0  0 .1 1 6 m / s  24 hr / day
3

A  At = 152 ha
2 4  6 .6 m m / d a y
t

15.2.2 Rotational

A rotational delivery system is also a rigid method. All the canals receive water by turns and
the farmers on the tertiary canals or watercourses receive water at a pre-set time and in the
allowed quantity. It is a highly efficient system from the operational point of view and
socially fair since it gives an equal chance to everyone.

Rotational supply, with rotation blocks of 50 to 250 ha, is well adapted to schemes with
single crop or simple cropping pattern. An advantage is relatively high conveyance efficiency
since canals are either fully filled or empty. Supply can be regulated from head-works and a
few check structures are required. Large variations in discharges are avoided. This relieves
the drainage system and minimises the effect of sedimentation and scour. A main
disadvantage can be that water supply to a diversified cropping pattern with distinct, different
irrigation requirements over area and time is very problematic. To solve this, the first solution
is the irrigation interval for deep rooted crops be a multiple of that of shallow rooted crops
and the second solution is use of relative area method, where gross depth ratio is used to
convert crop areas to relative areas. While the operation of the system, i.e. the supply
schedules, must meet the changing field requirements over the season, the design criteria are
based on supply requirements for the peak water use month.

A rotational system has advantage over a continuous delivery system because the supply rate
is large enough to manage and generally requires less labour from the farmers. A rotational
system also allows other field operations to occur more easily than continuous delivery. The
capacity of the primary delivery system is generally similar to that of the continuous flow
system, but the capacity of the system delivering water to the farm turn-out and on-farm
delivery system generally is large than that for continuous delivery.

There are several ways in which a rotational system can be implemented:


(i) The water is distributed by turns of equal duration throughout the irrigation season. The
farmer receives the water on a fixed day for an amount of time that is always constant,
regardless of the crops that he may plant.
(ii) The water is distributed by turns of different duration, longer at the beginning and end of
the irrigation season and shorter in the middle, according to crop demand. The order of
distribution within each turn is always the same and the amount delivered is constant
throughout the season.

143
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

(iii)The water is distributed by turns of different durations and the amount delivered also
changes throughout the season. The amount delivered is calculated according to the actual
crop water requirements.

The degree of technicality increases from (i) to (iii) and this not only refers to the actual
calculation of the amounts of water to be delivered but also to the design of the irrigation
network. For instance, method (iii) can only be applied if the irrigation network has water
measuring devices for each farm.

Method (i) is the simplest of the three and perhaps the most widely used. Socially speaking, it
is a fair method, since it gives every user an amount of water proportional to the amount of
land. If water is extremely scarce, the method is still quite efficient as the farmer must adapt
his cropping pattern to the fixed turns. Calculations for the water delivery are simple and can
be made easily.

Method (ii) requires a little more technical knowledge as the interval must be adapted more to
the actual needs of the crops. This can be very effective when the scheme is concerned with
monoculture but the more diversified the production, the less effective will it be.

Method (iii) technically offers the best opportunity to meet crop water requirements and
achieve greater water efficiency. However, it is difficult to implement. First of all, water
measuring devices are needed at the farm level in order to measure the amount of water that
must be delivered. Secondly, the management must have an excellent communication system
in order to inform the farmers well in advance about their turns. Thirdly, since the
calculations for the amounts of water to be delivered are made by the management and
changed from one irrigation event to the next, the system is very vulnerable to malpractice.
Fourthly, the calculation procedures are quite complicated and lengthy, needing qualified
staff for their execution. As a result of all these requirements, this method is rarely used, in
spite of its theoretical advantages.

The computation of the supply capacity and duration of irrigation of a given rotation area is
as shown in Example 15.2.

Example 15-2 Rotational delivery system


Given: Use the information from the example 12-1 and assume that the project is divided
into 10 irrigated blocks of 15.2 ha each.

Find: The supply capacity for each block and the duration of irrigation.

Solution:
1. With 10 blocks, a frequency of 0.1 days-1 could be used, thus each block
would be irrigated for a duration of 1 day. The gross irrigation would
be determined from the irrigation frequency (section 12.1.2.3) as:
I 6 . 6 mm / day
d g

g , max
 1
dg = 66 mm
freq 0 . 1 day
Since the allowable depletion is 96mm for this system, this depth is acceptable.

2. With a duration of 1 day, the flow rate to each block is determined from
the delivery duration:
Ai  F g 15 . 2 ha  66 mm
Q   Q = 0.116 m3/s
8640  t i 8640  1 day
The needed flow rate is exactly the same as that for continuous supply because no down
time or flexibility is designed into the system.

144
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Example 15-1 and Example 15-2 illustrate that the required capacity for the 152 ha area will
be 0.116 m3/s for either the continuous or the rotational delivery system. The difference
between the supply strategies comes in the size of the supply system needed to irrigate each
15.2 ha block. For the continuous system, a tenth of the delivery flow rate (0.0116 m3/s) was
supplied. With rotational delivery, each supply system must have enough capacity to carry
the delivery flow rate (0.116 m3/s) for 1 day and then will be dry for 9 days.

The supply system indicators, for a rotational system supplying acreages of 50-250 ha are as
follows:
-supply requirement factor (fi =Vi/Vmax) - 0-1
-supply factor (fs =Qi/Qmax ) - 1
-supply duration factor(ft = T/I) - 0-1
-design factor ( =86400 Qmax/Vmax) - 1-5 ( 50- 250 ha)
- 5  50 ha

The following are some average standards for water guards requirements for each 5000 ha
under rotational water distribution system, derived from projects in several parts of the world
(Sagardoy 1982):
-very large farms - 6
-large farms (5-10ha) - 8
-medium farms (2-5ha) - 10
-small farms (2ha) - 12

15.2.3 On-demand
On-demand system, a flexible method, is at the other extreme of the delivery methods. A pure
on-demand system allows users to remove an unregulated amount of water from the delivery
system at irrigator's convenience. The length, frequency, and rate of water delivery are totally
at the irrigator’s discretion.

On-demand irrigation systems are generally designed with high-level technology. The degree
of human intervention is minimal since they operate on automatic principles, i.e. when the
water level or pressure drops in a canal or pipe due to the opening of an inlet, the level or
pressure is immediately reinstated by an automatic device which calls for a greater supply,
provided by automatic gates or valves. The efficiency of these systems is very high (up to
90%) particularly when using pipes for the distribution.

There is very little actual operation in these systems and that is limited to some overall
control of the automatic system, which is usually done through remote control panels located
at the management office. Although the manpower requirements are small, the staff must be
highly qualified. Water guards are not applicable in this system (Sagardoy 1982).

The supply system indicators for an on-demand system supplying all acreages are as follows:
-supply requirement factor (fi =Vi/Vmax) - 0-1
-supply factor (fs =Qi/Qmax ) - 0-1
-supply duration factor (ft = T/I) - 0 -1
-design factor ( =86400 Qmax/Vmax) - 1

The great advantage of this method is that it allows the farmer to use the water when it is
most necessary for the crops. The main disadvantages are high costs, especially close to the
farm and the need for a high level technology in the construction and maintenance of the
systems. Thus the systems are more suitable for developed countries than developing ones
and rather unsuitable for the least developed countries. The extra cost of an on-demand
delivery system would hopefully be paid for through improved production on-farm or by
irrigating more area with the water saved from increased efficiency.

145
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

The success of on-demand systems in developing countries depends on many factors but, in
any case, the closed system (pipes) has better possibilities than the open canal system. The
reason for this is that in the latter the automatic gates are susceptible to blockage by
vegetation or misuse by human intervention and once the automatic regime of the canal is
disrupted the chances of overtopping are great, with subsequent damage. On the contrary,
pipe systems have the important advantage that they cannot be manipulated by anyone, and
thus operational and social problems (e.g. stealing of water, etc.) are reduced.

Example 15.3
Suppose that an on-demand system were implemented in Example 15.2. If the goal was to
supply the water to a block during 1 day (same duration as for the rotational system), the
maximum demand would occur where each block ordered the delivery flow rate on the same
day. Thus, the supply system to the 152 ha area must be 10 times the capacity of either the
continuous or the rotational delivery system. The supply capacity to each block would still
need to be 0.116 m3/s, obviously the cost of the on-demand system would be much higher
than that for the continuous system.

15.2.4 Semi-demand

Semi-demand delivery method varies between the rigid and pure demand methods. With
these supply method, either the rate, duration, or frequency, or all three, can be arranged
resulting in these system being perhaps the most common system of water distribution due to
its simplicity. A farmer requests the water from the water guard, who passes the information
up to the water master. He makes the necessary calculation to accommodate it with the
demands of the other farmers within the limited capacity of the canal. If the demand can be
met, the information is passed back through the water guard to the farmer with an indication
of the exact time of his turn. In irrigation systems where the canals have been designed with
certain flexibility, the request is usually met within a short time (2-3 days), although
sometimes 6-7 days can elapse in the case of canals with little flexibility and high demand.

The amount to be supplied to the farmer is usually fixed in relation to the number of hectares.
A known amount of water is an indispensable requirement for such a system; otherwise the
water master cannot calculate the time and flow needed for each request in order to prepare
the water distribution programme.

This form of distribution requires a well designed and constructed irrigation system since the
flows delivered by the canals should be well known, and the intakes should also be capable of
delivering the requested flow.

The water guard has a crucial role and he must be a man trusted and respected by the farmers.
It is preferable that he lives among the farmers of the canal in order to facilitate
communications, and he must be able to undertake elemental calculations to adjust demand
and supply.

The following are some average standards for water guards requirements for each 5000 ha
under semi-demand water distribution system, derived from projects in several parts of the
world (Sagardoy 1982):
-very large farms - 4
-large farms (5-10ha) - 6
-medium farms (2-5ha) - 8
-small farms (2ha) - 10

To avoid possible excess in water use by the farmers some restrictions may be imposed, such
as: the number of irrigations per year is limited, or a certain time (7-10 days) must elapse

146
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

between two requests for water, or those who in a given month have already received water
have a lower priority than those who have not, etc.

Another advantage of this system is that when the need arises (peak month) or during
exceptionally dry years, it can also function on a fixed rotation. The only disadvantage of this
distribution system is its low efficiency at times of low demand, because the opening and
closing of canals for few farmers could imply considerable losses. However, the problem is
mitigated since at times of low demand water losses are not so relevant.

Example 15.4
With the information given in Example 15.2. An example of a semi-demand delivery system
would be to require 2 day duration with a maximum of 5 blocks irrigated at any time. The
irrigator would need to place a water order in advance to allow time for the supplier to
provide the supply. The supplier might allow a maximum flow rate of 0.058 m3/s per block.
The irrigator could request any flow rate up to 0.058 m3/s and could request more than one
supply during a 10 day period.

Semi-demand delivery methods generally are more complicated because the probability of
various demands is needed to size the system and to manage the system once a project is on-
line. Clemmens (1986) showed that the flexibility allowed by semi-demand methods causes
the capacity needed in an irrigation project to be bigger than that for rotational systems at the
farm turn-out level, but that there was less effect upstream.

15.2.5 Canal Rotational and Free Demand

This is a combined system where secondary canals receive water by turns, for example every
7 days, and once the canal has water farmers can take the amount they need at the time they
wish. The system is adopted particularly where there is mixed control of management. The
public administration undertakes the operation of the main and secondary canals, and the
farmers either take the water freely from those canals or distribute it themselves from a
smaller canal.

The main feature of this system is that canals receive water in turns. These turns may be
changed during the irrigation season and vary from region to region. The duration of the turns
is generally the result of experience in the area. When the number of crops grown in the
irrigation project or region is fairly large, there is not much opportunity for rationalising the
duration of the turns. However, where the number of crop is limited or some crops clearly
prevail, the duration of the turns can be determined in a rotational way.

When it is their turn, the farmers take the water from the canals on free demand or they may
eventually establish some kind of rotation among themselves. In the first case, which is the
more frequent, the canal must be designed to cope with a concentration of demand at any
time. Water guards are not applicable in this system (Sagardoy, 1982).

This type of distribution can be handled easily by public organisation but it implies large
operational losses. The low efficiency of the system is mainly due to the fact that demand and
supply are disassociated: canals are filled every turn irrespective of the demand. Some
adjustments can be made in the canal water levels according to the demand pattern but still
operational water losses are bound to be high. The system can easily be improved by properly
organising the farmers along the watercourses and regulating the flows according to the
requests received from the groups of farmers on each watercourse. Under this hypothesis the
system could be much more effective.

147
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

15.2.6 Proportional Supply

This method of water distribution has some similarities with the rotation but is based in a
different principle. The basic idea originated in systems where the water flow at the
headwork is highly variable throughout the season and therefore the intention is to distribute
in the period considered, whatever amounts that arrives at the diversion point in a
proportional manner to the irrigated area. Secondary canals operate with the constant
continuous flow during the whole period considered, while rotation among plots takes place
at tertiary canals.

The basis for the water allocation is that all farmers get a volume share in proportion to his
irrigated area. The proportional share in water volume each plot will receive is normally
calculated in function of the irrigated area. The mechanism may however be improved by
taking into account the cropping pattern of the plot, or in the most equitable case, taking into
account the real areas under irrigation on the plot. It is obvious that in the latter case
operation of such a system requires very good programming, beneath the fact of having up to
date and accurate information of real cropping patterns encountered in the field. Good
estimations of crop water requirements in function of the prevailing climatic conditions, and
estimations of field wetness is also necessary.

In practice this system requires that all the canal off-takes are equipped with devices that
permit to divide the flow in a proportional manner to the irrigated area. The method is
efficient in the sense that it distributes available water in a proportional manner but it does
not take into consideration areas that are not cultivated or planted at the time of delivery.

148
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 16 - Surface and sub-surface drainage system


16.1 Introduction

Land drainage removes excess surface water from an area or lowers the ground water below
the root zone to improve plant growth or reduce the accumulation of soil salts. Land drainage
systems have many features in common with municipal storm-drain systems. Open ditches,
which are less objectionable in rural areas than in cities, are widely used for the drainage of
surface water, at a considerable saving in cost over that of buried pipe. Under suitable soil
conditions ditches may also serve to lower the water table. However, closely spaced open
ditches will interfere with farm operations, and the more common method of draining excess
soil water is by use of buried drains. Drains usually empty into ditches, although the modern
tendency is to use large pipe in lieu of ditches where possible. This frees extra land for
cultivation and does away with unsightly and sometimes dangerous open ditches. Since land
drainage is normally a problem in very flat or leveed land, a disposal works provided with
tide gates and pumping equipment is often necessary for the final removal of the collected
water.

Land drainage speeds up the runoff of water and, hence, increases peak flow downstream of
the drained area. The consequences of this increase should be considered in the planning of
drainage systems. Wetlands are important biological areas. They serve migratory waterfowl
and, in coastal areas, as nursery grounds for many important commercial species of aquatic
life. The consequence of draining such lands requires careful evaluation.

16.2 Design flows for land drainage


Land drainage is not as demanding in terms of hydrologic design as other types of drainage.
The purpose of land drainage is to remove a volume of water in a reasonable time. Where
sub-drainage is installed to remove excess water from irrigated land for salinity control, the
volume of leaching water to be applied in each irrigation event is known; and the drains
should be capable of removing this volume in the interval between irrigations.

Drainage installed for removal of excess water from rainfall is typically designed to remove a
specified quantity of water in 24 hours, commonly known as the rate or recharge or drainage
coefficient. In North-West Europe the rate of recharge is often assumed to be 7 mm/day,
however in tropical regions, where rainfall events are quite different in nature, analytical and
statistical techniques should be undertaken to generate a representative design discharge from
rainfall measurements. For example, a frequency analysis with return periods of different rain
depths can be used to determine a design discharge rate, subject to the criteria of the project.
The return period for on-farm design discharge will depend upon the value of the crop. Table
16.1 shows return periods for different types of cropping.

Table 16.1 Return period for different types of cropping


Crop value Example of cropping Design rainfall exceedance
Very high Special crops 1 year in 25
High Horticultural 1 year in 10
Medium Roots 1 year in 5
Intensive grass, cereals 1 year in 2
Low Grassland 1 year in 1

149
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Land protected by levees is often subject to excess water as a result of seepage from the river
through or under the levee. The design flow for a drainage system in this case should be
based on estimated seepage rates at river stages having a return period of 1 or 2 yr. If the high
stages are expected to be concurrent with the rainy season the drainage design discharge rate
should be added to the seepage.

16.3 Drainage ditches

A ditch-drainage system consists of laterals, sub-mains, and main ditches. Ditches are usually
unlined. Small ditches may be constructed with special ditching machines, while larger
ditches are often excavated with a dragline. Unless the excavated material (spoil) is needed as
a lavee to provide additional flow area in the ditch, it should be placed at least 4.5m back
from the edge of the ditch so that its weight does not contribute to the instability of the ditch
bank. Spoil banks decrease the cultivated area and prevent inflow of water from the land
adjacent to the ditch. If possible, this spoil should be spread in a thin layer, but where this
cannot be done, openings should be left in the spoil bank wherever a natural drainage channel
intersects the ditch and at least 150 m along the ditch.

The slope available for drainage ditches is small, and the cross sections should approach the
most efficient section as closely as possible. A trapezoidal cross section is most common,
with side slopes not steeper than 1 on 1.5. Slopes of 2:1 or 3:1 are required in sandy soils.
Occasionally, where the drainage water must be pumped, ditches are deliberately made with
an inefficient section to create as much storage as possible to minimise peak pumping loads.
The slope, alignment, and spacing of ditches are determined mainly by local topography. The
minimum practical slope is about 0.00005. The ditches generally follow natural depressions,
but where possible are run along property lines. Lateral ditches rarely need to be spaced more
closely than 800 m apart, although on level land a spacing of 400 m may be necessary. With
favourable land slopes a ditch spacing of 1600 m may be feasible.

Ditches are usually between 2 to 4 m deep. Where tile drainage is to be used, the lateral
ditches must be deep enough to intercept the under-drains, which are to discharge into it.
Similarly sub-main and main ditches must be deep enough to receive the flow of lesser
ditches. If the terrain is flat and the ditches quite long, an excessive depth may be required for
the main ditch and it may be advisable to divide the system into two or more portions to
shorten ditch lengths. In muck or peat soil, considerable subsidence may occur when the
water is drained from the soil, and ditches must be constructed proportionately deeper.

Ditch bottoms at junctions should be at the same elevation to avoid drops, which may cause
erosion. This may require some steepening of the last 30 m or more of laterals before they
join a sub-main. Right-angled junctions encourage local erosion of the bank opposite the
tributary ditch, and the smaller ditch should be designed to enter the larger at an angle of
about 30°. Erosion will also occur at sharp changes in ditch alignment, and long-radius
curves should be used where a change in line is necessary. With high flows or steep slopes
the radius of curvature should be 360 m or more, while for small ditches on flat slopes a
radius of 100 to 150 m may be satisfactory.

The most important factors controlling the value of Manning’s n for drainage ditches are the
neatness with which the ditch is constructed and the extent of vegetal growth in the ditch. If
brush and weeds are cut and burned annually, a value of n = 0.04 is typical for a well-formed

150
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

channel. If the channel section is very irregular or if several years’ growth of weeds and
brush develops, n may be 0.100.

16.4 Under-drains

Unglazed clay-tile, pvc pipe or concrete pipe are the most common materials for under-
drains, although wooden box drains and perforated steel pipe have also been used. Plain pipes
without special joints are used and are placed with the ends of adjacent pipes butted together.
Water enters the drain through the space between the abutting sections. The space should not
be so large as to permit soil to enter and clog the drain. If the pipe ends are so irregular as to
leave 3 mm openings, the joints should be covered with pieces of broken tile, tar paper, or
other material to prevent the entry of soil. Fabric and gravel drains may prove economical.
They are constructed by excavating a small trench, lining it with a synthetic fabric,
backfilling with gravel, lapping the fabric over gravel and backfilling the rest of the trench
with soil. The fabric serves as a filter to prevent fine soil material from entering the gravel
and clogging the drain.

Two-inch (5 cm) pipe was widely used in early drainage projects but modern practice favors
the use of four-inch (10 cm) minimum, and some projects have six-inch (15 cm) as minimum
pipe diameter. The larger size of pipe minimises the likelihood of clogging with sediment,
roots, or other materials. The slope of 10 cm tile should be not less than about 0.2 % to
provide a velocity of about 0.3 m/sec when flowing full to avoid sediment deposits in the
pipe.

The drains are usually spaced 10 to 50 m apart. Only the most permeable soils permit the
wider spacing, while spacing less than 15 m will usually be too expensive. Where a very
close spacing is required, the use of mole drains is sometimes feasible. Mole drains are
tunnels formed in cohesive soil by pulling a steel ball through it. If the soil is sufficiently
cohesive, mole drains may remain effective for many years. Perforated flexible plastic tubing
can be installed with a mole to maintain the drain if the added cost can be justified by the
longer drain life. Mole drains have been spaced as closely as 1.5 m apart in soil of very low
permeability.

If the land slope is such that drains can parallel the surface, there is no limit to drain length
except that imposed by the topography. On level ground a drain on the minimum slope of
0.002 will drop 0.3 m in 150 m, and this distance becomes about the maximum permissible
length. On steeper slopes drains are rarely more than 600 m long.

16.5 Flow of groundwater to drains

The flow of ground water to a drain pipe or ditch is governed by the same factors controlling
the flow to a well. Both ditches and drains create a water table like that shown in Figure 16.1.
The cone of depression about a well becomes a trough along the line of the drain. The
spacing of the drains must be such that the water table at its highest point between drains
does not interfere with plant growth. The water table should be below the root zone to permit
aeration of the soil. Higher water content also reduces soil temperature and retards plant
growth. The necessary depth to the water table depends upon the crop, the soil type, the
source of water, and the salinity of the water. In humid regions, a depth of 0.6 to 1.0 m is
satisfactory for most crops except orchards, vineyards, and alfalfa, which require a depth of 1

151
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

to 1.5 m. In general, the water table should be lowered more in heavy clay soils than in light
sandy soils. For land under irrigation the depth to the water table should be greater 1 to 1.5 m
for ordinary crops and good quality water, and 1.5 to 2.5 m for deep-rooted crops. If the
water contains excessive salts, the minimum depth for any crop should be 1.2 m. It is not
desirable to lower the water table far below the recommended minimum depth, for this
deprives the plants of capillary moisture needed during the dry season. Where deep ditches
are part of the drainage system, check dams should be provided to maintain the water level in
the ditch during low flow periods in order to avoid excessive lowering of the water table near
the ditch. Drains in peat and muck soils must be installed somewhat deeper than in other soils
because of the settlement which takes place as water is removed from the peat.

Figure 16.1: Groundwater in a drained field.

The variation in water-table elevation in the drained land should be not much over 30 cm.
This means that drains should be placed about 30 cm below the desired maximum
groundwater level. The spacing of drains must be such that, with the hydraulic gradient thus
established (approximately 30 cm in half the drain spacing), the drains will remove sufficient
water from the soil. In homogeneous soil the spacing can be determined by analysis similar to
that for a well in a homogeneous aquifer. In Figure 16.1 the hydraulic gradient at distance x
from the drain is dy/dx. Assuming the flow lines are parallel for a unit length of drain, the
cross-sectional area of flow is y. Assuming that the flow q toward the drain is inversely
proportional to distance from the drain, i.e., when x=L/2, q=0 and when x=0, q=1/2qD, where
qD is the design flow per meter of drain, the Darcy equation for the horizontal flow at x may
be written as:

2  L  qD dy
q    x  Ky
L  2  2 dx

Where; L is the drain spacing and K the coefficient of permeability. Transforming this
equation,
qD
L  2 x  dx  ydy
2 KL
and integrating,
 
2 2
qD 2x y
 Lx     C
 
2 KL  2  2

152
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

When x=0, y=a, and hence C = -a²/2. Substituting this value for C in the above equation and
solving for K,

q D Lx  x
2

K 

L y
2
 a
2

When x=L/2, y=b; hence,
4K b  2
 a
2

L 
qD

The most important single factor in determining drain spacing is the soil permeability. Since
drainage normally involves only a shallow depth of soil, small-diameter piezometers or open
auger holes may serve as test “wells.” It is convenient to calculate the permeability from the
rate of rise of water in such a hole after a portion of the water is pumped out.

Ditches, because of their larger dimensions, are somewhat more effective than pipe drains in
removing soil moisture. However, the increased effectiveness is not sufficient to permit ditch
spacing which will not interfere with farm operations. Moreover, ditches reduce the area
available for farming. Hence, open ditches for removal of soil water are used only
infrequently, and then only in the most permeable soils.

16.6 Layout of a tile-drain system

The plan for a system of tile drains is determined largely by the topography of the area.
Several possible arrangements are shown in Figure 16.2. If only isolated portions of the area
require drainage, a pattern similar to the natural stream system is most economical. This
natural system (a) is used in rolling topography where drainage is necessary only in small
swales and valleys. If the entire area is to be drained, a gridiron layout (b) is usually more
economical. Laterals enter the sub-main from one side only to minimise the double drainage
which occurs near the sub-main. The gridiron system might be used where the land is
practically level or where the land slopes away from the sub-main on one side. If the sub-
main is laid in a depression, the herringbone pattern (c) is used. The laterals join from each
side alternately. The land along the sub-main is double drained but, since it is a depression, it
probably requires more drainage than the land on the adjacent slopes. If the bottom of the
depression is wide, a double-main system (d) is often used. This reduces the lengths of the
laterals and eliminates the break in slope of the laterals at the edge of depression. If the main
source of excess water is drainage from hill lands, an intercepting drain (e) along the toe of
the slope may be all that is required to protect the bottom land.

Roots of trees can easily enter the open joints of under-drains. If possible, only the ends of
the laterals (f) should be exposed to this hazard. Mains and sub-mains should be kept well
away from trees, or bell-and-spigot joints with joint filler should be used if the drain must run
close to trees. Drainage systems often suffer considerable damage from the loads imposed by
farm vehicles. Consideration should be given to the loads in the selection of pipe for
drainage. A special heavy-duty grade of clay tile has been developed for severe conditions.

153
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Figure 16.2: Some arrangements of tile drains

16.7 Design of a land-drainage system

The basic procedure for design of land-drainage systems is not much different from that for
design of municipal storm drains. The steps in design may be summarised as follows:
- Prepare a detailed contour map of the area. A contour interval of 0.3 m is commonly
necessary;
- Select the location of the system outlet. If several outlets are possible, an economy
study of the alternatives may be necessary;
- Determine the design drainage discharge for the under-drains and estimate the amount
of water the ditch will intercept;
- Lay out a system of ditches (or pipe mains) of adequate size to carry the expected
flows;
- Determine the proper depth for tile drains and plan the tile-drain layout. Field drains of
the customary minimum size 10-cm will usually have adequate capacity, and it will be
necessary to calculate only required sizes for mains and sub-mains;
- The first trial layout of mains may require revision after the plan for under-drains is
completed. The entire system should be planned for minimum cost by use of shortest
possible routes for pipes and ditches; and
- Estimate project costs and proceed with legal steps necessary to undertake the project.

154
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Chapter 17: Environmental Impacts of Irrigation Projects


17.1 Introduction
Irrigated lands contribute significantly to the world agriculture output and food supply.
Estimates in 1986, indicated that about half of the increase in agricultural production in the
previous 35 years had come from irrigated land, about one-third of the world’s crops were
grown on the one-sixth of the cropped area which was irrigated, and the irrigated land was,
on average, more than twice as productive as rain-fed land. Figures from 1996 estimated that
developed countries, on average, irrigated 10% of their agricultural area, and countries in
development 23%, and combined they irrigated 18% of the total agricultural area. The share
of agricultural area irrigated in Latin America and the Caribbean (LAC) region in 1996 was
11%. United Nation’s predictions of global population increase to the year 2025 require an
expansion of food production of about 40-45%. Irrigation agriculture will be an essential
component of any strategy to increase the global food supply. However, overall growth of
new irrigated areas will be slow. Growth of irrigated land is expected to occur mainly in
developing countries, increasing their share from about 80% to 90% of the world’s irrigated
land. Much effort will be needed to increase the efficiency of existing irrigation systems.
Kenyan irrigation potential is estimated at 765,575 ha but only 141,900 ha (18.5%) is
currently developed (Table 17.1).

Table 17.1: Distribution of developed and potential irrigated areas per basin in Kenya
Basin Potential Developed by Developed as Balance of Balance as %
(Ha) 2010 (Ha) % of potential potential (Ha) of potential
Tana 226,224 64,425 28.5 161,799 71.5
Athi 91,006 44,898 49.3 46,108 50.7
Rift Valley 101,753 9,587 9.4 92,166 90.6
Ewaso Nyiro 49,379 7,896 16 41,483 84.0
Lake Victoria North 170,789 1,876 1.1 168,913 98.9
Lake Victoria South 126,424 13,218 10.5 113,206 89.5
Total 765,575 141,900 18.5 623,675 81.5

There is no question about the value of irrigated agriculture. Society has for long time
supported the development and improvement of irrigation. However, there is an increasing
trend to make irrigated agriculture accountable for its impact on the environment as well as to
critically evaluate the water use in the agricultural sector compared with other competing
uses. Improving the environmental performance of irrigation agriculture is also important for
its long-term sustainability. The aim of this chapter is to present different aspects of the
environmental impact of irrigation and avenues for improvement. A collage of materials, text
and figures from different sources were used to produce this article, which are listed in the
reference section.

17.2 Water, a scarce resource


Although water on earth is abundant, 97% of the world’s water lies in the ocean and seas, 2%
of all water is in glacial ice, and only 1% of all water is available for human use. About
108,000 km³ precipitate annually on the earth’s surface. It is estimated that 60% of the earth’s
annual precipitation evaporates directly back into the atmosphere, leaving 47,000 km³
flowing toward the sea. However, much of the flow occurs in seasonal floods. It is estimated
that only 9,000 to 14,000 km³ may ultimately be controlled. At present, only 3,400 km³ are
withdrawn for use.

155
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Another aspect of the water picture is the uneven geographical distribution of its availability
around the globe. Chronic scarcity of water is experienced or expected in large parts of Africa
and the Middle East, the northern part of China, parts of India and Mexico, the western part
of the USA, north-east Brazil, and in the former Soviet Union and the Central Asian
republics. Scarcity of water gives rises to conflicts across country borders (e.g. in the Nile
basin and in the Middle East) and between different interests (e.g. irrigation, industrial and
domestic use).

On the other hand, the availability of water resources must be contrasted with an increasing
world population. The world population in 1900 was 1.6 billion, increasing to 2.5 billion by
1950 and 6.1 billion by the year 2000. Despite a general decline in fertility rates, world
population is still growing rapidly. It is projected that world population will reach between
7.5 and 10.5 billion by 2050, depending on future growth rate scenarios. The population in
the LAC region should almost reach 700 million people by 2025 from 475 million in 1997.
The world population is expanding rapidly, with corresponding increased pressures on the
food supply and the environment. Competition for water is becoming critical, and
environmental degradation related to water usage is serious. The number of people living in
water-stressed countries is projected to climb from 500 million to three (3) billion by 2025.

Kenya is classified as a chronically water-scarce country. The country’s natural endowment


of freshwater is limited by an annual renewable freshwater supply of only 647 m³ per capita.
Globally, a country is categorized as “water-stressed” if its annual renewable freshwater
supplies are between 1,000 and 1,700 m³ per capita and “water-scarce” if its renewable
freshwater supplies are less than 1,000 m³ per capita (World Bank, 2000). Only 8.3% of
countries in the world are classified as water-scarce, while 9.8% of the countries are
considered water-stressed. By comparison, Kenya’s neighbours, Uganda and Tanzania have
annual per capita renewable freshwater supplies of 2,940 and 2,696 m³ per capita per year
respectively.

Agriculture is the major user of freshwater, with a world’s average of 71% of the water use.
There are large regional variations, from 88% in Africa to less than 50% in Europe. In Kenya
Irrigated agriculture is the largest water-using sector. Demand is predicted to rise from 3,965
million m³/day in 1990 to 8,138 million m³/day in 2010 – double the level in 1990. These
estimates are based on current water use efficiencies. Actual demand could be significantly
less if efficiency incentives are introduced. In the LAC region, 73% of the water use is for
agricultural purposes. However, the large and growing proportion of the population living in
urban areas will put considerable pressure for continued transfers of water out of agriculture
to supply growing urban centers. Other competing uses include hydroelectricity, protection of
aquatic ecosystems (e.g., environmental flows, restoration of native salmon runs in the US
Pacific Northwest), and recreation.

Meeting the crop demands projected for 2025 could require an additional 192 cubic miles of
water, a volume nearly equivalent to the annual flow of the Nile 10 times over. No one yet
knows how to supply that much additional water in a way that protects supplies for future use
and minimizes environmental impact. Severe water scarcity presents the single biggest threat
to future food production. Even now many freshwater sources (underground aquifers and
rivers) are stressed beyond their limits. As much as 8 % of food crops grow on farms that use
groundwater faster than the aquifers are replenished, and many large rivers are so heavily
diverted that they don't reach the sea for much of the year. As the number of urban dwellers

156
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

climbs to five billion by 2025, farmers will have to compete even more aggressively with
cities and industry for shrinking resources.

17.3 Sources of Environmental Impacts in Irrigated Agriculture


The benefits of irrigation have resulted in lower food prices, higher employment and more
rapid agricultural and economic development. The spread of irrigation has been a key factor
behind the near tripling of global grain production since 1950. But irrigation and water
resource development can also cause social and environmental problems.

Irrigation represents an alteration of the natural conditions of the landscape by extracting


water from an available source, adding water to fields where there was none or little before,
and introducing man-made structures and features to extract, transfer and dispose of water.
Irrigation projects and irrigated agriculture practices can impact the environment in a variety
of ways. For this chapter we will distinguish the following sources of environmental impact:
i) construction of irrigation projects, ii) water supply and operation of irrigation projects, and
iii) irrigated agriculture management practices.

17.3.1 Environmental impact derived from the construction of irrigation projects


Before 1900 only 40 reservoirs had been built with storage volumes greater than 25 billion
gallons; today almost 3,000 reservoirs larger than this inundate 120 million acres of land and
hold more than 1,500 cubic miles of water. The more than 70,000 dams in the U.S. are
capable of capturing and storing half of the annual river flow of the entire country.

In many nations, big dams and reservoirs were originally considered vital for national
security, economic prosperity and agricultural survival. Until the late 1970s and early 1980s,
few people took into account the environmental consequences of these massive projects.
Today, however, the results are clear: dams have destroyed the ecosystems in and around
countless rivers, lakes and streams. On the Columbia and Snake rivers in the north-western
U.S., 95 % of the juvenile salmon trying to reach the ocean do not survive passage through
the numerous dams and reservoirs that block their way. More than 900 dams on New England
and European rivers block Atlantic salmon from their spawning grounds, and their
populations have fallen to less than 1 % of historical levels. Perhaps most infamously, the
Aral Sea in central Asia is disappearing because water from the Amu Darya and Syr Darya
rivers that once sustained it has been diverted to irrigate cotton. Twenty-four species of fish
formerly found only in that sea are currently thought to be extinct.

Water is allocated in Kenya through permits. The responsibility for issuing these permits is
vested with the Water Apportionment Board, based on advice from District Water Boards,
Water Catchment Boards, and technical advice from the Director of Water Development.
Even when water is being abstracted legally, it can place a heavy burden on the available
resource. Thus, the horticultural industry around Lake Naivasha is reliant on large volumes of
lake water. The level of the lake fluctuates naturally because of changes in inflows and
evaporation losses. Even so, over a 14-year period beginning in the mid-1980s, modelling
results show a steady reduction in the lake’s storage capacity of about 800 million m³. This
loss is coincident with the abstraction for irrigation use from the lake and its associated
aquifers. In certain Kenyan districts there has been excessive abstraction of water from rivers.
The consequences are very apparent in Baringo, Laikipia and Nyandarua districts, where
downstream water users are denied essential water, important lakes and wetlands are severely
threatened, and aquatic biota on which the poorer population depends, such as fish, are in

157
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

decline. In Laikipia District, there has been a big increase in water use over the past decade.
The increase was mainly due to increased human settlement and expansion of irrigated
agriculture. The increased water use has led to severe social, environmental, and economic
problems.

The development of irrigation projects results in an alteration of the current condition of the
landscape. Depending on the nature of the projects, many questions regarding environmental
impact may arise. A few examples follow:
1. What is the social impact of relocating inhabitants of a given area to accommodate a new
irrigation project (for example, relocating those living on the area to be inundated by a
new reservoir)?
2. What is the impact of the new project on wildlife, particularly endangered species, and on
archaeological patrimony?
3. What is the impact of infrastructure associated with the construction and operation of the
project (roads, power lines, canals, etc)?
4. Will reclaimed and/or recycled construction materials be used, including aggregate,
lumber, and asphalt?
5. Will construction materials used be reclaimed and reused in future projects rather than
being disposed of?
6. Are there alternative materials available to reduce hazardous and toxic materials use
during construction?
7. Does the construction plan provide for erosion and sediment control, does it minimise the
disturbance of vegetation and soil, and does it include revegetation of disturbed areas?
8. Does the project use existing structures to the extent possible and avoid sensitive
habitats?
9. Will seepage be minimised or eliminated by selecting canal and ditch materials that
prevent seepage?

17.3.2 Environmental impact derived from water supply and operation of irrigation
projects
Irrigated agriculture depends on supplies from surface or ground water. The environmental
impact of irrigation systems depends on the nature of the water source, the quality of the
water, and how water is delivered to the irrigated land.

Withdrawing ground-water may cause the land to subside, aquifers to become saline, or may
accelerate other types of ground-water pollution. Withdrawing surface water implies changes
to the natural hydrology of rivers and water streams, changes to water temperature, and other
alterations to the natural conditions, sometimes deeply affecting the aquatic ecosystems
associated with these water bodies.

An excessive withdrawal of water for irrigation is clearly impacting the environment in some
areas. For example, the Colorado river often contains essentially no water by the time it
crosses the border into Mexico, owing to both urban and agricultural withdrawals. In fact, in
most years, the Colorado River doesn't make it to the ocean. This has consequences for the
river and its riparian ecosystems, as well as for the delta and estuary system at its mouth,
which no longer receives the recharge of fresh water and nutrients that it normally did. The
same is true for the Yellow River in China. The San Joaquin River in California is so
permanently dewatered that trees are growing in its bed and developers have suggested
building housing there. In the last 33 years, the Aral Sea has lost 50% of its surface area and

158
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

75% of its volume, with a concomitant tripling in its salinity, owing largely to diversion of
water from its feeding rivers for irrigating cotton. Groundwater is being mined. In the US,
about 20% of the irrigated agricultural area is watered by pumping in excess of recharge.
Texas has lost 14% of its irrigated acreage since 1980 as a result of aquifer depletion. In
India's Punjab, pumping exceeds recharge by 1/3, causing water tables to drop by 1 m/yr or
more.

Salinity in waters is a natural occurrence because of weathering of saline parent materials


derived from seawater deposits and other sources. The quality of the water supplies varies,
but water sources with significant salinity will impact the quality of the irrigated land and the
sustainability of agricultural production supported by this land, particularly when poor
management is present.

In 1982, salinization affected about 196,550 square km (0.7 %) of the agricultural soils in
Central America and Mexico, and 1,291,630 (7.6 %) in South America. Argentina and Chile
have about 35 % of their irrigated lands affected by salinity whereas 30 % or 250,000 ha, of
the coastal region of Peru under irrigation is impacted by this problem. In Brazil, 40 % of the
irrigated land in the northeast is affected by salinity as a result of improper irrigation. Natural
and man-induced salinity in Cuba covers about 1.2 million ha. In Mexico, about 12.4 % of
the country's irrigated acreage, was reported wholly or partially affected by salinization in
1980.

Due to water shortages, contaminated wastewaters are frequently used for irrigation. For
example, since the beginning of the century, about 90,000 ha of agricultural land in the Tula
Valley has been irrigated with wastewater from Mexico City. In Lima, 2,000 ha of vegetable
crops are irrigated with urban wastewaters. In Sao Paulo, the contaminated waters from the
Tietê River are used to irrigate vegetable gardens downstream from the urban core. In
Santiago, 62,000 ha of vegetables used to be grown using water from three courses located
downstream from Santiago’s sewage outflow. The reuse of wastewater for irrigation is
imperative in the Middle East. It is also the most popular form of wastewater and nutrient
recycling from dairy and other confined animal operations. The operation of irrigation water
supply systems can affect the environmental performance of irrigated agriculture. Systems
that deliver water continuously or in a fixed schedule are less efficient and/or limit
management options available for irrigators compared to on-demand water delivery
operations.

The operation and management of irrigation water delivery systems must include proper
monitoring and reduction of seepage and other water losses in the system, particularly if they
are a significant component of recharge of raising water tables. The combination of low water
quality supply and raising water tables will eventually lead to waterlogging and salinization,
threatening the sustainability of existing irrigation systems. Proper attention to the quality and
amount of irrigation return flows is also important to identifying and mitigating possible
impacts on receiving waters.

Questions regarding the irrigation water supply and environmental impacts that may arise
include:
1. Are there provisions to ensure that the quality of the supplied water does not contribute to
salinity build-up on the irrigated land?

159
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

2. Will groundwater extraction rates be kept at or below recharge rates to prevent drawdown
and related subsidence and habitat destruction?
3. Will surface water diversion reduce groundwater discharge?
4. Will the diversion rate have an adverse effect on downstream flow rates or downstream
water temperature?
5. Will water distribution systems and management be conducive to the implementation by
farmers of sound irrigation and agronomic practices that minimize the environmental
impact of irrigation?

17.3.3 Environmental impact derived from irrigated agriculture management practices


In the US, agriculture is the leading source of water quality impairment of rivers and lakes,
and third in importance for pollution of estuaries. Close to three quarters (72%) of river
length and 56% of lake area assessed by the US Environmental Protection Agency are
impacted by agriculture. Both dry land and irrigated agriculture contribute to this situation.

In addition to problems of waterlogging, salinization, and erosion that affect irrigated areas,
the problem of downstream degradation of water quality by salts, agrochemicals and toxic
leachates is a serious environmental problem. Salinization of water resources is possibly a
greater concern to the sustainability of irrigation than is that of salinization of soils, per se.

Six thousand years ago farmers in Mesopotamia started diverting water from the Euphrates
River. Thus, Sumerians went on to form the world's first irrigation-based civilization.
However, Sumeria was one of the earliest civilizations to crumble in part because of the
consequences of irrigation. Sumerian farmers harvested plentiful wheat and barley crops for
some 2,000 years, but the soil eventually succumbed to salinization. Many historians argue
that soil salinization and the decline in food supply figured prominently in the society's
decline.

The social, economic and ecological disaster that has occurred in the Aral Sea and its
drainage basin since the 1960s is the world’s largest modern example of how poorly planned
and poorly executed agricultural practices have devastated a once productive region. The
Aral Sea basin includes Southern Russia, Uzbekistan, Tadjikistan, and part of Kazakhstan,
Kirghiztan, Turkmenistan, Afghanistan, and Iran. Although there are many other impacts on
water quality in the region, improper agricultural practice is the root cause of this disaster.
Virtually all agriculture is irrigated in this arid area. Elements contributing to the problem
include increase in irrigation area and water withdrawals, use of unlined irrigation canals,
rising groundwater, extensive monoculture and excessive use of persistent pesticides,
increased salinization and salt runoff leading to salinization of major rivers, increased
frequency of dust storms and salt deposition, discharge of highly mineralized, pesticide-rich
return flows to main rivers, and excessive use of fertilizers. Some of the environmental
impacts include decline in Aral Sea levels by 16 m since 1960, decline of the Aral Sea
volume by over 70%, destruction of commercial fishery, major decline and extinction of
animal, fish, and vegetation species, salt content of major rivers exceeding standards by
factor of 2 to 3, high levels of turbidity in major water sources, high levels of pesticides and
phenols in surface waters, and several others.

The management of water application systems as well as the suitability of related agronomic
practices has a dramatic influence on the environmental impact of irrigated agriculture.
Constraints in the water delivery systems (e.g., continuous versus on-demand water supply),

160
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

extremely low water quality of the irrigation water supply, and limitations to investment on
improved technologies exacerbate the environmental damage derived from irrigation and
limit the options available to farmers for mitigating the problem. Regardless of the nature of
the problem, improved management and technologies are the main tools available to ensure
the sustainability and productivity of irrigated lands.

Salinity
All irrigation water contains dissolved salts derived as it passed over and through the land,
and rain water also contains some salts. These salts are generally in very low concentration in
the water itself. However, evaporation of water from the dry surface of the soil leaves the
salts behind. Salinization is especially likely to become a problem on poorly drained soils
when the groundwater is within 3 m or less of the surface (depending on the soil type). In
such cases, water rises to the surface by capillary action, rather than percolating down
through the entire soil profile, and then evaporates from the soil surface.

Salinization is a worldwide problem, particularly acute in semi-arid areas, which use large
amounts of irrigation water and are poorly drained. These conditions are found in parts of the
Mideast, in China's North Plain, in Soviet Central Asia, in the San Joaquin Valley of CA, and
in the Colorado River Basin. Salinization reduces crop productivity. In the US, salinization
may be lowering crop yields on as much as 25-30% of the nation's irrigated lands. In Mexico,
salinization is estimated to be reducing grain yields by about 1 million tons per year, or
enough to feed nearly million people. In extreme cases, land is actually being abandoned
because it is too salty to farm profitably.

Salinity has been associated with irrigated agriculture since its early beginnings. One reason
is that irrigation often exacerbates the effects of salinity, which occurs naturally. Estimates
indicate that roughly one-third of the irrigated land in the major irrigation countries is already
badly affected by salinity or is expected to become so in the near future. Present estimates for
India range from 27% to 60% of the irrigated land, Pakistan 14%, Israel 13%, Australia 20%,
China 15%, Iraq 50%, Egypt 30%. Irrigation-induced salinity occurs in large and small
irrigation systems alike. In recent years, many farmers have been abandoning their rice fields
in Sahelian irrigation schemes due to the incidence of salinity.

Salinity is often linked with the rise of groundwater tables resulting from excess irrigation
and poor drainage in large-scale, perennial irrigation systems. The resulting shallow water
tables bring salts to the upper layers of the soil profile. That salinity can also be induced by
the use of pumped groundwater of marginal or poor quality has been realized only more
recently. In these cases, the physical process underlying salinization is the absence of a
downward soil water flux of sufficient magnitude to leach the salts from the root zone.

Saline soils contain sufficient soluble salts to adversely affect the growth of most plants. With
a predominance of sodium on the exchange complex and a low concentration of salts in the
infiltrating water, the infiltration rate and permeability can be severely, and in some cases,
irreversibly reduced. Leaching and drainage cause salt loading of the water resource into
which the effluent is discharged. The volume needing disposal can be reduced through
improved irrigation management and reuse of drainage outflow for irrigation.

The technical problems that have led to irrigation-induced salinity include poor on-farm
water use efficiency; poor construction, operation and maintenance of irrigation canals

161
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

causing excessive seepage losses, and inadequate or lack of drainage infrastructure or, if
drainage facilities are present, their poor quality of construction, operation, and maintenance
(e.g., drainage systems in the Lluta Valley, Chile).

Waterlogging
Waterlogging usually results from overuse and/or poor management of irrigation water.
Lining and covering of water conduits from the storage dams to the point of delivery
improves water usage and at the same time reduces the risk of a rise in the water-table in
many irrigated areas. This procedure needs to be applied to the 11 million hectares of land in
Asia that have been degraded through waterlogging, but it would also benefit areas suffering
from salinization. Waterlogging and salinization impacts can be further reduced in most cases
by more investment in education and management capacity rather than in drainage and soil
improvement works.

Worldwide, about 10% of all irrigated land suffers from water logging. As a result,
productivity has fallen about 20% in this area of cropland. Drainage problems affect large
areas of land in Latin America; in many cases these problems are compounded by
salinization. Thus, in Argentina 555,000 ha are in need of drainage. In Peru 60,000 ha in the
coastal region and 34% of the cultivated lands in the upper jungle (Ceja de Selva) - or
150,000 ha - are affected by drainage problems; and in Costa Rica, projects for rehabilitation
through drainage exceed 60,000 ha.

Agricultural runoff
One of the main non-point sources of water pollution is runoff from agriculture. Runoff of
agricultural chemicals is primarily a localized problem where agricultural input use is high.
The consumption of fertilizers in LAC has increased rapidly over the last 30 years, from 16
kg per ha in 1966 to 62 kg per ha in 1996. However, fertilizer application is still below the
levels of developed countries. In 1996, the consumption of fertilizers per ha of farmland
amounted to 89 kg globally and 113 kg in the United States. However, the experience in LAC
is diverse. In a few countries, application rates are extremely low, including Bolivia with 5
kilogram per ha and Haiti with 9 kilogram per ha in 1996. In other countries of the region,
however (for example, Chile, Colombia, Costa Rica, El Salvador or Uruguay) the
consumption of fertilizers is similar to that of developed countries.

Not only nutrients and other chemicals are transported with irrigation runoff waters. Soil
erosion and subsequent transport of sediments (and adsorbed chemicals) is caused by runoff
of excess irrigation water from cropland. Soil erosion decreases the productivity of the land.
Furrow irrigation causes more erosion than sprinklers or drip irrigation. Sediments
transported by irrigation tail waters eventually return to streams and rivers, negatively
impacting canals and other water conveyance structures, causing sedimentation of reservoirs
and other structures, affecting the durability and uniformity of sprinkler and drip irrigation
systems, and creating significant problems to fish habitat and aquatic ecosystems.

The impact of agricultural runoff on water quality is well documented in the US and other
countries. Data on water pollution in developing countries are limited. Further, such data are
mostly aggregated, not distinguishing the relative proportion of point and non-point sources.
The water quality database that is available often is of little value in pollution management at
the river basin scale, and is not useful for determining the impact of agriculture relative to
other types of anthropogenic impacts.

162
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Impact on groundwater
Infiltration of irrigation water in excess of available root zone storage will penetrate beyond
the reach of roots and eventually recharge groundwater. Nitrates, salts, and other chemicals
dissolved in the soil water will move with the water. Crops with high water and N
requirements tend to increase the potential risk of nitrate pollution to groundwater. Light-
textured soils and intensive production of shallow-rooted crops under irrigation can lead to
considerable nitrate losses by leaching.

In the US, nitrate in groundwater is highest in areas of well-drained soils and intensive
cultivation of row crops. Areas such as the Northeast, Great Plains and along the West Coast
have been found to have high nitrate concentrations in groundwater. In the Northwest, the
Quincy-Pasco area of Washington State is reported as highly polluted by N, with non-point
sources from agriculture often cited as a significant contributing factor. This area, cultivated
with row crops on well-drained soils, receives high rates of irrigation and fertilizer
application. In the Central Columbia Plateau of Washington State, fertilizers are the source of
84 % of nitrate inputs. Other sources of nitrate include cattle feedlots, food-processing plants,
septic tanks, and treated wastewater.

Data on groundwater pollution in developing countries, resulting from excess chemical input
and irrigation, is not well documented, but it is likely to show an increasing trend as irrigated
agriculture worldwide becomes more intensively managed. According to various surveys in
India and Africa, 20 to 50 % of wells contain nitrate levels greater than 50 mg/l and in some
cases as high as several hundred mg/l. In some developing countries, it is wells in villages or
close to towns that often contain the highest N levels, suggesting that domestic excreta are the
main source, though livestock wastes are particularly important in semi-arid areas where
drinking troughs are close to wells.

Public health impacts


Polluted water is a major cause of human disease. According to the World Health
Organization, as many as 4 million children die every year as a result of diarrhea caused by
water-borne infection. The bacteria most commonly found in polluted water are coliforms
excreted by humans. Surface runoff and improperly designed rural sanitary facilities
contribute to this problem.

The use of untreated (or poorly treated) human wastewater for irrigation purposes contributes
to direct contamination of food. In many developing countries there is little or no treatment of
municipal sewage, yet urban wastewater is increasingly being used for irrigation. The most
common diseases associated with contaminated irrigation waters are cholera, typhoid,
ascariasis, amoebiasis, giardiasis, and enteroinvasive E. Coli. Crops that are most implicated
with spread of these diseases are ground crops that are eaten raw such as cabbage, lettuce,
strawberries, etc.

Nitrogen levels in groundwater have grown in many parts of the world as a result of
intensification of farming practice, particularly in irrigated lands. Nitrate levels have grown in
some countries to the point where more than 10% of the population is exposed to nitrate
levels in drinking water that are above the 10 mg/l guideline designed to prevent
methaemoglobinaemia to which infants and elderly are particularly susceptible. Although the

163
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

problem is less well documented, nitrogen pollution of groundwater appears also to be a


growing problem in developing countries.

There is a linkage between increase in malaria in several Latin American countries and
reservoir construction. Schistosomiasis, a parasitic disease affecting more than 200 million
people in 70 tropical and subtropical countries, has been demonstrated to have increased
dramatically in the population following reservoir construction for irrigation and
hydroelectricity.

Questions regarding environmental performance of irrigated agriculture management


practices
Improved field irrigation practices are a critical factor in mitigating the effect on the
environment of existing irrigation projects. Many questions may arise regarding the
environmental performance of field irrigation (and associated agronomic) practices. A few
examples follow:
1. Will measures be taken to limit pesticide, herbicide, and petroleum-based fertilizer use,
including using organic fertilizer, planting pest-resistant crops, and alternating crops and
planting cycles?
2. Will soil loss prevention measures be taken to prevent wind and water erosion, including
planting vegetative windbreaks, practicing contour ploughing, and maintaining soil
moisture? These practices reduce offsite sedimentation, nutrient pollution, and water
quality degradation.
3. Will irrigation periods be restricted to evenings, nights, and early mornings to prevent
excessive water loss due to evaporation and reduce peak power demands?
4. Will water application rates be minimized to prevent surface runoff, over watering, and
nutrient leaching?
5. Will soil moisture content, temperature, humidity, time of day, wind, and
evapotranspiration rate be considered in determining the most efficient time to irrigate?
6. Have drip irrigation and other water conserving application techniques been evaluated for
implementation?
7. Do water cost and allotment factors encourage conservation? Are financial incentives or
disincentives, such as an increasing per unit cost, provided to promote conservation?
8. Will agricultural runoff be managed to prevent impacts from excess nutrients and
chemical pesticides and herbicides?
10. Will filter strips or other methods be used to remove sediments, organic matter, and other
pollutants from runoff and waste water?
11. Can the drainage water be reclaimed and reused?
12. Will water from subsurface drainage systems be evaluated for contaminant levels and the
best management practice selected for its handling?

17.4 How can we mitigate the Environmental Impact of Irrigation?


As the world's population continues to grow, dams, aqueducts and other kinds of
infrastructure will still have to be built, particularly in developing countries where basic
human needs have not been met. But such projects must be built to higher standards and with
more accountability to local people and their environment than in the past. And even in
regions where new projects seem warranted we must find ways to meet demands with fewer
resources, minimum ecological disruption and less money.

164
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

FAO has estimated that the potential exists, based on physiography and soil conditions, for an
eventual total of 400 million hectares of irrigated land, three-quarters of which would be in
the developing countries. Irrigated areas are 2.5 times more productive than rain-fed
agricultural land, and there is a strong presumption that their extent (some 300 million
hectares at present) will increase. However, expansion beyond present levels is constrained
by the shortage of suitable land, limited water supplies and the high cost of installing large-
scale irrigation schemes. In many cases it is more effective to improve the management and
production efficiency of existing irrigated areas than to open up new irrigation schemes.

The largest single consumer of water is agriculture, and this use is largely inefficient. Water
is lost as it is distributed to farmers and applied to crops. Consequently, as much as half of all
water diverted for agriculture never yields any food. Thus, even modest improvements in
agricultural efficiency could free up large quantities of water. Growing tomatoes with
traditional irrigation systems may require 40 % more water than growing tomatoes with drip
systems. Even our diets have an effect on our overall water needs. Growing a pound of corn
can take between 100 and 250 gallons of water, depending on soil and climate conditions and
irrigation methods. But growing the grain to produce a pound of beef can require between
2,000 and 8,500 gallons.

Shifting where people use water can also lead to tremendous gains in efficiency. Supporting
100,000 high-tech California jobs requires some 250 million gallons of water a year; the
same amount of water used in the agricultural sector sustains fewer than 10 jobs, a stunning
difference. Similar figures apply in many other countries. Ultimately these disparities will
lead to more and more pressure to transfer water from agricultural uses to other economic
sectors. Unless the agricultural community embraces water conservation efforts, conflicts
between farmers and urban water users will worsen.

New approaches to meeting water needs will not be easy to implement: economic and
institutional structures still encourage the wasting of water and the destruction of ecosystems.
Among the barriers to better water planning and use are inappropriately low water prices,
inadequate information on new efficiency technologies, inequitable water allocations, and
government subsidies for growing water-intensive crops in arid regions or building dams.
Several types of interventions aimed at preventing, mitigating, or reversing soil and water
degradation at various levels within irrigated agriculture are possible. Some are applicable at
field or farm level, others at system, regional, or sub-regional level. Examples of possible
interventions are given below, categorized as policy, engineering, system management, and
irrigation/agronomic practice interventions.

17.4.1 Policy Interventions


1. Introduce water and power pricing that better represents the market value of water;
2. Introduce transferable water entitlements;
3. Set limits for allowable groundwater recharge (amount and quality) and introduce
penalties for exceeding these limits;
4. Provide incentives for land reclamation;
5. Require exhaustive environmental impact assessment for new irrigation projects; and
6. Provide incentives for monitoring and reduction of the environmental impact of existing
irrigation projects.

165
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

17.4.2 Engineering interventions


1. Incorporate environmental impact considerations in the design, construction, and
operation of new irrigation projects;
2. Improve maintenance of irrigation infrastructure;
3. Construct drainage facilities;
4. Improve maintenance of existing drains;
5. Reuse waste and drain water, and find alternative ways to dispose drainage effluent; and
6. Prevent or reduce canal seepage, i.e., through lining.

17.4.3 System management interventions


1. Improve the operation of existing irrigation and drainage infrastructure through
introduction of management information systems, etc;
2. Enhance farmers’ involvement in management and maintenance of irrigation and
drainage facilities; and
3. Evaluate the feasibility of implementing on-demand water delivery to farms.

17.4.4 Irrigation/agronomic practices interventions


1. Minimize water losses in the on-farm distribution system;
2. Improve irrigation systems performance to minimize deep percolation and surface runoff;
3. On-farm watercourse improvement and precision land levelling;
4. Implement more efficient irrigation methods (e.g. drip instead of surface irrigation);
5. Minimize sediment concentration in runoff water;
6. Grow different crops or introduce different crop rotations (i.e., less-water demanding
crops, more drought- and salt-tolerant crops);
7. Irrigate according to reliable crop water requirement estimates and leaching requirement
calculations;
8. Manage fertilizer programs so as to minimize nutrients available for detachment and
transport; and
9. Apply soil amendments and reclamation practices.

17.5 Final Remarks


The potential to increase substantially the irrigated area of the world is limited; however for
Kenya the potential is high. Globally gains from new capacity are expected to be largely
offset by losses such as waterlogging and salinization, as well as retirement of areas being
irrigated by pumping water in excess of rates of recharge. In fact, most new water capacity is
predicted to come from increasingly efficient use of existing supplies rather than harnessing
of new supplies. On the other hand, managing existing irrigation projects so as to minimize
their environmental impact is a requirement for long-term sustainability of irrigated
agriculture. Both, improved water use efficiency and environmental stewardship are indeed
complementary goals.

One area of concern is that developing countries could spend the next 50 years struggling to
provide safe drinking water and sanitation to their exploding urban populations and enough
irrigation water to maintain the high levels of food production needed to provide improved
diets to a growing population, at the expense of their ability to restore and maintain their
already damaged aquatic ecosystems.

The unresolved global warming issue could turn to be the major challenge for water
development during the next 50 years, or it may not be. There is so much uncertainty in the

166
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

scientific data and supporting models that it makes effective analysis very hard. Taken
together, the current state-of-the-science suggests a wide range of potential concerns that
should be addressed by national and local water managers and planners, climatologists,
policymakers, and the public. Prudent decision-making requires consideration of potential
climate change scenarios on long-term decisions regarding water use and environmental
impact.

References
Dougherty, T.C., Hall, A.W. and Wallingford, H.R. 1995. Environmental impact assessment
of irrigation and drainage projects. Irrigation and Drainage Paper 53, Food and Agriculture
Organization of the United Nations, Rome.
Gleick, P.H. 2000. Water: The potential consequences of climate variability and change for
the water resources of the United States. National Water Assessment Group. USGS and
Pacific Institute for Studies in Development, Environment and Security, Oakland, CA.
Gleick, P.H. 2001. Making every drop count. Scientific American, Feb 2001.
Kijne, J.W. 1996. Water and salinity balances for irrigated agriculture in Pakistan. Research
Report 6. International Irrigation Management Institute, Colombo, Sri Lanka.
Kijne, J.W., Prathapar, S.A., Wopereis, M.C.S. and Sahrawat, K.L. 1988. How to manage
salinity in irrigated lands: A selective review with particular reference to irrigation in
developing countries. SWIM Paper 2. International Irrigation Management Institute,
Colombo, Sri Lanka.
Ongley, E.D. 1996. Control of water pollution from agriculture. Irrigation and Drainage
Paper 55, Food and Agriculture Organization of the United Nations, Rome.
Postel, S. 2001. Growing more food with less water. Scientific American, Feb 2001.
Ringler, C., Rosegrant, M.W. and Paisner, M.S. 2000. Irrigation and water resources in Latin
America and the Caribbean: Challenges and strategies. International Food Policy Research
Institute, Washington D.C.
Rogers, P. 2000. Water Resources in the twenty and one half century: 1950-2050. Water
Resources Update 116, Universities Council on Water Resources.
Seckler, D., Amarasinghe, U., David, M., de Silva, R. and Barker, R. 1998. World water
demand and supply, 1990 to 2025: Scenarios and issues. Research Report 19. International
Irrigation Management Institute, Colombo, Sri Lanka.
State of Washington Department of Ecology and Washington State University. 1995.
Irrigation management practices to protect groundwater and surface water quality.
Washington State University Cooperative Extension, Pullman, Washington.
Williamson, A.K., Munn, M.D., Riker, S.J., Wagner, R.J., Ebbert, J.C. and Vanderpool, A.M.
1998. Water quality in the Central Columbia Plateau, Washington and Idaho, 1992-95.
Circular 1144, US Geological Survey.
World Bank. 2000. World Development Report 2000/2001: Attacking Poverty. Washington,
D.C.

Internet Resources
Human impacts on Ecosystems (www.orst.edu/instruction/bi301)
International Water Management Institute (www.cgiar.org/iwmi)
International Water Resources Association (www.iwra.siu.edu)
Pacific Institute for Studies in Development, Environment and Security (www.pacinst.org)
The World’s Water, Information on the World Freshwater Resources ( www.worldwater.org)
United Nations Food and Agricultural Organization, Land and Water Development Division
(www.fao.org/ag/AGL)

167
CVS 545E: Irrigation Engineering, Lecture Notes, and Instructor: Prof. Eng. E.C. Kipkorir
Department of Civil and Structural Engineering, Moi University
September 12th, 2017

Universities Council on Water Resources (http://www.uwin.siu.edu/ucowr)


Pollution prevention/environmental impact reduction checklist for agricultural irrigation
(http://es.epa.gov/oeca/ofa/pollprev/agric.html)

168

You might also like