Irrigation Engineering Principles
Irrigation Engineering Principles
Irrigation Engineering Principles
Gary P. Merkley
Copyright Notice
This material has been duplicated by special permission of the copyright holder.
It is not to be duplicated or used for purposes other than learning resource support for
Utah State University. Any violation of this agreement is punishable under existing
copyright laws.
Page
28.35 g/oz
15.85 gpm/lps (= 60/3.785)
7.481 gallons/ft3
448.86 gpm/cfs (= 7.481*60)
3.7854 liter/gallon
6.89 kPa/psi
1 cbar = 1 kPa
10 mbar/kPa, or 100 kPa/bar
2.308 ft/psi, or 9.81 kPa/m (head of water)
14.7 psi = 101.3 kPa = 10.34 m (head of water) = 1,013 mbar = 1 atm
62.4 lbs/ft3, or 1,000 kg/m3 (max density of pure water at 4C)
0.1333 kPa/mmHg
0.7457 kW/HP
1 langley = 1 cal/cm2
0.0419 MJ/m2 per cal/cm2
0.3048 m/ft
1.609 km/mile
2.471 acre/ha
43,560 ft2/acre
1,233 m3/acre-ft
57.2958 degrees/radian
3.14159265358979323846
e 2.71828182845904523536
C = (F 32)/1.8
F = 1.8 C + 32
I. Course Overview
No required textbook
Bound lecture notes are required
Some material will be taken from Irrigation
Fundamentals, by Hargreaves & Merkley, as well as
from several other books and sources
Some material will be on reserve in the USU
SciTech Library
IV. Tests
Irrigation has been practiced for more than 6,000 years on this planet and was
critical to the development of some early civilizations, as well as to the
sustainability and progress of modern civilization
Irrigation has been practiced in areas of the middle east and Egypt for
thousands of years, and for over one thousand years in parts of Europe and the
Americas
Many parts of Asia also have evidence of over
one thousand years of agricultural irrigation
Remains of a 6,000-year-old rice paddy with
vestiges of irrigation channels connecting it to a
well have been discovered in China
The first irrigation was by gravity diversion and from water lifters powered by
humans, animals, or by the flow of water
There are various definitions for irrigation, each depending on the perspective or
discipline from which it is derived
However, a common and general definition for irrigation is:
1. to add water to the plant root zone for plant growth (this is usually the
main objective);
2. to cool the soil and or air around the plants, creating a more favorable
micro-environment for plant growth;
3. to reduce or eliminate the detrimental effects of short-duration frost
(using sprinklers);
4. to maintain the relative humidity of the micro-climate around the plants;
5. to leach or dilute salts in the soil;
6. to soften hard pans and soil clods;
7. to delay bud formation through evaporative cooling;
8. to dispose of effluent from farms and ranches; and,
9. to apply fertilizers or soil amendments.
The above are not all of the objectives that can be cited for irrigation, and only a
few of those mentioned above would be applied in any given case, but
collectively they represent the major reasons for irrigating in agricultural areas
There are roughly 270 million irrigated hectares in the world today; this comes to
about 22 people per irrigated hectare
This irrigated area is about percent of the surface area of the planet, or 2% of
the total land area (excluding Antarctica)
Land and permanent ice covers about 27% of the surface of the earth
The pie chart below shows approximate percentages of irrigated areas, by
continent
Since 1950, the global irrigated area has kept pace with global population, so the
irrigated area per capita has not changed much over the past 50 years (about 22
people per irrigated hectare)
There has been a more than 100% increase in total food production in the
developing world during the past 30 years, but per capita agricultural production
More than 95% of the irrigated area in the world is by surface methods, and more
than half of the irrigated area is found in five countries: India, China, USA,
Russia, and Pakistan
Thus, most of the irrigated lands in the world are in Asia, where some 90% of the
world rice production area is located and approximately half of the population is
employed in the agricultural sector
There are also large areas in the world with agricultural production under rainfed,
or nonirrigated, conditions
Many areas in southeast Asia and the Indian subcontinent rely on rain for irrigation
water, but these regions also have hundreds of irrigation projects that supplement
rain and allow for dry season production
Even some of the rainiest regions of the world have periods of no rain and
occasional years of relative drought, and irrigation systems in these regions can
provide for much more consistent agricultural
production
Egypt is about 15th in terms of irrigated area, but
unlike the other countries with a relatively large
irrigated area, Egypt relies almost exclusively on
irrigation for agricultural production
Egypt is currently consuming nearly all of its legal share of the Nile River water, but
water development projects are being planned and implemented in upstream
countries, such as Sudan
Brazil is the fifth largest country in the world and is estimated to have about 8% of
the fresh water in the world, yet this country is not among the top ten in terms of
irrigated area
Of the total irrigated area in the world today, about 10% suffers from salinization
serious enough to limit crop yields; the salinization is currently increasing by about
two million hectares per year
Of course, not all food comes from irrigated agriculture:
1. Fisheries;
2. Rangeland; and,
3. Other non-irrigated lands.
The 20th century has had dramatic expansion of the total irrigated area in the world
It is estimated that in the year 1900 about 40 million hectares were irrigated
around the world
???
Irrigated Area (millions of hectares
250
200
150
100
50
0
1800 1850 1900 1950 2000 2050
Year
Thus, during the 20th century there has been a six- or seven-fold increase in
irrigated area
8
7
6
5
4
3
2
1
0
-2000 -1000 0 1000 2000 3000
B.C. A.D.
Postel, S. 1999. Pillar of Sand. W. W. Norton & Company, New York, N.Y. 313 pp.
The USA has some 20 million irrigated hectares (50 million acres), of which about
65% is irrigated by gravity or surface methods (furrows, basins, borders, etc.), and
30% or so is sprinkler irrigated
The USA has some 240,000 irrigated farms
About 17% of the irrigated area in the USA is in California, with Nebraska and
Texas having the second and third largest irrigated areas in this country
The above three states contain nearly half of all the irrigated area in the USA
Almost 90% of the irrigated area in the United States is in the 17 western states
Nebraska has about three times more land under sprinkler irrigation than
California; much of this is irrigated by center pivot
The irrigated area in Utah is about 1,200,000 acres, 2.4% of the national total, and
seven times less than that of California
In the USA, where does the irrigation water come from, and where does it go?
See the figures below
Irrigated agriculture in the western USA typically uses 80% to 85% of all fresh
water diverted or pumped for agriculture, industry, municipalities, and other
purposes
But much of the surface water from rivers and lakes is not diverted, partly
because of environmental concerns much of the water eventually flows out to
the sea
Therefore, irrigated agriculture uses only about 20 - 25% of all surface runoff and
pumped groundwater
The state of California has been the largest agricultural producer in the United
States for the past 50 years
Annual agricultural output is near $25 billion, with over 350 different crops and
commodities
California grows more than half the nations fruits, vegetables, and nuts
Agriculture accounts for nearly 10% of the jobs in the state
Development costs range from about USD $2,000 to $15,000 per hectare, or more
The lowest development costs tend to be in Asia (India in particular) and the
highest costs (by far) are in sub-Saharan Africa
The table below gives data from nearly 200 recent World Bank projects:
USD/ha
Region (average)
South Asia 1,370
Latin America 3,900
East Asia 4,300
North Africa 4,900
Europe 4,750
Middle East 5,100
Sub-Saharan Africa 18,300
Kind of Project
Groundwater 3,800
Gravity 5,600
Average of All 4,800
Drainage systems costs are typically from about USD $1,000 to $2,000 per
hectare
From another source, the global average cost for irrigation and drainage
development is perhaps USD $5,000 to $6,000 per hectare
Poor drainage and flooding are the primary constraints to agricultural productivity
in many areas of the world, and production and development costs are closely
linked to petroleum prices (energy and fertilizers)
Irrigation projects are often designed around an average flow rate capacity of
about one lps per hectare (8.6 mm/day at 24 hrs per day of operation) more on
this in a later lecture
In California, the towns of Santa Barbara and Avalon have begun using
desalinization methods to remove the salt from seawater and make it potable
The high cost of desalinization has kept it from being used more, as it costs
between $1,300-$2,200 per acre-ft to desalinate seawater, compared to about
$200 per acre-ft for water from normal freshwater supply sources
Reverse osmosis is the most common method of desalinization at this time
It is expected that more desalinization of sea water will be practiced in the future,
especially in areas such as California and the Middle East
Recently, there have been detailed proposals for floating desalinization plants
aboard modified tanker ships, with capacities of 50,000 m3/day per ship
It has been claimed that less than 1% of the water treated in municipal systems in
the USA is used for drinking and cooking (a lot goes to watering lawns and
gardens, flushing toilets, showers and baths in homes but even more of it is
used in industry, shops and restaurants)
I. Introduction
The term irrigation system has been applied by different people in various
ways
In one definition, an irrigation system includes the storage facilities that collect
runoff from a watershed, the conveyance and distribution system (canals and
pipelines), the on-farm application systems, and the drainage facilities in a large
irrigated area
In other definitions, one or more of these physical infrastructure components
are considered to be an irrigation system
To some, an irrigation system is considered to be the on-farm or field
application methodology and related equipment, such as pumps, pipes, and
sprinklers
The field application system is used to distribute water over (or under) the
ground surface where it can infiltrate
Irrigation water can be applied to the land in several different ways, and the
choice among alternative irrigation methods depends upon many factors,
including:
1. economics
2. crop type
3. soil type
4. water availability and quality
5. farming practices
6. legal considerations
7. and others
Basin irrigation or level border irrigation is used for the reclamation of saline
soils (leaching) and for a wide variety of soil textures and crops
This is also a surface irrigation method because water is distributed over the
field on the soil surface rather than though pipes
Water application efficiencies with this method can be very high (less than 10%
deep percolation losses) when the field surface is precisely leveled and the flow
rate is sufficient to cover the area quickly
In fact, the application efficiency in a level basin can easily be higher than that
obtainable in most sprinkler systems because there is no surface runoff unless
water is purposely transferred between adjacent basins through small openings
in the dikes
Basins may range in size from very small to quite large (several hectares).
Small basins used in the irrigation of orchards are often referred to as checks
In areas where significant rainfall occurs it is sometimes necessary to provide a
means of surface drainage, and in rice and sugar cane fields the water may be
passed from basin to basin by gravity flow through breaks or checks in the
dikes
For field and vegetable crops a system of ridges and furrows within the basins
provides what is essentially level furrow irrigation - the ridges provide aeration
for the plant roots while soil in the furrows is saturated
The furrow length and initial stream should be regulated so that water will flow
through the furrow rapidly but without erosion and then the flow should be
reduced so that little runoff will occur during the remainder of the irrigation
In furrow irrigation, crops are frequently fertilized by what is called side-
banding
Irrigation in the furrows then moves plant nutrients into the root zone of the crop
as water moves laterally from the furrows
Furrow irrigation is generally used on land with a fairly uniform surface and
slopes of less than one or two percent
However, furrow irrigation has been used extensively in Andean countries on
mountain slopes in which the furrows wind back and forth in the down slope
direction
In these cases, the zigzag furrows are well-formed and deep, with constant
vigilance to prevent severe erosion from a break in the furrow during irrigation
This is just one of many exceptions to the more traditional applications of
irrigation technologies
With natural sub-irrigation, the water table may typically be high due to
geographical and or topographical conditions, and the open channels or buried
drain pipes are used not for supply purposes, but rather to control the phreatic
water level
The phreatic water level is the level under the soil at which the soil water
tension is zero, and this essentially coincides with the water table
This method is used principally with shallow-rooted crops on peat or muck soils
in river deltas, but is also used in medium-texture soils for supplementing direct
rainfall
Less than 2% of the irrigated area in the United States uses this irrigation
method, and much of this area is concentrated in a few large tracts of land that
depend mostly on sub-irrigation for controlling the soil water content of the crop
root zone
Sprinkler irrigation can be restricted to crops that can withstand complete foliar
wetting
Some crops such as tomatoes cannot tolerate wetting when the fruit matures
because it would cause molding and other damage
However, in some cases fungicides and other chemicals are injected into the
irrigation water to prevent quality degradation in the fruit
In orchards it is often necessary to use under tree sprinklers to avoid wetting
the leaves and fruit, but in crops such as bananas a gun sprinkler can be used
for above-canopy supplemental irrigation
Some systems have been provided with automatic hydraulic controls, and
automatic metering devices are frequently used
Sprinklers have been classified by two main groups, set and continuous-move
(Keller and Bliesner 1990)
Sprinkler irrigation is frequently desirable on steep, rocky, or uneven land areas
or where the soils are shallow or too pervious for efficient use of other forms of
irrigation.
Pressurized irrigation may not be desirable where electrical energy or an
alternative energy source may be undependable or expensive
Field operations are easier because much of the soil surface remains dry
Weed growth is reduced, and uninterrupted orchard operations are possible
When crops are grown on beds, the furrows in which farm workers walk remain
relatively dry
Fertilizers are frequently applied in dissolved form through the irrigation water
Herbicides and soil sterilants can also be applied with the irrigation water
I. Introduction
Much of the information in this lecture was taken from Cadillac Desert, by
Marc Reisner (1986), Penguin Books (reprinted in 1993)
It is still available, and costs about $15.00 in paperback
There are many other sources of information on this subject, including books
written about specific projects and specific irrigation systems
There is also a lot of information from Internet sites on the history of the USBR
and other agencies, on irrigation districts and projects, and on regional
development of water resources in the western USA
Enormous amounts of money have been spent on developing water resources
in the western USA over the past 100 years, and there are many strong
opinions about how, where and why it was done
The roots of this development began in the mid-1800s, in something referred to
as Manifest Destiny, whereby the USA was somehow destined to extend the
national boundaries from the Atlantic to the Pacific
The development of irrigated agriculture in the western USA over the past 100
years or so was instrumental in stabilizing the western economy and allowing
the successful implementation of the expansionist concepts
The U.S. Army Corps of Engineers has developed water resources west of the
Mississippi river for more than 200 years, having begun in 1775, and also had
some involvement in water engineering projects in the western USA
But the U.S. Bureau of Reclamation (USBR) was the key player in the
development of irrigated agriculture in the western USA over the past century
It began in 1902 under Teddy Roosevelts administration as the Reclamation
Service after the Reclamation Act became federal law
In 1923 the name was changed to the USBR
When the USBR was created, the federal government had the express intention
of occupying the west and developing it
The USBR built several thousand dams in the western USA from about 1902
1990
In the 20th century, some 50,000 dams were built in the USA, about 2,000 of
which are considered large dams
In its heyday during the first half of the 20th century, the USBR attracted many
of the best engineering graduates from universities, and it was prestigious to
work there
The USBR was a very powerful federal agency in the western USA for many
years in the 1900s, but suffered a steady decline in funding and prestige
beginning about 1970
South
Oregon Dakota
Idaho
Wyoming
Nebraska
Nevada Utah
Nevada Kansas
California
Oklahoma
Arizona New Mexico
Texas
The USBRs current mission statement is quite different from what it was
originally, having shifted from an emphasis on design and construction of major
hydraulic works to operation & management, with a particular focus on
environmental concerns
Most irrigation water withdrawals in the USA are made in the west
Irrigation in Midwest and in the east is only supplementary to rainfall, but this is
not the case in most parts of the west
Most areas in the western USA could not produce agricultural crops without
irrigation, except for some low-value crops like wheat
There was virtually no irrigation in the western USA until about 1850
At that time, many thought the west was basically a waste land with little or no
water
But in the mid-1800s, as westward expansion accelerated, there were some
years of relatively abundant rainfall just west of the Mississippi river, just after
Beginning in 1904, Fred Eaton & William Mulholland of L.A. tricked many
Owens Valley farmers & ranchers into selling agricultural land, then took the
water some 250 miles to L.A. and the San Fernando Valley in aqueducts,
canals, tunnels & siphons
There was no pumping required for the inter-basin water transfer: it all flowed
by gravity from the Sierra Nevada mountains
The irrigated area in the San Fernando Valley increased from 3,000 acres in
1913 to 75,000 acres, while at the same time the irrigated area in the Owens
Valley decreased dramatically, as did the available water
Lake Tahoe
Owens Lake
Los Angeles
San Diego
75% of
precipitation
San Francisco
75% of
population
By the 1920s, Mulholland wanted to build an aqueduct from the Colorado River
to L.A. as he searched for more water
There were many frustrations in this quest; many obstacles
But by this time the value of the agricultural production in southern California
exceeded that of any other region in the USA
Southern California was producing many different kinds of crops, including high-
quality citrus fruit
Land values sometimes went from $10/acre to $500/acre after irrigation water
was made available: the value of agricultural land is still strongly influenced by
the availability of good quality water in sufficient quantity for irrigation
Some farmers sold their land at a great profit
The Colorado river water is nearly 100% allocated, meaning that it releases
almost no water to the sea
The Colorado river delta, in northwest Mexico, used to be a flourishing
agricultural area many years ago, but now it is mostly dry desert
Some privately initiated and funded water projects in the western USA have
become wealthy; for example, the Turlock and Modesto irrigation districts in the
central valley of California
The 1999 annual report for Turlock Irrigation District (TID) shows a net income
of $169 million dollars, generating and selling electrical power and selling some
water to municipalities
TID serves some 5,800 growers on about 150,000 acres of farmland
Much of the rest were federally or state financed, at low or no interest on the
loans
Originally, the farmers on government-financed irrigation projects were
expected to repay the entire cost of the infrastructure, but later it was seen that
in many cases it could not be realistically expected, especially in the upper
basin of the Colorado river
Later, the government decided that most of the dams had a greater scope in
providing recreation facilities, flood control, municipal water supplies, etc.
The USBR and the federal government have a tight grip on the operation of
many irrigation projects today, even those that have been paid off by the water
users
The rate of water resource development in the western USA has decreased to a
small fraction of what it was in the first half of the 20th century
Partly this is due to the ever higher development costs and environmental
concerns, but also due to the development of a significant portion of the
available water resources and the completed exploitation of the most favorable
sites for dams, irrigation systems, and other infrastructure
Environmental concerns were much less prominent in the first half of the 20th
century than they are now
There has been talk for years about bringing water from the Northwest and or
Canada down to the southwestern USA, but this would be a huge political
controversy and very, very expensive
Reisner, M. 1986. Cadillac Desert. Penguin Books, New York, N.Y.582 pp.
I. Introduction
Various manuals and texts are available on the selection and design of the
different types of agricultural irrigation methods (see references below)
Such sources provide considerable detail but cannot substitute entirely for
experience
There are many factors to consider when deciding upon an irrigation method
Initial development costs and annual operational costs are usually among the
most important economic factors in the selection of an irrigation method
The farmer or land owner should have knowledge of the conditions of soils,
topography, size and shapes of fields, cropping systems, and labor availability
Social considerations and traditional practices have also generally had a large
impact on the appropriateness and feasibility of a particular method
Sometimes, the selection is made by looking over the
fence to see what the neighbor is doing
In some places, a dependable supply of electricity is
not available pressurized irrigation may not be
feasible unless another source of energy is available at
a reasonable cost (or sufficient head is available due to
elevation difference)
Remember that there are exceptions to most rules in
irrigation method selection; the ingenuity of farmers
and irrigation engineers around the world has allowed for many notable
adaptations of irrigation methods under difficult conditions
1. Compatibility
3. Topography
Field shape
Field slopes
Uniformity of field slopes
Depth to the water table
Soil texture
Soil structure
Soil depth & layering
Ability to do land leveling
Soil salinity, acidity & sodicity
Soil spatial distribution
Water holding capacity
Water infiltration rate
Surface sealing characteristics
Tendency to form a crust
Tendency for cracking
Surface or groundwater
Reliability and flow rate of supply
Frequency and flexibility of delivery
Location and elevation of the water
supply
Salinity, and types of salts present
Level of pollution, contamination
Requirements for filtration
Specific toxicities
7. External Influences
Political considerations
Vendor propaganda
Local customs & tradition
Divisibility
This has to do with the flexibility of an irrigation method to fit specific field
shapes (geometries)
Maintain by
A factor to take into account the complexity of maintenance of the
irrigation method, whether by the farmer, trained technician, or other
Crop Risk
The crop risk factor is related to the ability to keep the system running,
especially during the peak-use period, and the equipment requirements
Management and O&M Skill
This takes into account the required level of skill of the irrigator and/or
manager in operating and maintaining the system
Management and O&M Effort
This considers the amount of time and effort that the irrigator and/or
manager must dedicate to operate and maintain the system
Ruggedness
A factor to account for the durability of the system, including equipment
and proneness to breakdowns (when irrigation is suspended)
One practical factor in irrigation system selection has often been the perceived
level of technology
Surface, or gravity, irrigation systems typically
have the least amount of hardware (pipes,
valves, pumps, filters, etc.) and have in some
cases been labeled as obsolete or inherently
inefficient
Because surface systems have less hardware,
there is less to sell to farmers
Many vendors of pressurized (sprinkle and
trickle) systems in various regions of the world
have produced brochures and convinced
farmers that they are not modern if they continue
to use surface irrigation methods
Surface irrigation systems can be and often are
more efficient than pressurized systems, and they are the most common types
of on-farm irrigation system even in many of the most developed countries
Burt, C.M., A.J. Clemmens, R. Bliesner, J.L. Merriam, and L. Hardy. 2000. Selection of irrigation
methods for agriculture. ASCE, On-farm irrigation committee of the Environmental and Water
Resources Institute (ISBN 0-7844-0462-3), 129 pp.
Keller, J., and R.D. Bliesner. 1990. Sprinkle and trickle irrigation. Chapman and Hall. (ISBN 0-412-
07951-1), pp. 605-625.
Walker, W.R., and G.V. Skogerboe. 1987. Surface irrigation theory and practice. Prentice-Hall, Inc.
(ISBN 0-13-877929-5), pp 3-6.
Doneen,.D., and D.W.Westcot. 1984. Irrigation Practice and Water Management. FAO, Irrigation and
Drainage Paper 1 (Rev.), Rome, 63 pp.
evaporation
precipitation
ET ET ET ET
precipitation
evaporation
ET
off
evaporation ru n a te r infiltration
runoff
d so il w evaporation
s to r e water
t a b le
grou
ndw
r a
wate flow ter lake
sea nd
grou flow
Note that various minor additions could be made to the above figure: what
might some of those be?
The concept of a complete hydrologic, volumetric balance is sometimes
referred to as a water budget
This is a simple concept, but many people become confused with the details
and draw incorrect conclusions about the hydrologic cycle
At small scales, it can seem that water is lost, but if you step back and look at
the bigger picture, the water is rediscovered
However, fresh water can be essentially lost in some ways
Naturally
Artificially
Reverse osmosis
Distillation
Water treatment plants (which may include some of the above)
Others
The data for this example are taken from Brickson et al. (1994)
Values in the figure are in millions of acre-feet of water per year
Of course, most of the flows are highly variable from year to year, due to
hydrological fluctuations, so the data shown are for what might be a typical year
For perspective, it is estimated that a family of five uses about one acre-ft of
water per year
Note that the inflows do not necessarily have to match outflows, because there
could be a change in the volume of water storage in the state
However, over a years time, the difference in inflow and outflow volumes over
such a large area should be very tiny
Note also that most of the inflow is from precipitation (rain and snow) within the
state, while most of the outflow is from evapotranspiration, which is not only
from agriculture
Finally, it is seen that the groundwater pumping is on the same order of
magnitude as the groundwater recharge, but the spatial distribution of the two
flows is not uniform; thus, there is overdraft of groundwater in some areas
r
on
.6 te
g
16 dwa
e o
Precipitation O r 1. 4 rad .8
o
un
193 C ol 4
o
Gr
re
23 ltu
u
CA
ric
Evap
Ag
otran
L
sp
121 iration
IF
O
RN
IA
n
Ocea
30.2
l
e nta
m
v iron .5 a da
En 21 N ev
t o
off .5
R un 1
Grou
ndwa M&I
ter re 4
14.6 charge
Millions of acre-feet
Inflows Outflows
Precipitation 193
Groundwater 16.6
Colorado river 4.8
Runoff from Oregon 1.4
Evapotranspiration 121
Runoff to Ocean 30.2
Agriculture 23
Environmental flows 21.5
Groundwater recharge 14.6
M&I consumption 4
Runoff to Nevada 1.5
Totals 215.8 215.8
Precipitation
Groundwater
Colorado river
Runoff from Oregon
Evapotranspiration
Runoff to Ocean
Agriculture
Environmental flows
Groundwater recharge
M&I consumption
Runoff to Nevada
The following figure shows inflows and outflows from a canal system
Over long time periods (weeks or months), the sum of the inflows will be equal
to the sum of the outflows, if all are taken into account
There is a maximum possible change in storage, but over time the volumes
from inflows and outflows accumulate indefinitely
Over short intervals (hours or days), the sum of the inflows may not equal the
sum of the outflows: this is because of changes in volumetric water storage
within the canals
seepage precipitation
inflow inflow at
headworks
runoff
inflow
unauthorized
l
na delivery
Ca
in
Ma
Latera
l Ca
seepage
outflow
na
overflow
l
authorized
delivery
rage
authorized
D is
evaporation delivery
n e sto
bu
tr i
tar
y
i n-li
spill
transpiration
(weeds, trees, ...)
spill
Now consider a field water balance (or water budget) in the figure below
Again, the basic components of the balance are: (1) inflows; (2) outflows; and
(3) changes in water storage
Note that runoff can come onto the field from uphill, and can also leave the field;
the net runoff over a certain period of time can be positive, negative or zero
Some agricultural fields depend entirely on runoff inflow from upstream
irrigators because rainfall is negligible and there is no other source of irrigation
water
ET
evaporation
transpiration
precipitation
irrigation
evaporation
runoff runoff
deep
percolation
water table
capillarity
The net vertical flow at the bottom of the crop root zone can be either positive or
negative
With subirrigation, the total water table contribution to the root zone (capillarity)
must exceed the deep percolation in magnitude
Given:
Suppose inflow and outflow measurements have been taken over a 48-hour
period on a 100 m2 field in an agricultural research station
Measurements with instruments indicate a net increase of 18 mm in soil water
storage over the 48 hours and the 100 m2
The crop ET was calculated to be a total of 14 mm, based on measured
weather data and precisely calibrated equations
The water application from irrigation, over the same period, was measured at
10,245 liters
There was no precipitation
The surface runoff was measured with a calibrated flume, giving a value of
1,350 liters over the 48 hours
No runoff came onto the field; it was all outgoing
The water table is located at a depth of 6 m
Question:
What is the estimated net deep percolation amount, in liters, for that particular
period of time?
Solution:
There was no rain and no incoming runoff, and the water table is deeper than
the crop root zone, so the total inflow to the field was from irrigation: 10,245
liters
Assuming the crop covered the entire 100 m2, all evaporation was taken into
account by the ET calculations
The volume of ET was 100 m2 x 0.014 m = 1.4 m3, or 1,400 liters
The volume of ET plus runoff was 1,400 + 1,350 = 2,750 liters
The increase in soil water storage was 100 m2 x 0.018 m = 1.8 m3, or 1,800
liters
Thus, the water balance is given as:
More than half of the applied water was lost to the field in the form of deep
percolation during the 48 hours
What might the consequences of this be?
A watt can also be defined as: current of one ampere across a potential
difference of one volt
The annual customer service charge, plus the monthly charge (prorated) and
the monthly power charge (see below), determine the minimum that a customer
must pay, even if he or she doesnt actually use any electricity
The monthly customer service charge is $14.75
The power charge is $8.10 per kW per month (double the year 2000 rate)
The voltage discount is $0.39 per kW
The energy charge is $0.067231 per kWh for On-Peak hours, and $0.020241
per kWh for Off-Peak hours
Obviously, if a farmer elects to go with the Time-of-Day program, he had
better minimize pumping during On-Peak hours; otherwise he will pay a
relatively high rate
Some farmers cannot choose the Time-of-Day program because they pump
from an irrigation district canal and they are unable to take water anytime they
Time Periods
Power Factors
These are mostly used for customers with large power requirements (e.g.
factories, USU, municipalities, and others)
Most agricultural users wont qualify for power factors in their billings
UPL determines whether or not it is going to apply power factors in billings;
the basic issue is their cost of supplying the electrical transformers and
equipment to provide a potential for a certain level of consumption. UPL
doesnt want to invest a lot in providing a certain power capacity at a point of
delivery if the customer isnt really going to need that capacity.
Thus, when UPL thinks it is appropriate, they will use power factors and will
apply what they call a 90% lagging rule to billings: if the actual power
consumption in any given month is less than 90% of the agreed upon (or
requested) capacity, UPL will bill based on of 1% for every 1% below 90%
consumption
This means that the customer will be billed (albeit a lower amount) for
electricity that was not even consumed, if power factors are used in the billing
and if the actual power consumption was less than 90% of a certain potential
value
Post-Season Rates
Most deep well (groundwater) pumps for irrigation in the Western USA are
operated by electric motors
But some remote agricultural locations may not have electrical power readily
available and diesel engines become an alternative to electric motors
Also, in the past two or three years the electricity rates have increased
dramatically, prompting many farmers to convert to diesel engines on surface
pumps and deep wells
The conversion usually requires the installation of a gear head to
accommodate the horizontal shaft from an engine (whereas the pump shaft is
vertical)
Many diesel engines used for pumping irrigation water in the USA are air-cooled
to reduce both risk of failure and maintenance costs
In the summer of 2000, Dr. Robert Hill and Bruce Godfrey (with help from Bill
Bullen & Sheridan Nicholas) estimated that the O&M costs per 1000 hrs of
operation for diesel engines is $611, whereby for electric motors it is only $40
(assume 50 HP capacity)
They estimated total annual costs as:
Electric motors have many fewer moving parts than diesel engines, they last
longer, and require much less maintenance
However, the cost to extend Utah Power & Light lines is $20,000 per mile
If extensions are necessary, an effective annual cost increment of $2,200 per
mile should be added to the electric power costs
In these cases, it may be more attractive to use a diesel engine, at least in the
short run
Other possible fuels for engines are gasoline, propane, and natural gas
Currently in Utah, electricity costs less per volume of water pumped than any of
the fuel-engine alternatives, but this situation could change as electrical rates
continue to rise
Electric motors come in many different sizes and of many different types, and
they have many advantages over other motors and engines:
a. relatively efficient
b. economical
c. easy to automate
d. compact in size
e. run quietly
f. do not pollute significantly
g. require low maintenance
h. do not tend to vibrate
i. operate under a wide range of ambient temperatures
1. split phase
2. permanent-split capacitor
3. shaded pole
4. capacitor-start/induction-run
5. capacitor-start/capacitor run
6. series (universal)
7. repulsion-start/induction run
8. three-phase, and
9. others
Some motors are good at starting under light loads, and others start better
under heavy loads
Single phase power is 110V in the USA; one hot wire and one common wire
Single phase power is what we have in homes in the USA
Two-phase power is 220V in the USA (but also called single phase in USA);
two hot wires and one common wire
Three-phase power has 3 hot wires and 1 common
wire
Three-phase is more common for larger motors, such
as agricultural pumps
Electric motors only pull the amount of electricity required, meaning that they
can be oversized and the operational consequences are small (relative to
internal combustion engines)
But if the motor and required load are not well matched, the motor may operate
at a lower efficiency
Electric motor efficiency is from about 90% to 98%, where newer and larger
motors tend to be in the higher end of range
Electric motors are rated (e.g. 100 HP, 200 HP, and so forth) by their output,
not their input
http://www.utahpower.net/
2. Modulus of Elasticity
A measure of compressibility
A function of pressure and temperature
High values of modulus of elasticity are characteristic of relatively
incompressible liquids (large change in pressure results in relatively small
change in volume)
dP P
Ev = V V (1)
dV V
3. Viscosity
du
=
dy (2)
The above equation says that the shear force is equal to the dynamic viscosity
multiplied by the relative velocity between the two plates
=
(3)
There are two distinct classes of flow for full pipes and open channels:
1. laminar flow
2. turbulent flow
Q = V1A1 = V2 A 2 (4)
This equation represents the different components of energy in liquid flow, per unit
mass of liquid. It is valid for the following conditions:
steady flow
incompressible flow
ideal flow (no friction loss)
P V2
+z+ = constant
2g (5)
A 3
= 0
V dA
3A
Vm (6)
V. Reynolds Number
FI L2 V 2 LV LV
NR = = = =
Fv LV (7)
where Fl are inertial forces (N); Fv are resistance forces (N); is the density of
water ( 1,000 kg/m3, or 62.4 lbs/ft3); is dynamic viscosity (cP); is
kinematic viscosity (m2/s); and V is mean flow velocity (m/s)
1. relative roughness, /D
2. Reynolds number, NR
The Moody diagram shows experimental results for the relationship between
f, NR, and /D
There are regression equations that fit all of the curves in the Moody
diagram very well (a set of equations)
2. Hazen-Williams
JL
hf =
100 (9)
1.852
Q
J = 1050 D4.87
C (10)
note also that the C factor is also a function of NR, even though this
is often ignored in practice
Darcy-Weisbach is often preferred over Hazen-Williams due to its
higher accuracy
V2
hf = k
2g (11)
L kD
k=f e or, Le = (12)
D f
1. Specific Energy
Specific energy is based on Bernoullis equation and the values are used
comparatively among two or more locations in a channel
For example, the specific energy equation is used in the calibration of broad-
crested weirs, upstream and at the weir crest
The units are in terms of head (m or ft, usually), and give an estimate of the
relative energy at a channel location
V2
E =h+
2g (13)
where h is depth of water (m); V is the average flow velocity (m/s); and
P
+z
(14)
P/
2. Uniform Flow
Uniform flow occurs when the slope of the water surface in the direction of flow
is constant and equal to the longitudinal bed slope of the channel
Manning Equation
1
Q= AR2 / 3 So
n (15)
where Q is flow rate (m3/s); A is flow cross sectional area (m2); R is hydraulic
radius, A/Wp (m); Wp is the wetted perimeter (m); So is longitudinal bed slope
(m/m); and n is a roughness factor
For Q in cfs and all linear dimensions in feet, the right side of the equation must
be multiplied by approximately 1.49, where
1
1.49
(0.3048 m / ft)1/ 3 (16)
The equation can be solved for Q or for any other variable (for example, depth
of flow, h)
For symmetrical trapezoidal channel cross sections:
A = h(b + mh)
Wp = b + 2h m2 + 1
(17)
where b is the base width (m); and m is the inverse side slope (m = 0 for vertical
walls, i.e. a rectangular cross section)
Q = CA RSo
(18)
A graph exists (see below) for specifying the C value as a function of Reynolds
number and relative wall roughness, much as with the Darcy-Weisbach
equation and Moody diagram; this makes the Chezy equation technically
preferable to the Manning equation
The Manning and Chezy equations have been used extensively in the static
(steady-state, full capacity) design of irrigation canals all over the world
The Chezy C is actually a kind of smoothness coefficient (because larger
values mean greater discharge)
0.01
50 0.02
0.03
Chezy C (V in m/s, R in m)
0.04
40 0.06
0.08
0.10
0.15
0.20
30 0.25
0.30
0.35
20
Laminar Flow
10
1.0E+02 1.0E+03 1.0E+04 1.0E+05 1.0E+06 1.0E+07
Reynold's Number (4RV/v)
3. Flow Regime
The specific energy equation can be differentiated with respect to flow depth, h,
(holding Q constant) and set equal to zero to define the minimum specific
energy conditions for a given flow rate
This produces the Froude number equation:
Q2 T
Fr2 =
gA 3 (19)
where T is the top width of flow (m); and Fr is the Froude number. For
symmetrical trapezoidal sections, T = b + 2mh.
4. Momentum Function
kg m kg m
s = (20)
s2 s
in words, force x time or mass x velocity, where kg/s is the mass flux
This function is used most often for the analysis of hydraulic jumps in open
channels, but also for other related purposes
Q2
M= + Ah (21)
gA
where M is the momentum function value (m3); Q is the flow rate (m3/s); A is
the cross-sectional flow area (m2); and h-bar is the vertical distance from the
free water surface down to the centroid of the cross-sectional area (m).
Upstream Downstream
subcritical
critical depth
supercritical depth
depth
conjugate depths
You can determine the value of h-bar by simple geometric equations, and the
formulas are given for many different section shapes in open-channel hydraulics
books
After determining conjugate depths, you can apply the specific energy equation
upstream and downstream to estimate the hydraulic energy loss across the
jump
Good quality, fresh water, is becoming more and more scarce as people exploit
water resources more aggressively, and as the world population increases
This increases the importance of water measurement
It is unlikely that the regional and global situations on water availability and
water quality will improve in the foreseeable future
Strictly speaking, all open-channel and most pipe flow measurement techniques
cause head loss
However, some methods incur negligible losses (e.g. ultrasonic)
It is usually desirable to have only a small head loss because this loss typically
translates into an increased upstream flow depth in subcritical open-channel
flow
In open-channel flow measurement, devices can operate under free flow and
submerged flow regimes
In modular flow we are concerned with the upstream head because critical flow
occurs in the vicinity of the flow measurement device. As long as this is true,
changes in downstream depth will not affect discharge at that location. In non-
modular flow we are concerned with a head differential across the flow
measurement device.
In this class, the terms flow rate and discharge will be used interchangeably.
Perhaps the most accurate method for measuring flow rate is by timing the
filling of a container of known volume
However, this is often not practical for large flow rates
Typical flow measurement accuracies are from 2% to 20% of the true
discharge, but this range can be much higher
Measurements of head, velocity, and area are subject to errors for a variety of
reasons:
1. Approach Conditions
3. Equipment Problems
4. Measurement Location
Local Measurements
Stream Gauging Stations (need steady flow)
5. Human Errors
The following methods are considered to be special methods, because they are
mostly simply and approximate, and because they are not usually the preferred
methods for flow measurement in open channels
Preferred methods are through the use of calibrated structures (weirs, flumes,
orifices, and others), and current metering
2. Measurement by Floats
Select a location in which the channel is fairly straight, not much change in
cross-section, smooth water surface, and no abrupt changes in bed
elevation or longitudinal slope
Note that wind can affect the velocity of the float, changing the relationship
between surface velocity and average flow velocity
Care should be taken to obtain measurements with the float moving near the
center of the surface width of flow, not bumping into the channel sides, and
not sinking
The float speed will be higher than the average flow velocity in the channel
You can estimate the average velocity in the channel by reducing the float
speed by some fraction
The following table is from the U.S. Bureau of Reclamation, but it is an
oversimplification of reality by making the coefficients a function of average
water depth alone, and often gives flow rate errors of 20% or more
It gives coefficients to multiply by the measured float velocity, as a function
of average depth, to obtain the approximate average flow velocity in the
channel
What happens to the above coefficient values when the average water
depth is below 1 ft or 0.3 m?
3. Dye Method
5. Radioisotope Method
V = 2gh
This method is best applied for higher flow velocities because it is difficult to
read the head differential at low velocities, in which large errors in the
estimation of velocity can result
It is also best for flows with a smooth, stable water surface
Following is a list of flow measurement methods for open channels that require
special calibrations and designs, or more expensive equipment than the simple
methods
These methods also generally require more expertise and technical knowledge
to apply successfully
1. weirs
2. flumes
3. ultrasonic
4. laser-doppler
5. underflow gates
6. current metering
I. Introduction
1. minerals;
2. organic material;
3. water; and,
4. air
Here, minerals include all solids except for organic (plant & animal) matter
Typically, about half of the soil volume is pore space, which is interstitial space
occupied by water and air
There are usually many different chemical elements present in agricultural soils, as
soil formation is primarily a process of mineral weathering with the addition of
organic constituents
Most agricultural soils have only a small percentage of organic material, and the
predominately mineral solids usually have a fairly consistent particle density of
about 2.65 gm/cm3
The various soil textural classes are based upon a system developed several
decades ago in which the relative percentages (by weight) of sand, silt and clay
are used to determine the class
The differences between sand, silt, and clay cannot be accounted for only by the
differences in size ranges, and the chemical composition of clays is usually quite
different than that of sands and silts
The figure below shows the soil texture triangle diagram that has been generally
accepted for use with agricultural soils, encompassing twelve different classes,
each with a descriptive name
90 10
80 20
70 Clay 30
y
60 40
Cla
Pe
r ce
nt
50 Silty 50
rce
nt
Sil
Pe
Sandy clay
t
40 clay 60
Silty clay
Clay loam loam
30 Sandy clay 70
loam
20 80
Loam
L Sandy loam Silt loam
10 oam 90
y sa Silt
Sand nd
100
100 90 80 70 60 50 40 30 20 10
Percent Sand
Some soil classifications refer to four components: mineral matter, organic matter,
air and water
Organic material is present in varying amounts in all agricultural soils, and those
with relatively large amounts of organic matter are sometimes called peats or
mucks
Soils with 20 to 30% or more of organic matter by weight have physical properties
dominated by the organic fraction, and not by the mineral fraction
The presence of organic material can have
a profound influence on the structure and
physical characteristics of a soil, as well as
the porosity and workability of the soil
Some clay soils can be very difficult to till
except within a very narrow range of water
contents - in one instance the soil may be
too dry and hard, but with too much water it
is slippery and sticky
Organic material almost always improves
the physical characteristics of a soil, and
increases the porosity
Agricultural soils are often compared and distinguished according to texture and
structure, which are two essentially independent classifications
Standard textural soil classifications include three inorganic constituents: clay, silt
and sand
The relative percentages of these three constituents determines the textural
classification, such as sandy loam, clay, and silty clay loam
The structure of a soil refers to the looseness or compactness (amount of pore
space), and the alignment of small colloidal particles
Both texture and structure are directly related to the permeability of a soil, or the
rate with which water can infiltrate downward into the soil from the surface
There are several methods for evaluating soils or classifying lands with respect to
their agricultural potential
A. profile: Soil profiles are divided into nine groups based on soil or
profile development, geologic position, and source of soil forming
material. Within each profile group the percentage rating varies
principally with soil depth and the degree of profile development.
B. surface texture: Fine sandy loams, loams, and silt loams are
rated 100% except when they are gravelly or stony, in which case the
maximum is 70 to 80%. Coarse textured soils and fine textured or
heavy textured soils are given lower ratings.
C. slope: Soils that are nearly level (0 to 2%) are rated 100%. Very
steep slopes (45% and higher) are rated 5 to 30%.
Factor Rating
A. Recent alluvial soil with unconsolidated profile 100%
B. Highly calcarious silty clay 80%
C. Undulating (3 to 8%) 90%
X. Fair nutrient level 95%
Composite Storie Index (100% x 0.80 x 0.90 x 0.95) 68%
Class Description
1 Lands most suitable for development
2 Lands less desirable because of soil, topographic, or drainage characteristics
3 Lands which are considered least suitable and which have the lowest potential
repayment capacity
4 Lands with excessive deficiencies but considered capable of repaying costs upon
development
5 Lands not suitable for irrigation under present conditions but worth segregating for
further study
6 Lands considered to be permanently non-arable
In many land classification studies only classes 1, 2, 3, and 6 have been mapped
Definite specifications are established for each major project area prior to land
classification
Generally, geology and the type of parent material have more influence on the
characteristics of younger soils
Soil physical and chemical properties are determined largely by the nature of the
parent material, the soil age, and the amount of leaching and decomposition that
has taken place
Many of the better soils for irrigation development are fairly young alluvial deposits
from calcareous marine parent material or from basic igneous rocks
Topography and position frequently relate to soil texture
Where the flow of surface water slows, sands and coarse materials are deposited
first, while clays are usually deposited in locations where water movement has
become very slow or even stagnant
The medium-textured soils (e.g. loams and silt loams) are rated highest for
agricultural development
The quality of a soil for irrigation is influenced by acidity or alkalinity, salinity, base
exchange capacity, base saturation, nutrient (fertility) level, and by other conditions
or elements
The optimum range of soil pH for irrigated crops varies somewhat. A pH of 6.5 is
optimum for many crops, although the full range of suitable values varies by crop
The pH should not be less than 5.0 or more than 8.5 for the most agricultural crops
Alkaline soils are those with a pH of greater than 7
More rain falls in the eastern USA than in the west, and agricultural soils tend to be
much more acidic in the east
Lime is often applied to acidic soils to raise the pH, but this is almost never done
on soils in the western USA
VII. Definition of pH
The acidity or alkalinity of a soil water solution is commonly quantified by the base
10 logarithm of the reciprocal of the hydrogen ion concentration in moles per liter
The resulting value is called the pH of the solution
In equation form: pH = log (1/H+)
The pH scale runs from 0 to 14, where 0 is for a strong acid solution and 14
corresponds to a strong base
A pH of 7 is neutral
Several equations are given below to define the common terms related to soil
water content
These are important in irrigation system design, evaluation and operation
Porosity, , is simply the percent pore space (air and water) related to the total
soil volume
Porosity usually ranges from 0.3 to 0.6, or 30% to 60%
Porosity tends to be higher for fine-textured (clay) soils, but it is also a function
of the soil structure
Vair + Vliquid b
= = 1 (1)
Vtotal s
where the volume terms are defined in the following figure, with Vliquid being the
volume of water in the soil; b is the bulk density (defined below); and s is the soil
particle density
Soil particle density is approximately 2.65 gm/cm3 for most agricultural soils
The void ratio is defined as:
Vair + Vliquid
e= (2)
Vsolid
Saturation occurs when the pore space of a soil is completely filled with water
Saturation means essentially no air, which is a problem for most crops if the
condition persists for more than a few hours
But, in reality, even if the soil is saturated the soil water will contain some air
Saturation, S, can be defined as a function of volumetric water content and
porosity:
Vliquid v
S= = (3)
Vair + Vliquid
The dry mass fraction water content is the ratio of the mass of water in a given
soil volume to the mass of solid particles:
w Vliquid
m = (5)
s Vsolid
where m is the dry mass fraction water content; and w is the density of water (1
g/cm3; 1 kg/liter; or 1,000 kg/m3)
Dry mass fraction is sometimes used, but for our purposes we are more
interested in v
The density of water depends on the chemicals dissolved in the water, but can
usually be taken to be slightly higher than 1.0 g/cm3
Bulk density can be defined as the ratio of the mass of solid soil particles (dry
mass) to the bulk soil volume:
s Vsolid
b = = s (1 ) (6)
Vtotal
The available water (AW) is difference in water content at field capacity and
wilting point (or permanent wilting point), multiplied by the crop root depth, Rz
Ignoring the water between saturation and field capacity can be considered a
safety factor in design calculations for irrigation system capacity
In equation form, the available water is:
AW = ( FC WP ) R z (7)
The following figure shows two sample soil water content curves: one for the soil
water distribution in the root zone immediately following an irrigation, in which the
soil water content at the surface is at or near saturation, and the other for the
redistributed soil water after a few days
The following figure shows a generalized curve of the soil water content above a
water table, some 1 to 3 days after an irrigation at the soil surface:
t = z + p + s + m (8)
s 0.36EC (9)
The solute potential increases as the soil water salinity increases, making it
more difficult for plant roots to extract water from the soil
Solute potential can be directly measured by:
1. Freezing-point depression
2. Vapor-pressure osmometers
3. Thermocouple psychrometers
Matric potential is related to the adsorptive forces of the soil, and can be
measured using a tensiometer
Matric potential and volumetric water content are sometimes assumed to be
uniquely related for any given soil, but because of hysteresis this is not quite
true
Matric potential is nearly zero for saturated soil conditions
IV. Hysteresis
I. Introduction
Several methods can be used to determine soil water content, and new
methods are being developed
It is important to have measurements of soil water
content to determine when to schedule irrigations
V. Neutron Probe
The presence of other dielectric molecules in the soil will bias the reading
Therefore, often TDR must be calibrated to a specific soil
Also, the presence of salinity in the soil will change the dielectric constant for
the soil-water solution
Therefore, a correction for salinity must be made
The dielectic constant is a weak function of temperature, so that a temperature
correction may be needed for high accuracy
There is generally a limit of about 0.3 m for the lengths of wave guides
This is because the signal can become too weak on long guides
In addition, as the electromagnetic signal passes through changes in soil
texture or large changes in soil water content, these changes can cause false
reflections to occur, and it becomes impossible to determine the true reflection
Therefore, TDR is usually used only for the top 30 cm of soil
For deeper depths, you may need to dig a trench and then insert the TDR wave
guides horizontally
Care must be taken to not disrupt the soil or roots where these may impact
subsequent infiltration or water extraction
Some companies are making a long probe that is comprised of a series of TDR
wave guides
This probe is installed permanently into the soil profile, the depth of the effective
root zone, to assist in evapotranspiration (ET) studies
A more literal title for TDR that translates more readily into other languages is:
"Time-based measurement of electromagnetic reflection"
Other sensors use techniques that are similar to TDR, however, instead of
relying on a reflection of a signal, they merely measure the time for a signal to
travel a certain distance along guides
These guides are in the form of wires, and should be about 10-m long to make
a loop back to the source
The benefit of these sensors is that they measure a long integrated path of soil
This is valuable where the crop root extraction or the infiltration of irrigation
water is highly variable
It may be possible to use these sensors to measure the average water content
of a full root zone
VIII. Tensiometers
The porous ceramic cup regulates the flow of water into and out of the
tensiometer, from the surrounding soil
After some time it will reach equilibrium with the tension in the soil
The reservoir tube holds a supply of water, which the tensiometer needs to
operate
The water should be de-aerated, if
possible (boil it first, then let it cool)
A hand-operated vacuum pump is
used to evacuate air from the
tensiometer after it is filled with
water
An end cap seals the unit and
allows negative pressures to be
maintained inside the tensiometer
The dial gauge provides a reading
of the soil water tension
Tensiometers function within a
certain range of negative
pressures, in the soil vadose zone: about zero to 100 cb (0 to 14.7 psi) under
ideal conditions, but only up to about 75 cb tension under many practical
conditions
http://www.earthsystemssolutions.com/assets/TensiometerUsersGuide.html
http://www.sowacs.com/feature/deltat/swp-swt.html
http://inelext1.inel.gov/science/prestige.nsf/ineel/Tensiometer
http://home.t-online.de/home/unidata/5242eng.htm
http://weather.nmsu.edu/Teaching_Material/soil698/tensio.html
http://www.tensio.de/english/steck.html
http://www.irrometer.com
I. Infiltration Rates
For example, the cerrado soils in central Brazil have relatively high infiltration rates
even though their texture is predominately clay
The following table indicates the approximate bulk density (dry weight density) and
porosity of three different soil types
Soil infiltration rates can be measured using cylinder ring infiltrometers, ponds,
inflow-outflow in furrows, blocked (ponded) furrows, and through the use of special
equipment such as recycling furrow infiltrometers
A cylinder ring infiltrometer, as shown below, is usually made of steel and can be
with single rings or concentric double rings
In sprinkler irrigation the impact of the water drops on the soil surface can cause a
partial sealing and reduced infiltration rate
This is especially true with sprinklers that produce large drops on heavy textured
(clay) soils
The velocity of the drops as they arrive at the ground surface is also a factor in the
partial sealing phenomena, but most drops from sprinklers reach their terminal
velocity before hitting the ground (Keller and Bliesner 1990)
The amount of water available in the soil that can be used for crop growth and
development depends upon several parameters
These variables include the crop effective root density and depth (how well the
root system explores the soil), as well as the ability of the soil to store available
water
The principal soil characteristics used to estimate the
amount of potentially available water storage include
the field capacity (FC), the permanent wilting point
(PWP), and the available water (AW)
Field capacity may be defined as the soil water
content just after rapid drainage due to gravity,
following an irrigation in which the soil was
temporarily saturated
Field capacity will typically be reached one to three
days after an irrigation in which the soil in the root
zone is completely refilled
The field capacity can be measured directly as a percentage by volume, which can
be expressed in terms of mm/m of soil depth
Field capacity can also be calculated from the bulk density and the fraction of the
weight of water in a soil sample
The bulk density may change with irrigation, passage of farm equipment, or tillage
practices
For agricultural soils, it can vary from 1.1 gm/cm3 for highly structured clays to 1.8
gm/cm3 for fairly compact coarse sand
For estimates in connection with irrigation scheduling, an average value of 1.4 or
1.5 is frequently assumed for medium soil textures
The field capacity (FC) is the amount or percentage of water by volume of soil that
can be held against drainage by gravity
This usually occurs between 1/10 atmosphere of tension for coarse soils and 1/3
atmosphere for heavy soils
A practical procedure for determining FC is to select an area with no plants on it,
flood the area to saturate the soil, cover with canvas or plastic to prevent
evaporation, and take soil samples after the soil has drained to the field capacity
The drainage time required is usually one day for sandy or coarse textured soils
and may be as much as four days for fine textured or heavy textured soils
Soil texture is not a very good indication of AW the variation in water holding
characteristics within a given texture is too great
In one study, a linear regression was made for values of volumetric water content
from 8 to 41% at FC
The equation is:
The coefficient of determination (r2) was 0.98 for the data used to calibrate the
equation
Allen (1994) developed an equation from the data given by Jensen, et al. (1990):
AW = 1.55 (FC )
0.66
(2)
When plants are using water from the upper 50% of AW there is little stress
In the lower range, plants experience increasing water stress, and when the PWP
is reached, water use by the plant approaches zero
For irrigation scheduling, it is assumed that soil water should be replaced when
between 25 and 75% of available water is depleted
The actual recommended depletion fraction or percentage depends upon the crop
potential evapotranspiration and the crop type
When cumulative soil water infiltration (or intake) is measured, the data typically
take the form shown below, for vertical (one-dimensional) intake
Note that more points are taken at small times, where the curvature is greatest
0.015
0.01
0.005
0
0 10 20 30 40 50
Intake Opportunity Time
The slope of the curve tends to reach a constant value after some time
The slope of the above curve can be plotted with respect to time, to obtain another
typical curve shape for infiltration in agricultural soils
0.0030
0.0025
0.0020
0.0015
0.0010
0.0005
0.0000
0 10 20 30 40 50
Intake Opportunity Time
The infiltration rate is very high during the first instants of water application to a soil
The nearly constant infiltration rate after some time is called the basic intake rate
If the soil is very wet already, the measurements may start at the basic intake rate,
which remains constant
Thus, to develop the full curves shown above, it is best to start with a fairly dry soil
Various equations have been developed for predicting soil water infiltration
properties
The most commonly-used equations for irrigation design, evaluation and
management purposes are presented below
1. Kostiakov Equation
Z = ka (3)
Intake opportunity time means the elapsed time since the water first arrived at the
surface of the soil
For most surface irrigation methods, the value of varies spatially due to the time it
takes for water to travel over the surface of the field to different locations
That is, intake opportunity time is something like elapsed irrigation time minus
advance time
The Kostiakov-Lewis equation has an extra term added, and better approximates
field measurements of infiltration:
Z = ka + fo (4)
I = aka 1 + fo (5)
Eqs. 4 and 5 are commonly used because of their simplicity and good ability to fit
most field data
2. Green-Ampt Equation
IMD
I = K 1 + S (6)
z
3. Richards Equation
v h
= K(v ) 1 S(z,t) (7)
t z t
Soil hydraulic conductivity usually depends on the direction of flow (e.g. vertical,
horizontal, or other direction) because of soil layering and structure
Hydraulic conductivity also depends greatly on the soil water content, being
highest for saturated conditions, and decreasing rapidly to zero for dry conditions
There are various methods for measuring K in soils, including the auger hole test
Saturated hydraulic conductivity ranges from a low of 0.25 cm/hr (clays) to a high
of 25 cm/hr (sands) in most agricultural soils [0.1 to 10 inches/hr]; for loam soils it
may be around 2 cm/hr
Permeability is a general term and hydraulic conductivity is specific to the
movement of water in soils, but they are often used synonymously
Darcys law states that the flow of water through a porous media (e.g. soil) is
proportional to the hydraulic gradient and to the hydraulic conductivity of the media
This is somewhat analogous to Ohms law in electricity
This can be expressed as:
Q = KiA (8)
The negative sign in front of the right side of the above equation accounts for flow
direction water flows from higher to lower potential, or opposite to the potential
gradient
That is, the gradient decreases in the direction of flow
Note that for vertical flow of water in a soil, H = L, and i = 1.0
Below is a table of cylinder ring infiltration data from Walker (1989, FAO #45)
Note that the readings are most frequent at the start of the test, and that this one
goes for about 24 hours (not all tests run so long)
50
40
30
,
20
10
0
0 200 400 600 800 1000 1200 1400 1600
Intake Opportunity Time (min)
In the last few data points, the fo term dominates the Z value
Solving for the basic intake rate, fo, we can perform linear regressions on the
data points taken toward the end of the test, where the rate is roughly linear
Following are the regression results for different numbers of data points,
starting from the last point and going back in time
Using the last six values gives a high correlation and what is likely to be a
representative value of the basic intake rate, and fo 0.031 mm/min
With seven or more values, the correlation steadily declines, as we move into
the curved region of the data
0.040 1.000
Basic Intake Rate (mm/min)
0.035 0.998
0.030 0.996
Correlation (R2)
0.025 0.994
fo (mm/min)
0.020 0.992
Correlation
0.015 0.990
0.010 0.988
0.005 0.986
0.000 0.984
0 2 4 6 8 10 12 14
Number of Points
In the first few data points, the ka term dominates the Z value
ln() ln(Z)
0.000 1.386
0.693 1.609
1.386 1.792
1.792 1.946
2.303 2.079
2.996 2.303
3.401 2.398
4.094 2.639
4.787 2.890
5.193 3.045
A linear regression on the above data gives a = 0.31 and k = e1.37 = 3.94, with a
correlation of 0.998
Thus, the calibrated equation for these data, with in minutes and Z in mm, is:
This equation predicts the measurements fairly well, but an fo value of about
0.025 will give a better fit at high values
We could have just determined fo directly from the last two data points, and in
this case the results would have been better
I. Introduction
8
Leaf Area Index (LAI)
100%
6
m LAI alue)
im u v
Opt given R s 75%
(for a
4
50%
25%
2
0
0 100 200 300 400 500 600 700
2
Incident Solar Radiation (Rs) (cal/cm /day)
For GDD40:
The same rules are applied in the case of GDD50; it is just that the cardinal
values are different
Other definitions for GDD have been proposed and used, but they are similar to
the two described above
Crop yield models have generally been developed for predicting dry matter
production in grain and forage crops and are available for a limited number of crop
types
Also, most of these models focus exclusively on the effects of water deficit in the
crop root zone, not on the many other factors
Some models, however, relate only to the effects of salinity in the root zone
Much research work continues today in this area of study
Doorenbos and Kassam (1979), in FAO Paper 33, proposed a procedure for
estimating the crop yield response to soil water availability
Reduction of crop yield below the maximum for prevailing conditions was assumed
to be proportional to the reduction in actual evapotranspiration (ETa) below the
maximum potential evapotranspiration (ETm)
ETm is the evapotranspiration under conditions where soil water availability does
not limit crop yield
n
Ta i
Yrel = 100 (4)
i=1 Tmax i
1. Establishment
2. Vegetative
3. Flowering
4. Yield Formation
5. Ripening
The above concepts have been incorporated into various crop yield or crop growth
models
Some models use estimated crop transpiration (T), others use crop ET, and some
estimate interactions of fertility and drainage on yields
That is, 40% of the total water extraction occurs in the first quarter of the root
zone depth, 30% from the second quarter, etc.
These values are generalizations from experimental measurements for various
crops
The actual extraction pattern varies according to crop type and growing
conditions
Ashcroft, G.L., D.T. Jensen, and J.L. Brown. 1992. Utah Climate. Booklet published by the Climate
Center at Utah State University, Logan, Utah. 127 pages. (http://climate.usu.edu/)
Doorenbos, J. and Kassam, A.H., 1979. Yield Response to Water, Irrigation and Drainage Paper 33, FAO,
Rome, Italy.
Hill, R.W., R.J. Hanks, and J.L. Wright. 1987. Crop yield models adapted to irrigation scheduling
programs. Irrigation systems for the 21st century, ASCE conference proceedings, Portland, OR,
July 28-30, pp. 699-706
Stewart, et al. 1977. Optimizing crop production through control of water and salinity levels in the soil. Utah
Water Lab. PRWG 151-1, Logan, Utah, p. 191.
Rn = ET + H + G (1)
where,
Rn = net solar radiation at the ground and plant surfaces
ET = latent head energy used in the evapotranspiration process
H= sensible heat transfer to the atmosphere
G= sensible heat transfer to (or from) the ground
If you measure Rn, H and G, you can solve the above equation for ET
You can measure G directly using soil heat flux plates, but these are expensive
and mostly used in research (G can be estimated based on air temperature)
Net radiation is basically the amount of incoming short wave radiation minus
outgoing long wave radiation
This involves weather conditions (e.g. cloudiness) and the determination of
surface reflectance, or albedo
Surface conditions and crop types are factors in the magnitude of the albedo, ,
which is typically between 0.23 and 0.30 (the fraction that gets reflected)
Incoming short wave radiation can be measured with a radiometer
Common units for radiation are MJ/m2/day, or langleys/day (this is power per
unit surface area; note that a J/s is a watt)
A langley is a calorie per cm2
ET is commonly given units of equivalent water depth (mm or inches) per day,
week, month or year
For purposes of irrigation planning, design, and management, computations are
usually made to determine reference evapotranspiration (ETo)
ETo is multiplied by a crop coefficient (Kc) to determine the evapotranspiration of
a particular crop at a given growth stage
Much has been written about crop water requirements and many different
equations have been developed and used for calculating ET
Many of the equations for ET are complicated and require several weather
measurements that may be unavailable or of questionable accuracy at a given
site
It is often desirable that methods selected for calculating crop water
requirements require a minimum of measured climatic variables, and be simple
and easily understood
Also, even though you may have the required climatic parameters for a more
complete ET equations, you have to make a judgment as to whether the
accuracy of the data justify the use of such and equation (sometimes the data
are complete, but incorrect)
We will look at some equations in the next lecture
V. Daily ET Magnitudes
The peak ET rate, which typically occurs during the hottest period of the
growing season, for most mature, healthy agricultural crops is from about 5 to
10 mm/day (0.20 to 0.39 inch/day)
Peak ET 5 to 10 mm/day
Most agricultural crops have peak ET rates within the above range
For immature, diseased, pest-stricken or poisoned crops, the actual peak ET
will tend to be less than it could be
More frequent irrigations will tend to increase the evaporation component of ET
Crops with low LAI values (e.g. sparsely populated) tend to have lower ET rates
in which is the latent heat of vaporization in calories per cm2; and T is the air
temperature in C
o temperature-based methods
o radiation-based methods, and
o combination methods
Crop coefficients generally include both transpiration from plant stomata and
evaporation from wet soil and foliar surfaces, with the assumption that the
availability of water is not a limiting factor to plant growth and development
Frequent (e.g. daily) irrigation usually causes higher ET rates than those
normally considered in the development of crop coefficient values because of
higher wet soil evaporation (especially with sprinkler and surface irrigation) and
possibly higher transpiration as well
Crop coefficients depend principally on crop type and stage of growth
A crop coefficient for grass based on a grass reference is often less than 1.0
because grass is usually kept shorter than a 12-cm height
K cb K c
daily Kc
daily Kc
daily Kc
daily Kc
Kcb
Crop Coefficient
Kcb
full cover
Kcb
0.0
Elapsed Time in Growing Season
The reason all the daily spikes peak at Kc = 1.0 is that for a reference crop
(usually grass or alfalfa) Kc equals unity by definition
The above Kcb curve increases as the crop matures and the LAI increases for
an annual crop
However, the curve generally takes the same shape for perennial crops
because of the significant effects of weather conditions on ET
Kc values generally range from about 0.2 to 1.2 for a grass reference crop, and
the value for many agricultural crops never exceeds that of grass, even during
the peak use period
But because of the significant difference between grass and alfalfa ET, it is
important that Kc values be qualified as grass or alfalfa based
When this distinction is not explicitly made it is best to assume that the values
are based on a grass reference
The figure below gives generalized crop coefficients and the growth stage,
which is approximately correct for weekly or longer calculation periods, and for
basal crop coefficients
(Kc)3
(Kc)1
Initial Rapid Late
Growth Growth Mid-Season Season
A B C D E
Crop Growth Stage
Very large lysimeters may have an access ladder to allow for inspection and
maintenance, but many lysimeters are too small for this kind of feature
It is often difficult to notice a lysimeter because most of it is underground, and
the vegetation on the surface hides the upper edges
Changing water content in the lysimeter is measured by weighing, by comparing
applied water with the amount of drainage, or by other suitable methods
The ET of various grasses grown in lysimeters has been used to develop and/or
calibrate numerous equations for estimating ETo
However, due to the wide variations in the ET of grasses and in the
management and design of lysimeters, there has been considerable variation in
the calibration of equations for computing ETo
The emphasis on the use of lysimeters has shifted more toward the
determination of crop coefficients than reference crop evapotranspiration
because equations (most notably, the Penman-Monteith equation) have been
shown to predict reference ET with excellent accuracy for most agricultural
locations around the world
The principal climatic factors used include incoming solar radiation, net
radiation, extraterrestrial radiation, air temperature, the temperature range,
relative humidity, vapor pressure, sunshine hours, and wind speed
Most of these methods require some degree of local calibration for general
application
There have been made of at least eight versions of the Penman combination
equation
This equation makes use of a wind run function, which is the average wind
speed over a specified period of time, multiplied by that elapsed time to obtain a
value in kilometers or miles, for example
If the time period is a day, it is called daily wind run
The required net radiation is usually estimated from global radiation (RS)
ETo = K pEpan
(1)
where Epan is measured evaporation over some specified time period (usually in
mm or inches)
In Eqs. 2 & 3:
The wind speed at 2 m above the ground can be estimated from the wind
speed, Uz, at a height z above the ground:
2 0.67hc
ln
0.123hc
U2 = Uz (4)
z 0.67hc
ln
0.123hc
There are a lot of ET equations in the technical literature, and some have many
variations (see http://www.fao.org/docrep/X0490E/x0490e00.htm)
Following is a list of some of the more well known ET equations (in alphabetical
order):
a. Blaney-Criddle, FAO-24
b. Blaney-Criddle, SCS TR-21
c. Businger-van Bavel
d. Christiansen-Hargreaves Pan Evaporation
e. FAO-24 Pan Evaporation
f. Hargreaves
g. Jensen-Haise
h. Kimberly Penman
i. Penman
j. Penman FAO-24
k. Penman-Monteith
l. Priestly-Taylor
m. Radiation Method
n. Thornthwaite
o. Turc
Hargreaves, et al. (1985) and Hargreaves and Samani (1985) proposed the use
of an equation for estimating ETo from air temperature and latitude:
in which ETo and Ra are in the same units of equivalent water evaporation
(often in mm), Ra is extraterrestrial solar radiation (see the next lecture)
900 U2 ( es ea )
0.408 (Rn G ) +
ETo = T + 273 (2)
+ (1 + 0.34U2 )
where ETo is in units of mm/day for a grass reference crop; is the slope of the
saturation vapor pressure function (kPa/C); is a psychrometric constant
(kPa/C); Rn is net solar radiation (MJ/m2/day); U2 is the wind speed (m/s) at 2.0
m height (above the ground surface); T is the mean daily air temperature (C);
es-ea represents the vapor pressure deficit (VPD) of air (kPa); and G is the soil
heat flux density (MJ/m2/day)
d ( es ) 4,098 es
== (3)
dT (T + 237.3)2
A single value of the saturation vapor pressure, es, can be estimated from air
temperature measurements as follows (FAO 1998):
17.27 T
es = 0.611 exp (4)
T + 237.3
The term exp() means e raised to the power of the term in parentheses
For 24-hour periods the saturation vapor pressure should be calculated as the
average of ea at Tmax and ea at Tmin for the day
Thus, Eq. 4 is applied once with Tmax, again with Tmin, in C, and the two
resulting vapor pressure values are added and divided by 2, as follows:
es ( Tmax ) + es ( Tmin )
es = (5)
2
The actual vapor pressure, ea, is estimated by multiplying the maximum relative
humidity by the vapor pressure at Tmin
The justification is that the minimum daily air temperature is usually associated
with the maximum daily relative humidity
Thus, ea at Tmin (Eq. 4), multiplied by RHmax is:
17.27 Tdew
ea = 0.611 exp (7)
Tdew + 237.3
When the air temperature drops below the dewpoint, water from the air
condenses on blades of grass and other surfaces
If dewpoint temperature is available, use Eq. 7 (preferred), otherwise use Eq. 6
e
Rh = 100 a (8)
es
cp P
= (9)
where has units of kPa/C; cp is the specific heat of moist air at constant
pressure (equal to 0.00101 MJ/kg/C); P is atmospheric pressure (kPa); is the
ratio of molecular weights of air to water (equal to 0.622); and is the latent
heat of vaporization (about 2.45 MJ/kg)
5.26
293 0.0065 z
P = 101.3 101.0 0.0103 z (10)
293
where P is in kPa; and z is the elevation above mean sea level (m)
The daily value of soil heat flux density, G, in Eq. 2 can estimated as:
(
G = 0.38 Tday T3 ) (11)
Other equations for estimating G can be found in FAO Irrig & Drain Paper 56
G is positive when the soil is warming, negative when the soil is cooling
That is, G is positive when the soil is adsorbing energy
The magnitude of G is almost always small compared to Rn, and can often be
considered negligible
I. Introduction
Radiation comes from other celestial sources, but most (by far) is from our sun
In general, the term radiation refers to any process which carries energy
through space
Some common types of radiation are:
1. Electromagnetic
2. Alpha and beta particles
3. Gravitational
Radiation as discussed with regard to the main energy source for crop
evapotranspiration is mostly of the electromagnetic variety
A glimpse of the electromagnetic spectrum:
2 (284 + J)
= 0.4093 sin (2)
365
2J
dr = 1 + 0.033 cos (3)
365
s = cos1 ( tan tan ) (4)
On a leap year the constant 365 in the denominator of Eqs. 2 and 3 can be
replaced by 366, but the difference in the results may not be significant
The inverse cosine argument in Eq. 4 is valid for latitudes between 55 N and
55 S (0.95993 -0.95993)
For latitudes outside this range, the argument should be less than or equal to
2.0, during the winter
If less than zero during the summer, it should be evaluated as [tan tan - 2.0]
As noted above, the value of is positive for northern latitudes and negative for
southern latitudes
Equator
40
Extraterrestrial Solar Radiation (MJ/m2/day)
35
Equator
30
52N Latitude
25
20
15
10
autumnal equinox
summer solstice
vernal equinox
winter solstice
0
0 50 100 150 200 250 300 350
Day of the Year (1 is Jan 1st)
Note that the relative distance from the earth to the sun is given as:
1
d= (5)
dr
where d is the distance from the earth to the sun, and dr is a function of J
Multiply d (from Eq. 5) by 150 to get millions of km between earth and sun (see
the plot below)
120
100
80
60
40
20
0
0 20 40 60 80 100 120 140 160 180 200 220 240 260 280 300 320 340 360
Day of Year
Of course, Eq. 6 represents a shortcut calculation and is mostly for use when
you are not using a computer
n
Rs = Ra 0.25 + (7)
2N
where n is the actual hours of sunshine during a day (hrs); and N is the
potential number of hours of sunshine in a day (hrs)
24 s
N= (8)
Eqs. 9 & 10 are versions of Eq. 7 whereby n/N = 1.0 (cloudless sky)
Note that the effect of elevation is small, compared to the 0.75 constant
The approximate minimum value of the term in parentheses in Eq. 7 is 0.25,
when there is heavy cloud cover all day long
In the figure below, the radiometer is probably out of calibration, because many
daily radiation values are significantly above the 0.75Ra curve
25.000
Solar Radiation (MJ/m2/day)
20.000
15.000
10.000
5.000
0.000
0 30 60 90 120 150 180 210 240 270 300 330 360
Day of Year (Jan 11, 1999 to Jan 10, 2000)
In winter months, some of the daily values that fall below the 0.25Ra curve may
be due to snow cover on the pyronometer
In the summer, measured values below 0.25Ra may be due to dust
accumulation or bird droppings
R
4
Rn = (1 )Rs Tmax
2 (
4
+ Tmin )(
a1 b1 es,dew a s + b
Rso
) (11)
Rns
Rnl
J + C
= 0.29 + 0.06 sin (12)
57.3
where C is 96 for northern latitudes, and 276 for southern latitudes
Estimate a1 as follows:
(
a1 = 0.26 + 0.1 exp [0.0154(J 180)]2 ) (13)
Use b1 = 0.044 for es, dew in mb, or b1 = 0.139 for es, dew in kPa
When Rs/Rso 0.7 (not too cloudy): a = 1.126; b = -0.07
When Rs/Rso < 0.7 (kind of cloudy): a = 1.107; b = -0.06
FAO Irrig & Drain Paper 56 (1998) suggests the following values for Eq. 11:
a1 = 0.34
b1 = 0.14
a = 1.35
b = -0.35
Following is another way to estimate net daily solar radiation (Allen, et al. 1993):
n
Rns = 0.77 0.25 + Ra (15)
2N
or,
and,
n
(
Rnl = 2.45(10)9 0.9 + 0.1 0.34 0.14 ea
N
) (Tkx4 + Tkn4 ) (17)
I. Introduction
Date.....July 15
Latitude.....42.4N
Elevation (above sea level).....1,195 m
Anemometer height.....3.66 m
Alfalfa height around the station.....12 cm
Temperature sensor height.....1.35 m
Dewpoint sensor height.....1.35 m
Maximum air temperature.....30.0C
Minimum air temperature.....12.8C
Dewpoint temperature.....10.0C
Average wind speed (24 hrs).....205 km/day
Global solar radiation (Rs).....641 cal/cm2 per day
Mean pan evaporation.....8.4 mm/day
Lysimeter-measured ET.....7.92 mm/day
2 (284 + 196)
= 0.4093 sin = 0.3756
365
2 (196)
dr = 1 + 0.033 cos = 0.9679
365
1.939 sin(0.740)sin(0.3756)
Ra = 37.6(0.9679) = 40.78
+ cos(0.740)cos(0.3756)sin(1.939)
17.27(10.0)
ea = 0.611 exp = 1.228 kPa
10.0 + 237.3
19.08(21.4) + 429.4
es = exp = 25.40 mb
21.4 + 237.3
or 2.54 kPa
196 + 96
= 0.29 + 0.06 sin = 0.234
57.3
a = 1.126
b = -0.07
b1 = 0.044
Finally, determine Rn
Rn = (1 0.234)(641)
11.71(10)8
2
(
[30.0 + 273.16]4 + [12.8 + 273.16]4 )
(0.354 0.044 )
12.28 (1.126(0.88) 0.07 ) = 329 langley/day
or, 329 x 0.0419 = 13.8 MJ/m2/day (because there are 0.0419 MJ/m2 per
cal/cm2, and one langley is defined to be equal to a cal/cm2)
This could be estimated based on the past three days air temperatures,
but in this case we will assume it to be negligible (which it often is)
5.26
293 0.0065(1195)
P = 101.3 = 87.95 kPa
293
or, P = 879.5 mb
4,098(25.40)
= = 1.555 mb/C
(21.4 + 237.3)2
(0.241)(879.5)
= = 0.584 mb/C
(0.622)(583.5)
2 0.67(0.12)
ln
0.123(0.12)
U2 = 205 = 181.7 km/day
z 0.67(0.12)
ln
0.123(0.12)
Ra is 40.78 MJ/m2/day
= 2.50 0.002361 (21.4) = 2.449 MJ/kg
Then, in equivalent depth of water evaporation,
(40.78 MJ / m2 / day)(1,000 mm / m)
Ra = 3
= 16.65 mm / day
(2.449 MJ / kg)(1,000 kg / m )
V. Penman-Monteith Equation
19.08(12.8) + 429.4
( es )Tmin = exp = 14.78 mb
12.8 + 237.3
= 0.795
I. Definition
Following are four principal methods for delivering irrigation water to users and
or groups of users
Several variations on these four methods exist
In fact, you can find site-specific peculiarities in almost any irrigation system,
district, or project
Here, we are talking about delivery from the conveyance and distribution
system, which may be canals and or pipelines
1. On-Demand
2. Modified Demand
3. Rotation
4. Continuous Flow
When the primary source of irrigation water is from privately-owned deep wells,
the delivery method is essentially on-demand: the farmer can usually irrigate
whenever desired and for as long as necessary
Among western USA irrigation districts, the vast majority have operational
policies espousing variations of modified demand schedules
A user typically is required to order water 24 hours in advance, and that user
usually gets the water at the desired time and with the desired flow rate
However, in water-short years, deliveries may be prorated
In most of these districts, users are charged according to volumetric
consumption of water, in addition to some fixed annual fees
Very few of these districts use system-wide gate automation
where dseed is the depth of seed placement below the soil surface (cm); and
(Rz)init is also in cm
The period for application of the initial Kc value ends at LAI 0.1, when roots
have grown down to about two times the seed placement depth
A rule of thumb on seed planting depth is that it should equal to 4 to 6 times the
seed diameter (however, this is not true for all crop types)
Maximum root depth is a function of crop type & variety, soil structure and
depth, soil layering, depth to water table, soil temperature, and water
management
The allowable depletion, dx, increases with Rz for annual crops, assuming MAD
& Wa remain the same:
dx = MAD Wa R z (2)
where Wa is soil water holding capacity (mm/m); and MAD is the management
allowed deficit
Of course, it could be argued that dx should be based on soil water tension, not
soil water content, v
Note that the equivalent Wa can change with Rz if the soil is layered, having
different textures
MAD can also change during the season, based on crop sensitivity to water
deficit (stress) in different plant phenological stages
where J is the day of year; and Kc(J) is the crop coefficient on day J
Of course, the value calculated from Eq. 3 must be less than or equal to (Rz)max
At full cover (maximum LAI), or effective cover, Rz (Rz)max
J Jplanting
R z (J) (R z )init + ( R z )
max ( z )init
R (4)
Jfullcov er Jplanting
When soil water from the root zone is extracted to the limit set by the MAD,
typical patterns of depletion are as shown below, where about 70% of water
extraction is from the upper one half of the root zone
When you plot ETc versus time, or In versus time, where In = ETc Pe, you
usually observe a large variability over time
In is the net applied irrigation depth, and Pe is effective precipitation
Thus, it seems difficult to determine an irrigation schedule that maintains soil
water above within the range of field capacity to allowable deficit, at the same
time minimizing the number of irrigations
Total costs of items such as labor, capital, and energy may increase with higher
numbers of irrigations per season
ETc Pe
depth per day = (5)
f
Eq. 5 can be used to determine an average depth per day for calculating
irrigation system capacity, and also to determine the total number of irrigations
per season for estimation of water requirements, and labor and energy
consumption
This procedure assumes that the season begins with soil water at field capacity
in the initial root zone, but that no moisture is contributed by the soil below the
initial root zone as the roots deepen with time (this may be a poor assumption)
Therefore, this method will be used to estimate a seasonal irrigation water
requirement, which is larger than that required by an amount equivalent to
about:
With real time irrigation scheduling, you need to keep a daily budget of the soil
moisture depletion
Alternatively, you could use a weekly time step, but there would be a greater
risk of experiencing crop water stress due to potentially late irrigations
The soil water depletion balance is accomplished by keeping an account of
daily (or weekly) ETc, effective precipitation, root zone depth, and net irrigation
application values
The soil moisture balance can take the following form:
where all terms are in units of depth of water (mm or inches), and
If DP > 0, resolve the soil water balance equation again to find the new Di (it
should be equal to zero when a positive DP value has been calculated)
This is because if there is deep percolation, the root zone water content is at or
above field capacity
Di1 = R z ( fc i1 ) (2)
where fc is average volumetric soil water content at field capacity in the root
zone; and i-1 is the average measured volumetric soil water content on day i-1,
in the root zone (both values as fractions)
Updating Di-1 via Eq. 2 also helps to correct for a deepening root zone
Often, an irrigator is required to submit an order for water from a canal system
one or more days before it is delivered
Therefore, it is necessary to forecast the irrigation date ahead of time, by
calculating the projected ET, P and D (for i+1, i+2, i+3, etc.)
ET can be forecasted by examining ET for the past week, or for the same
period in past years (if such data are available)
As an alternative, one can forecast ET and precipitation by using forecasts of
the parameters used in ET equations, such as air temperature, relative
humidity, percent sunshine or cloudiness and wind speed
These parameters can be predicted fairly accurately up to 7 days in advance in
some areas by using satellite data and atmospheric models
1. Kc versus time
2. ETo from weather data
3. Rz versus time
4. Wa
5. MAD
6. Infiltrated precipitation
7. Net application of water (on low quarter of field)
8. Occasional soil moisture measurements to update depletion estimates
If you know the net depth of applied water on the low , you will also know it for
the whole irrigated area (the low will be discussed in a later lecture on
irrigation efficiency and evaluation of system performance)
With sprinklers, the net application of water can be estimated as qt (i.e. volume
of water) from nozzle discharge-pressure relationships
With surface irrigation, soil moisture measurements must be made, unless the
irrigation is long enough to ensure recharge of soil water to field capacity
Alternatively, you could use an inflow-outflow measurement for the entire field
V. Supplemental Irrigation
1. Computer Programs
2. Voice Synthesizers
Irrigation projects may publish ETc in newspapers and farmers can maintain
their own water balance
TDR
neutron probe
gravimetric
gypsum blocks (threshold readings to predict irrigation)
tensiometers (threshold values)
"feel" by hand
6. Infrared Thermometry
I. Introduction
Irrigation water quality can involve many chemical and biological factors,
including salinity, presence of pesticides, and others
This lecture focuses primarily on irrigation and soil water salinity
All irrigation water contains salts, which can accumulate to problematic levels in
the crop root zone if not managed through leaching and drainage
The presence of salts can have detrimental effects on plants in different ways
For example, some of the chemical constituents in salts can be toxic to certain
types of plants even when present in very small concentrations
Salts can also harm the soil structure and increase the osmotic potential,
thereby making it more difficult for plants to withdraw water from the soil
Following are three main reasons why soil water salinity can be harmful to
agricultural production:
Most discussions of soil water salt refer in general to dissolved minerals in the
water, not just sodium chloride (table salt)
These commonly include compounds of calcium, magnesium,
potassium, sodium, chloride, sulfate, carbonates and bicarbonates,
and nitrates
All soil water has some dissolved minerals, though the chemical
composition varies greatly
High concentrations of salts are detrimental to crop growth &
production
There is a wide range of crop tolerance to salts; some crops are very tolerant to
high soil water salt concentrations, and others are very sensitive
Problems with soil water salinity are usually worse in arid and semiarid areas
For most crops and conditions, 5 to 10% of the applied water must pass down
through the soil to carry away excess salts
This can be accomplished by annual rains in many cases, eliminating the need for
the irrigation system to perform the leaching
FAO Irrig & Drain Paper 29 recommends the following equation for determining
the leaching requirement (LR), based on
where ECw is the EC of the irrigation water (dS/m); and ECe is of the
soil water extract for a 10% yield reduction (dS/m)
Electrical Conductivity,
Effect on Crop Yield
ECe (dS/m)
0 1 Reduction in yield will usually be negligible
24 Yield of sensitive crops will be restricted
48 Yield of many crops will be restricted
8 16 Yield is satisfactory only for salt tolerant crops
100
80
Relative Crop Yield (%)
Threshold Salinity
60
40
20
0
0 Increasing Salinity
80
60
unacceptable
for crops
40
tolerant
moderately
20 tolerant
moderately
sensitive sensitive
0
0 5 10 15 20 25 30 35
ECe (dS/m)
Bean Apple
Carrot Pear
Strawberry Almond
Citrus Avocado
The following table (compiled from several sources) shows threshold and slope
values for several crop types
By the slope values in the third column, you can get an idea of how fast the
yield tends to decrease with increasing salinity
It is usually easier to remove salts from sandy soils than from silty clay soils, and it
should be recognized that leaching works well only with adequate drainage to
carry the salts away
Compacted zones and changes in soil texture reduce downward water movement
through the soil and result in less leaching
Salt must always be managed so it accumulates in zones away from germinating
seeds and plant roots
Drip and other low flow irrigation methods can be used to flush salts away from
seeds and or plant roots
These methods are also effective, particularly on sandy soils, in decreasing deep
percolation and in reducing the farm irrigation water requirements
When salts accumulate between the plant rows, they may be leached during
periods of sufficient rainfall or periodic irrigation to remove them from the depth of
soil usually explored by plant roots
Thus, irrigation system capacity does not necessarily have to include LR because
sufficient capacity may be available for leaching during non-peak crop
consumptive use periods
Na+
SAR = (2)
(
0.5 Mg ++
+ Ca +
)
in units of meq/liter
Na +
SARadj = ( 9.4 pHc ) (3)
(
0.5 Mg ++
+ Ca +
)
where,
( )
pHc = pK '2 pK 'c + p ( Ca + Mg) + p ( Alkalinity ) (3)
An adjusted SAR value of less than 6 is usually acceptable, but a value greater
than 9 is usually unacceptable
The exchangeable sodium percentage, ESP, is another definition of sodium ion
content in a soil, and is approximately (through a statistical relationship) defined
as a function of SAR:
Saline soils have relatively high levels of soil water salinity and can be
reclaimed by leaching
Alkaline soils are those that have a pH greater than 7 (they can be either sodic
or calcium carbonate-rich soils), usually with relatively high SAR values
Black alkali soils are formed when saline-alkali soils are irrigated and the salts
are leached out, leaving dissolved organic matter; they have a high pH and low
permeability, and the soil surface may be slick; ESP is greater than 15%
Sodic soils are those with SAR or exchangeable sodium percentage, ESP,
greater than 15, or in other words soils with sodium as the dominant cation
The reclamation of sodic soils may involve gypsum (calcium sulfate, CaSO4),
but if there is a source of calcium carbonate in the soil, sulfuric acid may be
used to dissolve calcium carbonate and exchange the sodium with calcium,
whereby the sodium could be leached downward and beyond the root zone
Boron is present in water from most sources, but the concentration varies from
trace amounts to several parts per million
Boron is a micronutrient, essential to plant growth, but become toxics at
concentrations only slightly above the optimum level
Many plants manifest significant yield reduction effects when the boron
concentration is 1.0 ppm or more
A salt-affected soil is a soil which has enough dissolved minerals that it cannot
support optimal crop growth
Many soils in arid and semi-arid regions are considered to be salt-affected, and
they require management to sustain agricultural productivity over the long term
The major types of soils with respect to salts are:
1. Normal
2. Saline
3. Sodic
4. Saline-Sodic
The types of salt-affected soils have been defined as follows (James et al. 1982):
Saline soils are sometimes called white alkali because of the presence of a white
crust of salt on the soil surface, but not all saline soils have such a crust
The pH of a sodic soil can be as high as about 10, indicating a high level of
alkalinity
The presence of excess salt in saline-sodic soils helps maintain the permeability
of the soil, compared to that of a sodic soil
Thus, leaching of salts from a saline-sodic soil can (but not necessarily) cause the
soil to become sodic only
As you have probably surmised by now, the science and management of salt-
affected soils is a very deep and complex subject in which a person could dedicate
their entire professional career
James, D.W., R.J. Hanks, and J.J. Jurinak. 1982. Modern irrigated soils. John Wiley &
Sons, N.Y. 235 pp.
I. Introduction
net output
efficiency , or (1)
gross input
Some of these are discussed in this lecture, others will be defined in the
subsequent lecture
Many other definitions have been proposed and applied
Vd
Ec = 100 (2)
Vs
where Ec is in percent; Vd is the volume of water delivered to farms at turnouts
or lateral/distributary canal inlets; and Vs is the volume of water entering the
conveyance system from a source (river, lake, well, reservoir, other)
1. flow rate
2. total reach length
3. conveyance system age
4. type and condition of lining material
5. adequacy of infrastructure maintenance
6. type of soil adjacent to the channels or pipes
7. hydraulic radius & wetted perimeter (open-channel flow)
Vz
Ea = 100 (3)
Vd
where Ec is in percent; Vd is the volume of water delivered to farms at turnouts
or lateral/distributary canal inlets; and Vz is the volume of water stored in the
crop root zone
1. surface runoff
2. deep percolation
3. evaporation and wind drift
The above are considered to be losses from a field perspective, and sometimes
from a farm perspective, but they are not all necessarily losses at larger scales
It can be said that the value of Ea is a function of:
1. irrigation method
2. irrigation timing and duration
3. irrigation system design & current condition
4. uniformity of water distribution at the surface
5. soil type and structure
6. soil infiltration rate
7. irrigator skill
It is not enough to evaluate irrigations based only on the above definition for Ea
because it can be misleading
For example, if you practice deficit irrigation (incomplete irrigation), most of the
applied water may be stored in the root zone, but the average depth was not
nearly enough to bring v to field capacity
On the other hand, the application uniformity might be perfect (same depth
applied all over the field), but the irrigator ran the system too long and applied
too much water, so the Ea looks bad, but it is not due to system design nor
maintenance
Other efficiency terms are necessary to more fully qualify the Ea value
Vz
Es = 100 (4)
V reqd
where Es is in percent; and Vreqd is the required volume, which is equal to the
net irrigated area multiplied by the soil water deficit:
V reqd = ( FC v ) R z A (6)
where v is the soil water content just before beginning an irrigation; FC is the
water content at field capacity; and A is the net irrigated area
Deep percolation is the water that infiltrates into the soil from irrigation and
precipitation, passing downward through the crop root zone
Some of the deep percolation can later move upward, reentering the root zone,
but this is usually negligible
Almost all complete irrigations have some deep percolation, where a complete
irrigation is one in which the root zone is refilled to field capacity over 100% of
the field area
Runoff is the water that leaves the irrigated area in the form of surface
drainage, usually at the low end of the field, or the tail of the field
Runoff is sometimes referred to as tail water
Efficiency terms can be presented to quantify deep percolation and runoff
losses separately, relative to Vd, thereby helping to explain the magnitude of Ea
Vz
Ep = 100 (7)
Vs
This is defined as the percentage of water diverted to an irrigation project (or
district) which is stored in the root zone and is usable to satisfy crop water
requirements
This is not really an efficiency term, rather a water application uniformity term
The following equation is the definition for Christiansens coefficient of
uniformity, CU:
n
zi z
Ed = 100 1 i=1 (8)
nz
2.5
65%
2.0
70%
75%
80%
85%
1.5
Relative depth
90%
95%
1.0
0.5
0.0
0 10 20 30 40 50 60 70 80 90 100
Percent area
Example #1
head tail
no runoff
zreqd
Example #2
head tail
no runoff
zreqd
no deficit
no deep percolation
Ea = 100% Es = 100% Ed = 100%
Example #3
head tail
no runoff
zreqd
no deficit
deep percolation
The dashed horizontal lines in #1 and #3 are the depths of infiltrated water
Examples #2 and #3 represent complete irrigations (no deficit)
zreqd
deficit
Example #2
head tail
runoff
zreqd
deficit
deep percolation
runoff
zreqd
no deficit
deep percolation
Water enters the field on the surface at the head, and any runoff leaves at the
tail, to the right
The dashed curves represent infiltrated depths; they are curved because of the
effect of intake opportunity time
Note that for surface irrigation, uniformity tends to increase with increasing
depth, because as seen by the Kostiakov equation, the infiltration rate
decreases towards the basic rate, fo, with opportunity time
That is, the infiltration curve becomes more horizontal with infiltrated depth, and
Ed increases
zreqd
Example #2
zreqd
average depth = required
zreqd
no deficit
deep percolation
You can search the technical literature and find dozens of articles related to
irrigation efficiency, uniformity and related topics
Willardson, Allen & Frederiksen (1994) wrote a paper entitled Elimination of
Irrigation Efficiencies, instead proposing fractions of water
That paper objected to some inappropriate applications of traditional efficiency
terms, whereby mostly nonexistent water savings were to be diverted
elsewhere
Andrew & Jack Keller presented a similar paper in 1994
Willardson previously published other papers on irrigation efficiency
Many others have also written papers on this topic
For example, PELQ and PELH
z
PELQ = 100 25% (9)
z
where z25% is the average depth of water stored in the crop root zone when the
average depth infiltrated in the lowest quarter of the field area is equal to the
soil water deficit, zreqd; and z is the mean gross application depth over the
whole field
According to the above definition, the irrigation is not complete, but about 75 +
(25) = 87.5% of the field area is over-irrigated (has some deep percolation)
One-eighth (12.5%) of the field area has some amount of deficit both before
and after the irrigation
Thus, PELQ is an efficiency term based on the uniformity of infiltration in the
low of the field
For sloping surface-irrigated fields, the low of the irrigated area is usually at
the tail end of the field
PELQ is a uniformity-based quantification because it uses the mean gross
application depth as the basis (Eq. 1), which explains the qualification that the
average depth infiltrated in the lowest of the field area is equal to zreqd
runoff
zreqd
low
deficit
deep percolation
75%
87.5%
z
PELH = 100 50% (10)
z
where z50% is the average depth of water stored in the crop root zone when the
average depth infiltrated in the lowest half of the field area is equal to the soil
water deficit, zreqd
According to the above definition, the irrigation is not complete, but about 50 +
(50) = 75% of the field area is over-irrigated (has some deep percolation)
One-quarter (25%) of the field area has some amount of deficit both before and
after the irrigation
PELH is also an efficiency term based on the uniformity of infiltration
Both PELQ and PELH account for evaporation, wind drift (sprinklers), surface
runoff, nonuniformity, and irrigation adequacy
The greater the variability of irrigation depths within a field, the greater the
difference between the net depth of the lowest and the average gross depth
applied
Therefore, for any given evaluation conditions, the PELQ < PELH
But, PELQ requires that 87.5% of field area be adequately or over-irrigated,
essentially ensuring adequacy, even though the soil water content in about
12.5% of the area is somewhat less than field capacity
If the average net application depth of the lowest of the area is different than
the soil water deficit, then the AELQ can be used
This is the application efficiency of the low quarter
AELQ is defined as:
z
AELQ = 100 25% (11)
z
where Z25% is the average depth stored in the root zone in the 25% of the area
with the least amount of infiltration from irrigation
AELQ is similar to DU
1. Sprinklers
Sprinkler PELQ (%)
Type Range Average
Hand-move 50 - 70 65
Side-roll 50 - 70 65
Solid Set 60 - 75 70
Center Pivot 70 - 85 75
Linear Move 65 - 85 75
Big Gun 55 - 65 60
Traveler 60 - 80 70
2. Trickle/Drip
70 - 95% depending on design, management, and maintenance
3. Surface: Borders
60 - 80% for well graded and well managed systems
30 - 60% for poorly graded and or poorly managed systems
4. Surface: Furrows
50 - 70% for well-graded and well-managed systems
30 - 50% for poorly-graded and or poorly-managed systems
60 - 90% for surge-flow irrigation with tail water recovery
I. Introduction
1. Open ditches
2. Buried pipelines, or
3. Gated pipe laid on the surface
Laser equipment has been used for more than 30 years in the construction
industry, and it is also used for laying pipelines, excavating ditches and canals,
and lining canals with concrete
Lasers are used for conventional surveying jobs, except that in many cases
only one person is required to complete the work
The laser is mounted vertically, and a mirror at a 90-degree angle rotates above
This essentially establishes a plane of light over the field surface, and when the
job is finished, the field surface is parallel to the plane of light
Problems will arise when two separate laser sources overlap in a field because
the sensor doesn't know which one to use
Most laser equipment is sensitive to movement (e.g. from gusty winds), and the
mirror will stop rotating if the laser unit is shaken slightly, resuming again after
automatically re-setting to the specified slopes
Equipment can work day and night, but heavy fog and thick dust can obstruct
the laser beam enough to prevent operations
These problems are worse when the sensor is far from the laser itself (in fields
of about 20 ha or more)
Hot and dry conditions can also cause problems when the sensor is far from the
laser and the air has strong currents, causing the beam to diverge and fluctuate
This in turn causes the sensor to feed fluctuating signals to the controller, giving
a "washboard" effect to the cut and fill areas
The equipment can be used for both surveying the existing field elevations, and
for doing the leveling work
The survey job is not as accurate as a conventional manual survey, but it is
quick and easy
Leveling calculations are slightly complicated by the fact that the survey is often
performed using a sloping reference plane
This is done because the telescoping mast upon which the sensor is mounted
has limited vertical travel, and some parts of the field could go out of its range
The operator must check the sensor position at the high and low corners before
beginning the surveying process
Three common modes of operation are:
With deep cuts, the operator may need to make a few initial passes in manual
mode, then make the final cuts in automatic mode "on-grade"
Otherwise, the scraper and tractor may be forced to bite off too much at a time
The equipment is usually taken down and stored when not in use
Re-setting the equipment in the field before a job is completed can be a
problem for the following reasons:
1. It may be difficult to line up the main (in the direction of furrows or borders)
and side slopes to exactly the same directions as before;
2. The plane of laser light may be at the same slopes as before, but the height
above the field surface has changed, requiring a corresponding change in
the elevation difference between the sensor and the scraper blade
Eq. 1 gives the ground slope in the x direction, as shown in the figure below
The slope in the y direction would be:
Thus, for an angle of zero ( = 0) the x and x axes coincide, as do the y and y
axes, and from Eqs. 1 and 2 the slopes are simply Sx and Sy
For any other angle the slopes are different because the coordinate system is
rotated
1. Advance
2. Wetting (or Ponding)
3. Recession
The figure below shows a case that is more typical of a blocked-end border, in
which the recession phase lasts much longer
This is in contrast to a free-draining border, where surface runoff would cause
the water to recede from the ground surface much faster
Other curve shapes and configurations can be obtained in practice; the two
figures shown here are idealized examples
For example, the inflow rate could be varied (cutback) to reduce runoff
Or, there could be cracks in the soil or obstructions, causing discontinuities in
the advance curve
Or, the advance curve could become vertical, indicating that total infiltration
matches inflow (advance stops)
Notice that the intake opportunity time, as used in the Kostiakov equation, is
equal to the vertical difference between the advance and recession curves, at
any given distance
The shapes of the advance and recession curves affect application uniformity
If the recession curve is fairly flat, better uniformity will be achieved by having a
quick advance
On the other hand, if the advance is too quick, soil erosion will occur
The time of cutoff also affects uniformity insofar as the infiltration rate various
with opportunity time
The time of cutoff also affects the average applied depth of water, as it does in
pressurized irrigation methods
Z = ka + fo (1)
x = p tr (2)
This simple function approximates the advance curve very well in most cases
It can be calibrated by logarithmic transformation and regression, or by a two-
point method
It is possible to write a simple volume balance equation for the advance phase
in which all inflow goes to soil infiltration and surface storage
Evaporation losses are neglected and the inflow rate is assumed constant
Here is the equation:
N
( ) f t
Qo t = y A o x + z kt a x + o x
1 + r
(3)
inflow surface
water infiltrated
water
where the left side of the equation is the inflow volume; Qo is the inflow rate; t is
the elapsed time since the introduction of water into the furrow; y is a surface
shape factor; Ao is the cross-sectional flow area at the furrow inlet; x is the
advance distance; and z is a subsurface shape factor for the first term in the
Kostiakov equation
Note that the intake opportunity time, , is equal to the elapsed time, t, because
the infiltration terms are defined for the furrow inlet (where the advance time is
zero)
The surface storage term, yAox, is also defined for the furrow inlet
The term y is usually between 0.7 and 0.8, and is often taken to be 0.77
The subsurface shape factor z is defined from geometrical considerations as:
a + r(1 a) + 1
z = (4)
(1 + a)(1 + r)
1
Qo = AR2 / 3 So (5)
n
where R is the hydraulic radius (A/Wp); n is the hydraulic roughness; and So is
the longitudinal furrow slope
Rearranging Eq. 5,
1/ 2
Q2n2
Ao = o (7)
1So
where the value of the roughness, n, is assumed based on the size and
condition of the furrows; Qo is in m3/s; So is in m/m; and Ao is in m2
Qin Qout
fo = (8)
L
Flow rates are generally measured using portable flumes which are installed at
the furrow inlet and outlet before beginning the test
It is important to wait until the outflow becomes constant before applying Eq. 8
Taking Eq. 3 and writing it once for x = L, then again for x = L/2, the following
are obtained:
( ) f t
Qo tL = y A oL + z k tLa L + o L L
1+ r
(9)
and,
L
2
( )
L f t L
Qo tL / 2 = y A o + z k tLa / 2 + o L / 2
2 1+ r 2
(10)
VL =
Qo t L
L
f t
1+ r
( )
y A o o L = z k tLa (11)
and,
VL / 2 =
2Qo tL / 2
L
f t
(
y A o o L / 2 = z k tLa / 2
1+ r
) (12)
where the terms VL and VL/2 are calculated based on the known values between
the two equals signs
The unknowns in Eqs. 11 & 12, a and k, are at the far right side, after the
second equals sign
Combining Eqs. 11 & 12, the Kostiakov exponent a can be determined:
log VL log VL / 2
a= (13)
logtL logtL / 2
VL
k= (14)
z tLa
Dont show more than 2 or 3 significant digits in your calculation results, but
dont round intermediate calculated values
W.R. Walker and G.V. Skogerboe (1986). Surface Irrigation Theory and Practice. Prentice-Hall.
Z = ka + f o (1)
where Z is the cumulative infiltration (m); is the intake opportunity time (min); k
is an empirical coefficient; a is an empirical exponent; and fo is the basic intake
rate (m/min), as previously defined
For a complete irrigation, the least-watered area of the field must have the
required intake opportunity time
The least-water area of a surface-irrigated field is usually the tail end
Thus, the time of cutoff, tco, is equal to the time of advance to the tail end, plus the
required intake opportunity time
The time of advance can be determined from a calibrated advance equation:
L = p tra,L (2)
where L is the furrow or border length (m); ta,L is the time of advance to the end of
the field (min); and p and r are empirically fitted parameters
where,
r
Ap = p ta (3)
where Ap is the percentage of the total basin area (%); ta is the advance time
(min); p is an empirical coefficient; and r is an empirical exponent
Note that Eq. 3 is calibrated in terms of surface area, not advance distance
1/0.83
100
t a100% = = 164 min (4)
1.45
(b) infiltrated depth at the basin inlet when the advance finishes
0.565
Zo = 0.0029(164 min ) = 0.052 m (5)
(e) volume ponded on the surface at the end of the advance phase
The value of tc is calculated assuming that the infiltrated volume of water during
the advance phase is negligible, thus assuring that the irrigation is "complete" (all
parts of the basin receive the required application, or more)
(a) Calculate the time of advance for each of 20 sub-areas. The time of advance
is determined from the values of p and r.
(b) Calculate the intake opportunity time and infiltrated depth for each sub-area.
The intake opportunity time at the end of the advance phase is equal to the time
of advance at 100% of the area minus the time of advance to each sub-area.
The infiltrated depth is calculated based on the respective opportunity times and
the values of a, k, and fo.
2 3
V inf = (0.6972 m)(0.05)(5420 m ) = 189 m (12)
(f) volume of water ponded on the surface at the end of the advance phase
3
V surf = 768 - 189 = 579 m (14)
(h) required irrigation time according to the difference between Vreq y Vsurf
In the example, the respective cutoff times for methods 1 and 2 are nearly equal
In the first method a value of z = 0.65 was assumed, and this value can be
calculated for the second method as follows:
The average application was 0.6972 m/20 = 0.0349 m, and,
Therefore, the value of 0.65 used for the first method was almost correct, and this
is why the results from the second method were very close to those of the first
Finally, suppose that instead of calculating the time of cutoff as shown in these
examples, the time of cutoff were computed as for furrows or borders
In this case, the time of cutoff would have been:
1/0.565
0.113m
req = = 654 min (19)
0.0029
and,
t c = 164 + 654 min = 818 min (20)
This is much greater than the required time of about 171 minutes
The time of cutoff could also be determined by dividing the required volume of
water by the inflow rate:
612 m3
tc = = 131 min (21)
(60 s / min)(0.078 m3 / s)
A time of cutoff of 131 minutes gives the correct volume, but does not take into
account the fact that there will be some deep percolation near the inlet location,
and some deficit at the far end of the basin
Most irrigators would not choose a time of cutoff of 818 minutes for the previous
example of a level basin; they would choose a time closer to 171 minutes, even if
they had not done any calculations
The runoff from free-draining borders and basins could be reused in lower areas,
or pumped back onto the same field; in these cases, it may be preferable to use
the maximum nonerosive stream size so that the time of advance is minimized
A good irrigator may be able to come very close to an optimum balance between
deep percolation and runoff losses, but most focus exclusively on runoff because
that is what you can see, and is often how others judge the irrigators ability
A thorough irrigation system evaluation will involve three components: (1) field
observations, (2) field measurements and data analysis, and (3) an interview
with the farmer or irrigator
The third component can be a very important part of an evaluation because it
can provide indications as to why the system is maintained and operated the
way it is, and can uncover additional problems that may not have been
discovered during the field observations and measurements
The interviewer should ask the farmer what kinds of irrigation problems are
normally experienced, and what are his or her perceptions as to how the
performance of the irrigation system might be improved
Is the soil very dry or very wet (use soil augers or a shovel)? Is the
soil dry enough that an irrigation should begin, or so wet that irrigation
in progress should stop?
Are there any signs of soil erosion, especially at the head end and tail
end of the field? Are the flow rates too high or are the field slopes too
steep?
Are the headworks (ditches, structures, pipes, etc.) well maintained?
Are the furrow or basin inlets eroded? Is gated pipe discharging
directly into the furrow without any energy dissipation?
Are there low spots in the field (ponded water, wet and perhaps
weedy areas)? Does the field need to be graded or leveled?
Are there sandy areas in the field that tend to dry much quicker than
the rest of the field?
Are there heavy clay areas that dry much slower?
For borders and basins, are the dikes in good condition?
For furrows, are the furrows deep enough and without obstructions
(clods, straw, vines, etc.)?
Are there deep cracks in the furrows due to soil shrinkage?
For furrows, is the soil crusted on the beds (did the beds become
inundated during the irrigation)?
Does it appear that water jumped across furrows during the
irrigation?
For sprinklers, is the water being blown off the field by strong winds?
Are the sprinklers rotating evenly, and are there any visible leaks in
the pipes?
For sprinklers, is the water droplet size obviously very small (high
pressure, high evaporation, low uniformity) or very large (low
pressure, soil surface sealing, low uniformity)?
Are salts visibly accumulating on the soil surface?
Are there any barren areas that dont even support weed growth
(alkaline soils, sodic soils, lack of water, etc.)?
Does the crop growth appear to be uniform across the field?
Field measurements are also an important part of an evaluation, but care must
be taken to present the results in a comprehensible way such that the results
can be applied toward irrigation system improvements
General measurements may include soil infiltration rate using cylinder ring
infiltrometers or other devices, soil water, and water quality analysis
Merriam and Keller (1978) and Walker and Skogerboe (1986) describe detailed
technical procedures for performing irrigation system evaluations
Field length can be modified, but the possibilities are usually very limited. For
example, the field length can be divided by two when the advance rate is too
slow, then later in the season it may be feasible to irrigate over the entire length
Field topography can be modified by land grading or leveling, which can
significantly affect the uniformity, labor requirements, weed control, and time to
dry between irrigations
Infiltration rates can be affected by mechanical or chemical means, or by using
surge flow
Farmers will sometimes run furrows with a tractor to crush clods and reduce
the infiltration rate, thereby allowing faster water advance and better uniformity
Or, it may be necessary to chisel the soil to provide increased infiltration and
aeration
x = p tra (1)
Z = ka + fo (2)
where Z is the cumulative infiltrated depth (m2); is the intake opportunity time
(min); k is an empirical coefficient; a is an empirical exponent; and fo is the basic
intake rate (m2/min)
Zreq L
Ea = 100 (3)
Q tc
Zreq x d + Vzd
Ea = 100 (4)
Q tc
where, Ea is the application efficiency (%); Zreq is the required application (m2); L
is the length of the furrow or border (m); Q is the discharge per furrow or per
unit width (m3/min); tc is the time of cutoff (min); xd is the distance to the
beginning of the deficit (m); and Vzd is the volume infiltrated beyond xd (m3)
Vz Zreq L
DPR = 100 (5)
Q tc
(2) Irrigation with deficit:
Vz Vzd Zreqx d
DPR = 100 (6)
Q tc
Zreqx d + Vzd
WRE = 100 (8)
Z L
req
The values Vz and Vzd can be calculated by numerical integration using the
equations for advance and infiltration
The time of cutoff is the number of minutes since the beginning of the irrigation
until the inflow is discontinued at the head of the field
The value of WRE will be equal to 100% for a complete irrigation, and will be
less than 100% when the irrigation is incomplete
All units are in meters and minutes
It is assumed that the recession phase is nearly instantaneous, and that the
inflow rate, Q, is constant
Also, it is assumed that the soil is uniform with the same infiltration
characteristics everywhere, and that there are no obstructions (e.g. straw or
other organic matter) or large cracks in the soil
The Kostiakov equation can be calibrated from cylinder ring infiltrometer tests,
or from advance tests, as described previously
The p and r parameters for the advance function can be determined from
advance tests
The required input data for the above evaluation procedure are:
Get a copy of Farm Irrigation System Evaluation: A Guide for Management (J.L. Merriam & J.
Keller, 1978)
Get the book Surface Irrigation Theory & Practice (W.R. Walker & G.V. Skogerboe, 1986)
Check out the ASAE standards
Advantages of Sprinklers
Disadvantages of Sprinklers
1. Systems are expensive to purchase and install ($1,000 to $6,000 per ha)
2. Susceptibility to clogging of emitters, which usually have very small
openings so, it is important to spend time and money on maintaining the
system, applying chemicals, and keeping filters clean
3. Possibly low distribution uniformity due to low operating pressures and
possibly due to steep slopes, especially along laterals
4. Where laterals are on steep slopes, the water will drain out the downhill end
at every startup and shut-down.
5. Soils with very low intake rates will exhibit ponding and runoff
6. Salt tends to accumulate at the soil surface and around the wetted area --
when it rains, these accumulated salts may be driven into the root zone
7. These systems tend to require more capable and diligent management
because of the susceptibility to clogging, and because the systems are
usually designed to operate continuously during peak ET periods (cant
afford to let the system shut down during these periods). These systems do
not usually take full advantage of the soil storage (buffer) capacity.
Sprinkler irrigation can be restricted to crops that can withstand complete foliar
wetting
Some crops such as tomatoes cannot tolerate wetting when the fruit matures
because it would cause molding and other damage
However, in some cases fungicides and other chemicals are injected into the
irrigation water to prevent quality degradation in the fruit
In orchards it is often necessary to use under tree sprinklers to avoid wetting
the leaves and fruit, but in crops such as bananas a big gun sprinkler can be
used for above canopy supplemental irrigation
Water is applied to the place where the active plant roots are to be found
There are many types of emitters and misters that are capable of supplying
water directly to the plant root area
The soil water may be maintained near field capacity at all times
The system may be located on or under the soil surface, or elevated above
ground along rows of trees or vines
Operating pressures are usually low (one half to one atmosphere or 7 to 15 psi)
The installation may be nonportable, semi-portable or portable
Water can typically be applied more frequently and efficiently than with other
methods, and evapotranspiration is not reduced by reducing the soil water
below field capacity
In some cases surface evaporation is reduced but transpiration may be
somewhat increased due to maintenance of low water tension in the root zone
Therefore, the influence on evapotranspiration can be considered to be
negligible
Deep percolation and surface runoff can be reduced
to minimum values
The low irrigation rate also makes the method
suitable for low-infiltration soils
Water can be saved by not wetting the area
between rows or between plants.
Field operations are easier because much of the soil surface remains dry
Weed growth is reduced, and uninterrupted orchard operations are possible
When crops are grown on beds the furrows in which farm workers walk remain
relatively dry
Fertilizers can be injected into the irrigation
water to avoid the labor for ground application
Greater control over fertilizer placement and
timing may lead to improved fertilizer
efficiencies
The possibility of groundwater contamination
due to deep percolation can be significantly
reduced
Fertilizers are usually applied in dissolved form through the irrigation water
Herbicides and soil sterilants can also be applied with the irrigation water
Crop yields are often higher and of better quality
Sometimes the timing of harvest can be improved
Localized irrigation is expensive, but in many cases increased quantity and
quality of yields together with improved timing of harvest has justified the
expense
IV. LEPA
1. Write notes for the test and include the names of the evaluators, the date, the
location, the sprinkler type and any available technical information about the
sprinkler, plus the wind direction and speed. Include any relevant comments
about the site conditions, and if possible, take photographs of the location.
2. Include a plan view sketch of the layout, showing the sprinkler location and the
can locations, and showing the cardinal directions.
3. Record the catch-can spacing and the number of cans to each side of the
sprinkler. The spacing is usually from 1 to 2 m.
4. Measure and record the riser height, hr, which is the height of the sprinkler
above the ground.
5. Turn on the pump (or open a valve) and either cover other sprinklers or prevent
them from rotating into the test area. In some cases the evaluation is for a
single sprinkler, and for others it is for a line of overlapping sprinklers. The
evaluation could also be for multiple lines (laterals) operating simultaneously.
6. Measure the sprinkler flow rate, possibly using a section of inner tube, a bucket
and a stopwatch. This can be done twice: once at the start of the test, and
again at the end of the test.
7. Measure and record the pressure at the sprinkler nozzle with a pitot tube
gauge, or with another type of gauge on the riser pipe.
8. Adjust the valve so the pressure is within the manufacturers recommended
range, or whatever pressure is desired. However, the idea is usually to do the
evaluation for representative field conditions, whatever they might be.
1. This is much the same as for sprinklers, except that a determination of emission
uniformity (EU) is done instead of a catch-can test.
2. Emitter performance can be evaluated separately by determining the
manufacturers coefficient of variation of flow rate (for a constant pressure).
3. Also, with drip irrigation, the percent wetted area can be measured to determine
the adequacy of the system in meeting crop water requirements.
4. Trickle systems always have filters, so a part of the evaluation is to check the
condition and adequacy of the filters and screens. For example, the filters may
be dirty and in need of maintenance, or there may be leaks. Alternatively, the
filters may be clean and in good condition, but the capacity is insufficient,
causing higher than normal pressure loss.
5. Another evaluation aspect is the chemical storage and injection equipment: are
they in good condition?; are there leaks? are chemicals stored safely and
according to state or local regulations?
n
j=1 (
abs z j m )
CU = 100 1.0
n
z
j=1 j
(2)
n
n ( dr )
i=1 i i
r d i=1 i i
n
r
CU = 100 1.0 i=1 i (3)
i=1( dri i )
n
Eq. 3 is different than Eq. 2 because different sprinklers along a center pivot
lateral cover different areas of the field, and the catch-can data must be
weighted appropriately
ASAE Standards. (1997). American Soc. of Agric. Engineers, St. Joseph, Michigan.
Farm Irrigation System Evaluation: A Guide for Management. (1978). J.L. Merriam & J. Keller,
BIE Dept., USU, Logan, Utah.
I. Introduction
These valves, or vents, serve the double purpose of removing entrapped air
from the line upon filling, and introducing air upon the evacuation of water
(dewatering) when an upstream valve is closed or a pump is turned off.
It is recommendable to remove air from the line during filling because air can
accumulate at high points and bends, acting as a full or partial blockage of the
water flow
Trapped air can completely block the flow,
even while a pump is operating
An air release valve should close, or seal, at
a pressure of only 2 or 3 psi
Pockets of air can suddenly dislodge and
move downstream, causing pressure and discharge surges due to the relatively
high compressibility of air
Air can enter a pipeline in these ways:
It may be necessary to allow air to enter the pipeline under the following
conditions:
Some valve designs are only for vacuum relief, not for air release
If air does not enter the pipeline to replace the exiting water, the pipe can
collapse (due to internal pressures below atmosphere pressure), or the
temporarily low pressures can cause water to reverse direction and slam into a
valve or pump, causing surges or water hammer
These kinds of problems occur in many pipelines, including pressurized
irrigation systems
Locations at which it is recommended to install combination vacuum relief - air
release valves:
Stand pipes are vertical pipes connected to a buried pipeline, and open to the
atmosphere
These serve all of the purposes mentioned above, and also serve to limit
pressure in the line (they will overflow when the pressure head in the pipeline
exceeds the height of the stand pipe)
Stand pipes should have at least 0.3 m of free board so that they do not
overflow frequently
This means that the height should be at least 0.3 m above the normal operating
pressure head of the pipeline
However, the free board should not be so high as to allow the pipeline or pipe
joints to burst
Pressure relief valves protect the pipeline and other hydraulic components from
excessively high pressures, which may result from surges, water hammer, or
improper pump and valve operations
The valves open automatically according to preset pressure values.
Two basic pressure relief valve designs are:
1. Spring-actuated
2. Diaphragm
Spring-actuated valves are less costly, but have the disadvantage that they do
not fully open until the pressure increases significantly above the initial opening
pressure
That is, they begin to open at a specified pressure, and incrementally increase
to full opening if the pressure continues to increase
But they are almost always adequate for pressurized irrigation lines
Some have outlets for diverting discharge away from the valve location
Some have manually-operated release levers
The required valve size depends on the flow rate of the pipeline
It is important to not install pressure relief valves that are too large because
they can open and close rapidly, causing surges or water hammer in the line
This can occur when a large valve opens and suddenly lowers the local pipe
pressure, closes, then opens again when the pressure increases, etc.
High pressures can occur due to surges in the pipeline
In most cases, these surges result from:
Another option to control hydraulic transients is to use stronger pipe (lower SDR
or stronger material)
However, it should be noted that although steel pipe has relatively high burst
pressure, it can still collapse due to pressures below atmospheric unless steps
are taken to strengthen the pipe (e.g. used welded rings on the exterior)
Some industrial applications use rupture discs which break open at a certain
pre-set pressure
When a disc ruptures, it must be replaced with a new one
These are rarely, if ever, used in irrigation systems
1. Check valves
2. Back-flow prevention valves
3. Vacuum-breaker valves
4. Air-gap separation
When filling a pipeline, the water should enter slowly, at a flow rate much lower
than the normal operating discharge of the system
In-line and outlet pipe valves should be opened and closed
slowly (some gate valves are geared down to require many
turns of a wheel for opening and closing, and they cannot be
slammed shut)
To help avoid water hammer, the mean pipe velocity should usually not exceed
1.5 to 2.0 m/s (unless it is an open pipeline)
This is also an economic consideration, because pumping costs will increase
significantly at higher flow velocities, as the head loss due to friction is
proportional to the square of the velocity
The use of air chambers can help control intermittently high pressures
The use of stand pipes at the discharge side of a pump can protect the pipeline
from high pressures during operation and during filling of the line
The stand pipe should be of sufficient diameter that it can store water during
temporary pressure increases without quickly overflowing (this is only an option
for low pressure pipelines)
In some cases, air is continuously injected into a pipeline to dampen the effects
of water hammer, effectively reducing the wave speed and causing a high
pressure wave to more quickly decrease in amplitude
I. Introduction
1. direct rainfall
2. runoff from uphill areas
3. overflowing streams & rivers
4. irrigation
5. intentional leaching
1. Surface drains
2. Subsurface drains
Subsurface drains can be tile drains, concrete pipe, perforated plastic pipes,
mole drains, and others
Tile drains have gaps or openings between pipe segments to allow water to
enter from the surrounding soil
4KH
L2 = ( 2de + H) (1)
V
where L is the spacing between parallel drains; K is the saturated hydraulic
conductivity in the horizontal direction (mm/day); H is the vertical distance from the
shallowest depth to the water table to the drain depth; de is the equivalent depth to
the impermeable layer; and V is the drainage coefficient, or steady-state recharge
rate (same units as K), as defined below in Eq. 4
d
de = (2)
8d 8d
1+ ln
L 3r
where r is the outside radius of the drain pipe, or outside radius of the drain
envelope (if present); and d is the depth from the drains to the impermeable layer
(all length dimensions should be the same, usually m or ft)
The depth to the impermeable layer defines the value of d in Eq. 2, where an
impermeable layer can be taken as any soil layer whose hydraulic conductivity is
between one-fifth and one-tenth of the average hydraulic conductivity of the
overlying layer(s)
Thus, for the purposes of applying Eq. 1, the barrier is often a relatively
impermeable layer through which water does pass, but very slowly
To identify the barrier, and to obtain a representative value of hydraulic
conductivity, K, for all soil layers, use Eq. 3
ETmax (1 + LF )
V= ETmax (4)
Eirrig
where LF (the leaching fraction required for salinity control) can be taken as 0.05,
unless rain during the year will refill the soil profile and cause leaching of salts, in
which case LF is zero; and ET is the rate of crop consumptive use, or
evapotranspiration (mm/day)
deep percolation
Eirrig = 1 (5)
infiltration
Thus, if the deep percolation (infiltrated water passing downward through the root
zone) is zero, the efficiency is 1.0, or 100%
Eq. 5 does not account for surface runoff because the design is for subsurface
drainage
Note that this is only one of many definitions of irrigation efficiency, but it is useful
for determining the drainage coefficient
Both Eqs. 1 and 2 have the term L, which is the drain spacing
The two equations must be solved simultaneously (iteratively) to determine the
value of L for a given drain design, and this can be accomplished readily on a
programmable calculator, in a spreadsheet, or with a simple computer program
In summary, to apply Eq. 1 for the calculation of drain spacing the following seven
parameters must be known or estimated: K, H, d, ET, LF, r, and Eirrig
1
Q= AR2 / 3 So (6)
n
where Q is the flow rate (m3/s); n is a roughness coefficient; A is the cross-
sectional area of flow (m2); R is the hydraulic radius (m); and So is the slope of the
pipe (m/m)
Apply Eq. 6 at the downstream end of drain lines (note that Q varies with distance
along the drain lines)
The Manning roughness value is usually between 0.016 and 0.020 for corrugated
plastic drain pipes
The hydraulic radius is equal to the cross-sectional area of flow, A, divided by the
wetted perimeter, Wp
For a circular cross section, the geometry parameters are related as follows:
D2
A= ( sin ) (7)
8
D
Wp = (8)
2
and,
D
h= 1 cos (9)
2 2
where the terms in Eqs. 7 - 9 are defined in the following figure
2h
= 2cos1 1 (10)
D
Note that when h D, the hydraulic condition changes from open-channel flow to
pressurized pipe flow
Some sample applications of the drain spacing equations are given below
These data can be used to verify a computer program that applies Eqs. 1 - 5
Suppose subsurface drains are to be installed at a depth, z, of 1.2 m from the
ground surface, with an outside drain radius of 0.1 m and a minimum water table
depth of 0.6 m
The depth, d, to a relatively impermeable boundary is 1.5 m, the maximum ET rate
is 5 mm/day, the leaching fraction is 0.05, the irrigation efficiency is 75%, and the
hydraulic conductivity of the soil is 10 mm/day
For these data, the steady-state drain spacing would be about 4.99 m
If the depth to the impermeable layer were increased from 1.5 to 5.0 m, the
spacing would decrease to 4.74 m, or if the drain depth were increased from 1.2 to
1.5 m, the spacing would increase to 6.85 m
If the hydraulic conductivity were 15 mm/day, the spacing would increase to 6.37
m
I. Introduction
In the western USA there are two basic types of water rights that have been
established over the past 150 years, and these persist to the present time
The earlier of the two is known as a riparian right, which governs the private use
of water on riparian lands, which are properties adjacent to (bordering) a
natural stream or lake
Riparian water rights are usually considered to be attached to the land and the
land owner can use the water whenever desired
However, in some areas it is considered illegal for a riparian water user to store
water on the property for more than 30 days (typically), or to import water from
a hydrologic basin that does not border the property
Riparian rights cannot be transferred to other areas - they stay with the land
The other basic type of water right is the appropriative right, which is acquired
either by prior appropriation (first use) of a water resource or by some type of
permit
Unlike riparian rights, appropriative rights can exist without any particular
relationship between the proximity of land and the water resource
Also, holders of appropriative rights are generally allowed to store water on their
property for an indefinite duration of time
In the case of a prior appropriation, the user makes a claim on the water with
the argument that it was not previously being used by anyone else
Prior appropriations were common in mining areas of the western USA during
the 1800s
Appropriation by permit is more common today because the available water
resources are becoming more fully exploited (some argue that they already are
fully exploited)
Washington North
Dakota
Montana
South
Oregon Dakota
Idaho
Wyoming
Nebraska
California
Oklahoma
Arizona New Mexico
Texas
Riparian rights are usually older and have precedence over appropriative rights,
but appropriative rights actually control a greater volume of water in the western
USA
Some holders of riparian water rights do not make use of the available water,
creating a dormant riparian right, but such dormancy does not necessarily
imply forfeiture of the right, even if many decades have passed
Those with appropriative rights on water with dormant riparian rights always run
the risk of the riparian rights being reasserted, and this has occurred in more
than one case
Riparian-type water laws are more predominant east of the Mississippi river,
and appropriative water rights are more common west of the Mississippi
Riparian and appropriative rights tend to be conflicting and countless legal
disputes have arisen between the holders or claimants of such rights
As a result, there have been many court cases and much adjudication of water
resources
Water law is now an area of specialization for many lawyers in the western
USA, and the number legal battles over water rights is not likely to diminish in
the foreseeable future
Water rights are generally interpreted as usufructuary, which means that people
can have rights to use water, but not to own it
Furthermore, many legal definitions now require that all water use, regardless of
the type of right, must be reasonable and beneficial
This implies that water waste should be avoided and that water quality should
be conserved, but the actual legal definition of reasonable and beneficial is not
explicitly defined by current laws
Thus, this is one of a number of points that are subject to broad interpretation
by individuals and by judicial courts (some people believe that in water disputes,
it is not important who is right, rather, who has the best lawyer)
Some foresee an increasing emphasis on public water rights through the
doctrine of public trust, which asserts that specified resources (such as water)
are the property of the general public
I. Introduction
For example, the code of ethics of the American Society of Civil Engineers
(ASCE) contains the following seven canons:
1. Engineers shall hold paramount the safety, health, and welfare of the
public in the performance of their professional duties
2. Engineers shall perform services only in areas of their competence
3. Engineers shall issue public statements only in an objective and truthful
manner
4. Engineers shall act in professional matters for each employer or client as
faithful agents or trustees and shall avoid conflicts of interest
5. Engineers shall build their professional reputation on the merit of their
service and shall not compete unfairly with others
6. Engineers shall act in such a manner as to uphold and enhance the
honor, integrity, and dignity of the engineering profession
7. Engineers shall continue their professional development (continuing
education) throughout their careers and shall provide opportunities for
the professional development of those engineers under their supervision
In the state of Utah, registered professional engineers must now have 24 hours
of continuing education per two-year period to retain their registration as a PE
The ethical conduct of an engineer should be based on the protection & benefit
to the following groups and individuals, ranked below from highest to lowest
priority:
1. society
2. the law
3. the engineering profession in general
4. the engineers client
5. the engineers firm
6. other engineers
7. the engineer