Sei sulla pagina 1di 49

COLLEGE OF ENGINEERING, DESIGN, ART AND

TECHNOLOGY

SCHOOL OF ENGINEERING

DEPARTMENT OF CIVIL AND ENVIRONMENTAL ENGINEERING

YEAR THREE SEMESTER I 2011-2012

HYDROLOGY I (CIV3104)

HANDOUT 3: EVAPOTRANSPIRATION
7. EVAPORATION/EVAPOTRANSPIRATION
7.1 Introduction
The term evapotranspiration describes all the processes by which liquid water at or near the
land surface becomes atmospheric water vapor. It is all impossible in practice to separate
evaporation (from wet surfaces) and transpiration (water evaporation inside the plants) from
each other. Hence we focus on the combined quantity, evapotranspiration. Evapotranspiration is
a significant water loss from a watershed. Looking at the Global averages, two-thirds (about
62%) of the precipitation that fall on the continents is evapotranspired. Of this amount, 97% is et
from land surfaces and 3% is open-water evaporation. It is through the process of
evapotranspiration that the sun’s energy is introduced to drive the hydrological cycle.

The two ingredients for the phase change from liquid or solid water to water vapour are energy
and water. Hence, evapotranspiration is where the surface-water balance and surface energy
balance meet. Because both solar energy and available water are necessary to cause
evaporation (and transpiration), energy will limit the rate of evapotranspiration at some times
and water availability, will limit the rate at other times.

Most plants have openings (stomata) on their leaves to allow them to take up carbondioxide
from the atmosphere. When the stomata are open, plants transpire water. Unlike evaporation,
transpiration is not controlled solely by physical conditions because plants regulate the rate at
which water is released in transpiration in a manner that varies by plant type. Of the water taken
up by the plants roots, about 95% is transpired through the stomata. The remaining 5% or so is
converted to biomass through photosynthesis. Hence, to first order, the water taken up by the
roots is converted to vapour and lost to the atmosphere. When soil water is in limited availability,
the plants conserve it by restricting flow to the atmosphere through contraction of the stomata.
However, the degree of restriction varies considerably across plant species and even
throughout the year for a given specie. This renders quantitative treatment possible only in an
average sense, say, with an entire stand of trees treated as a single “big leaf” for which the
resistance to water flow is handled as a simple function of soil moisture content and time of the
year. Here, hydrology and ecology must be explored simultaneously, to incorporate a
quantitative description of plant use characteristics into the framework of hydrology.

Types of vegetation and land use significantly affect evapotranspiration, and therefore the
amount of water leaving a watershed. Because water transpired through leaves comes from the
roots, plants with deep reaching roots can more constantly transpire water. Likewise, the leaf

1
structure influences the rates of transpiration. For example, Conifer forests tend to have much
higher rates of evapotranspiration than deciduous forests. This is because their needles give
them superior surface area, resulting in more pores for transpiration, and allowing for more
droplets of rain to be suspended in and around the needles and branches, where some of the
droplets can then be evaporated. Through evapotranspiration, forests reduce water yield,
except for in unique ecosystems called cloud forests. Trees in cloud forests condense fog or low
clouds into liquid water on their surface, which drips down to the ground. These trees still
contribute to evapotranspiration, but often condense more water than they evaporate or
transpire.

In areas that are not irrigated, actual evapotranspiration is usually no greater than precipitation,
with some buffer in time depending on the soil's ability to hold water. It will usually be less
because some water will be lost due to percolation or surface runoff. An exception is areas with
high water tables, where capillary action can cause water from the groundwater to rise through
the soil matrix to the surface. If potential evapotranspiration (PET) is greater than actual
precipitation, then soil will dry out, unless irrigation is used.

Evapotranspiration can never be greater than PET, but can be lower if there is not enough water
to be evaporated or plants are unable to readily transpire.
Potential evapotranspiration (PET) is the amount of water that could be evaporated and
transpired if there was sufficient water available. This demand incorporates the energy available
for evaporation and the ability of the lower atmosphere to transport evaporated moisture away
from the land surface. PET is higher in areas closer to the equator, because of the higher levels
of solar radiation that provides the energy for evaporation. PET is also higher on windy days
because the evaporated moisture can be quickly moved from the ground of plants, allowing
more evaporation to fill its place.

Potential evapotranspiration is usually measured indirectly, from other climatic factors, but also
depends on the surface type, such as free water (for lakes and oceans), the soil type for bare
soil, and the vegetation. Often a value for the potential evapotranspiration is calculated at a
nearby climate station on a reference surface, conventionally short grass. This value is called
the reference evapotranspiration, and can be converted to a potential evapotranspiration by
multiplying with a surface coefficient. In agriculture, this is called a crop coefficient. The

2
difference between potential evapotranspiration and precipitation is used in irrigation
scheduling.

Average annual PET is often compared to average annual precipitation, P. The ratio of the two,
P/PET, is the aridity index.

7.2 Factors Affecting Evaporation


The physical process in the change of state from liquid to vapour operates in both Eo and Et and
thus the general physical conditions influencing evaporation rates are common to both. The
plant's growth stage or level of maturity, percentage of soil cover, solar radiation, humidity,
temperature, and wind are some of the main factors that affect evapotranspiration as highlighted
below;
(a) Solar Radiation
Latent heat is required to change a liquid into its gaseous form and in nature this is provided by
energy from the Sun. The latent heat of vaporization comes from solar (short-wave) and
terrestrial (long-wave) radiation. The incoming solar radiation is the dominant source of heat
and affects evaporation amounts over the surface of the Earth according to latitude and season.
(b) Temperature
Temperature of both air and the evaporating surface is important and is also dependent on the
major energy source, the Sun. The higher the air temperature, the more water vapour it can
hold, and similarly, if the temperature of the evaporating water is high, it can more readily
vaporize. Thus evaporation amounts are high in tropical climates and tend to be low in Polar
Regions. Similar contrasts are found between summer and winter evaporation quantities in
mid-latitudes.
(c) Saturation Deficit
Directly related to temperature is the water vapour capacity of the air. A measure of the amount
of water vapour in the air is given by the vapour pressure, and a unique relationship exists
between the saturated vapour pressure and the air temperature that is the saturated vapour
pressure decreases with decreasing temperature. Evaporation is dependent on the saturation
deficit of the air, which is the amount of water vapour that can be taken up by air before it
becomes saturated. Hence more evaporation occurs in inland areas where the air tends to be
drier than in coastal regions with damp air from the sea.

3
(d) Wind Speed
As water evaporates, the air above the evaporating surface gradually becomes more humid until
it is saturated and can hold no more vapour. If the air is moving, however, the amount of
evaporation is increased as drier air replaces the humid air. Thus wind speed at the surface is
an important factor. Evaporation is greater in exposed areas that enjoy plenty of air movement
than in sheltered localities where air tends to stagnate.
It will be noted that the temperature and wind speed factors may be in conflict in affecting
evaporation since windy areas tend to be cooler and sheltered areas are often warmer. Over a
large catchment area, it is the general characteristics of the prevailing air mass that will have the
major effect on evaporation (apart from the direct solar radiation).
The principle influences on the physical process of evaporation enumerated above are in their
turn affected by wider considerations. The following factors outline more generally larger-scale
influences.

(e) Weather Pattern


The prevailing weather pattern indicated by the atmospheric pressure affects evaporation. The
edge of an anticyclone provides ideal conditions for evaporation as long as some air movement
is operating in conjunction with the high air pressure. Low atmospheric pressure usually has
associated with it damp unsettled weather in which the air is already well charged with water
vapour and conditions are not conducive to aid evaporation.
(f) Nature of surface
The nature of the evaporating surface affects evaporation by modifying the wind pattern. Over a
rough, irregular surface, friction reduces wind speed but has a tendency to cause turbulence so,
with an induced vertical component in the wind, evaporation is enhanced. Over an open water
surface, strong winds cause waves which provide an increased surface for evaporation in
addition to causing turbulence. As wind passes over a smooth, even surface there is little friction
and the evaporation is affected predominantly by the horizontal velocity.
(g) Salinity of the water
Evaporation may also be affected by the salinity of the water or by the presence of pollutants
such as oil films, but these are of relatively minor importance.

Variations in some of the dominant factors operating over different surfaces can result in
noticeable changes in evaporation rates over small adjacent areas in short time periods.
Diurnal fluctuations are considerable since during the night there is no solar radiation. However,

4
evaporation totals over neighboring areas show relatively smaller differences over periods of a
week or a month.
Evaporation is necessarily dependent on a supply of water and thus the availability of moisture
is a crucial factor. With all the other factors acting favorably, once the body of water disappears
then open water evaporation Eo ceases. For Et, the availability of water is not so easily
observed. Plants draw their supply from the soil where the moisture is held under tension, and
their rate of transpiration is governed by the stomata in the leaves which act like valves to
regulate the passage of water through the pores according to the incidence of light. The pores
are closed in darkness and hence transpiration ceases at night. When there is a shortage of
water in the soil, the stomata regulate the pores and reduce transpiration. Thus Et is controlled
by soil moisture content and the capacity of the plants to transpire, which are conditioned by the
meteorological factors.

7.3 Measurement of Evaporation

7.3.1 Continuity Methods

i) Water Budget Approach


Evapotranspiration may be estimated with a water balance approach, if the change in storage
and all the inputs and outputs except et are known. (See equation 1.3 in Handout1). The water-
budget method is the simplest of the analytical methods and is also the least reliable.

= p (t ) + rsi (t ) + rgi (t ) _ rso (t ) _ rgo (t ) _ et (t )


dV
7.0
dt
In this equation the time rate of change in this volume of water stored within the control volume
(left side) is balanced by the difference between the inputs and the outputs. Each of the terms
on the right side of the equation has units of volume per time [L3T-1], and all are functions of
time. Equation (7.0) may be simplified for some problems, for example by neglecting rgi and rgo
when considered unimportant. Writing equation (7.0) for average annual quantities reduces it to
equation (1.3). Once all the terms except et have been all measured, then et may be computed
as the residual to force a balance that conserves mass in the system. In many cases however
the water balance approach may suffer from the accummulation of errors in the measured
variables. As an example, consider a lake with a very rapid throughflow rate, which over the
course of the year does not see a net change in storage. Suppose that the measured average
annual fluxes and associated errors for the lake are:

5
p = 10 7 ± 5 × 10 5 m 3 yr −1 ( ± 5% )

rsi = 10 9 ± 1.5 × 108 m 3 yr −1 ( ± 15% )

rso = 9.95 x108 ± 1.5 × 108 m 3 yr −1 ( ± 15% )

If we neglect ground water inflows and outflows, we can use these values and equation 7.0 to
solve for et. The result, accumulating the errors as we go, is 1.5 x10 7 ± 3 × 108 m 3 yr −1 (Bear in
mind that for a lake such as this, the et is a relatively small part of the budget)
It is unrealistic to expect to be able to quantify accurately, all the terms in a water balance for a
catchment to solve for et, especially over short time periods when storage changes are both
significant and very difficult to measure or predict. In view of the various uncertainties in the
estimated values and the possible errors in measured variables, the water-budget method
cannot be expected to give very accurate results. Further more, this is a diagnostic rather than a
predictive approach. In many cases we need to predict that rate of water loss to the atmosphere
for certain anticipated conditions.

ii) Field Site


On a small plot scale, it may be feasible to provide sufficient soil moisture instrumentation to
determine ∆S (Change in water storage) in addition to P (daily precipitation), Vis (daily surface
inflow), Vos (daily surface outflow) and hence estimate E. The main difficulty is to account for
possible lateral sub-surface flows and inflows (capillary rise) and outflows at the base of the
measured profile. If tensiometers are used, a ‘zero flux plane may be identified, above which
water is lost to evaporation and below which water passes to percolation and drainage.

iii) Lysimeters (Evapotranspirometer)


A lysimeter is an artificially enclosed volume of soil for which the inflows and outflows of liquid
water can be measured and usually, changes in storage can be monitored e.g by weighing. A
lysimeter can be used to measure the quantity and quality of percolated water, or monitor the
water budget, from an isolated block of soil in the field.
Basically, it is a soil tank with growing plants inserted into the ground to simulate natural
conditions (Figs. 7.1). Rain or irrigation water percolated through the soil profile is collected at
the bottom of the tank or pumped to the surface through installed pipes for analysis. The
amount of evaporation is measured by the difference between the precipitation and the change
in soil moisture. It is necessary to ensure that the thermal, hydrological and mechanical
properties of the sample soil are similar to those of the parent material.

6
Rainfall input is measured, drainage is collected, and soil moisture storage determined through
gravimetric (weighing lysimeters), volumetric (floating lysimeters), or water-budget (percolation
lysimeters) approaches. Although the weighing lysimeters are the most complicated and
expensive, they are generally considered the most accurate means of direct measurement of
actual evapotranspiration (Dugas et al. 1985). Their high fabrication costs may limit the number
of replicates available for statistical evaluation.

Lysimeters have been widely used in research for testing hypotheses on plant-water
relationships and calibrating theoretical and empirical evapotranspiration models. Their
accuracy depends on the sensitivity to and representativeness of the surrounding areas.

Fig 7.1a An evapotranspirometer (or drainage lysimeter) Fig 7.1b A Lysimeter is covered with
grasses
7.3.2 Energy Balance Methods
The energy balance method is an application of the law of conservation of energy. The energy
available for evaporation is determined by considering the income energy, outgoing energy and
energy stored in the water body over a known time interval.

According to the first law of thermodynamics, the net radiant energy received at the land surface
must be conserved. For the conservation of mass, we consider a control volume and seek to
balance inflows with outflows and net changes within the system. For convenience, we establish
an imaginary control volume near the land surface that includes a very thin layer of top-soil (say,
10-mm thick), the vegetation, and the immediate surrounding air (Fig. 7.2(a)). Thermodynamic
principles hold that the net radiant energy arriving across the boundary of this system must be
exactly balanced by other energy fluxes across the boundary and the net change in energy held

7
within the volume. The energy may form among its possible forms (radiant, thermal, kinetic and
potential), but it must be conserved.

Fig 7.2 (a) A Control volume for energy conservation. Solar Radiation (Rn)
entering the control volume must be balanced by fluxes of energy out of the

volume and by the time rate of change in energy stored ( dQ )


dt

All matter has internal energy Eu (ML2T-2) due to the kinetic and potential energy associated
with individual molecules. Internal energy is an extensive property (it depends on the amount of
material) and is expressed in units of calories or joules. There are several types of internal
energy. Here, we concern our selves with sensible and latent heat (energy).

1. Sensible heat is the portion of the internal energy that is proportional to temperature. As
the name implies, this is the heat that you would “sense” by contact or touch.
The specific heat capacity Cp ( L2 Θ −1T −2 ) provides a measure of how a substance’s
internal energy changes with temperature:

dE u / m
Cp = .......................(7.a )
dT

8
Where m is the mass and T (K) is the absolute temperature (Note that absolute temperature
is represented in degree Kelvin). Basically, the specific heat capacity is a measure of the
amount of heat required to raise the temperature of a unit mass by 1 oC. We can use the
specific heat capacity to determine how the temperature of a given mass of water will
change if we add energy (heat the water). The heat capacity of water at 20 oC is
approximately 4.2 x10 3 JKg −1 K −1 or 1.0 ca g-1K-1. If we have 1.0 Kg of water and add 12kJ
(kilojoules) of energy, the temperature change can be calculated as:
dE u / m 12000 j / 1 Kg
dT = = = 2 .9 K
Cp 4.2 x10 3 JKg −1 K −1
Or the water temperature will increase by approximately 3 oC. We have assumed in this problem
that we have not evaporated any of the water.

2. Latent Heat is the portion of the internal energy that cannot be sensed or felt. Instead,
the latent energy is the amount of internal energy that is released or absorbed during a
phase change, at a constant temperature. Evaporation involves a liquid or vapour
conversion, which requires energy to be added to the water. This energy is called the
latent heat of vaporisation:
λv = 2.45 x10 6 JKg −1 ( at 20 o C ) ........................(7.b )
This tells us that we need to add about 2.5 million joules of energy to evaporate 1 Kg of water.
Other types of latent heat energy include:
(i) Latent heat of melting: λm = 3.34 x10 5 JKg −1 (at 0 o C )

(ii)Latent heat of fusion: λ f = −3.34 x10 5 JKg −1 ( at 0o C )

(iii) Latent heat of sublimation: λ s = 2.83 x10 6 JKg −1 ( at 0o C )


Note that evaporation of ice or snow involves a change in phase from solid to vapour. This may
occur in either two separate steps, melting and then evaporation, or in one step as sublimation,
which is the direct phase change from ice to water vapour. In either case, latent heat is added to
support the sum of the two unique phase changes, as evidenced by the fact that λ s = λv + λ m .

The received radiation (Rn) warms the exposed surfaces inside the system. When water is
present, some of this heat energy is absorbed by the water to support the phase change from
liquid to vapour (evapotranspiration). Evapotranspiration will not typically absorb all the energy,
and so the surface continues to warm. With the surface becoming warmer than the air and the
underlying soil, heat will be conducted from the hot surface to the air (H), and from the surface

9
down into the soil (G). During this process, both heat and water vapour have been added to the
air inside the control volume. The moist, warm air becomes less dense than the surrounding air
and tends to rise up and out of the control volume. Therefore, we see that the main energy
outlets of the sun’s energy that heats the Earth’s surface are conduction into the soil, conversion
of liquid water to water vapour, and heating of the overlying air. The energy used to evaporate
the water is stored in the water vapour, and is removed from the system as the water vapour is
removed. Finally, the latent heat (evaporating energy) is converted back to thermal energy when
and where the water recondenses.

The energy balance approach to evaporation is based on the fact that evaporation involves an
energy flux (i.e of latent heat energy to the atmosphere), and therefore the rate of evaporation
can be described in the context of an energy balance equation (Figure 7.2(a)).

dQ
= Rn − G − H − E l ………………………………………….. (7.c)
dt

Where: Rn = the net (solar) radiation input; G = energy output through conduction to the ground;

H = net output of sensible heat to the atmosphere; El = output of latent heat to the atmosphere
(the latent heat flux), due to evaporation; Q= the amount of heat energy stored in the control
volume per unit area of surface. Each of the inputs and outputs of the right side of equation 7.c,
is in the form of an energy flux (i.e energy per unit area per unit time). In SI units, this would be
Jm-2s-1 or W m-2. The convention adopted here is to express Rn and G as positive when directed
downward and H and El as positive when directed upward. We can re-arrange equation 7.c to
solve for El , the latent heat flux:

dQ
E l = Rn − G − H − ........................................(7.d )
dt

The latent heat flux is related to the rate of evapotranspiration through the latent heat of
vaporization:

El
et = .......................................................(7.e)
ρ w λv
Where et = the evapotranspiration rate [LT-1]; ρ w =density of water [ML-3]; λv = latent heat of

vaporization at the temperature of interest [L2T-2]. We can substitute equation equation (7.e) into
equation (7.d) and re-arrange to solve for the evapotranspiration rate:

10
Rn − G − H − dQ / dt
et = ......................................( 7. f )
ρ w λv
Let us consider a simple example of how we might apply the energy balance approach to
calculate the daily evaporation from a forest on a sunny day (Rn= 200 W m-2). If we can neglect
G, H and assume that the temperature (thermal energy content, Q) within the forest remains
approximately constant, then equation (7.f) becomes:
Rn 200W m−2
et = = = 8.0x10−8 ms−1 = 0.7 cmday−1
ρ wλv (1000Kg m )(2.5x10 JKg )
−3 6 −1

However, in practice we cannot neglect H. Also, for averaging times significantly less than a day
we cannot safely neglect G either. We need to consider the state (wetness) of the surface to
understand and quantify how the received energy is partitioned.

When water is in limited supply, the surface becomes warmer than in the wet cases and more of
the energy is removed from the control volume through conduction into the soil and heating of
the air. Hence, in this case, the surface properties rather than the atmospheric properties are
controlling the rate of evapotranspiration. Under wet surface conditions, the rate of
evapotranspiration is governed by the supply of radiant energy, the relative dryness of the air
and the efficiency of the wind in removing the water vapour from the surface. This is evidenced
by how quickly water evaporates from wet clothes on a dry day, as compared to a humid day.
We know intuitively that for the same example, higher winds will increase the evaporation rate,
and reduced solar input, say from heavy cloud cover, will reduce the evaporation rate.

The rate of et that occurs under the prevailing solar inputs and atmospheric properties, if the
surface is fully wet, is commonly referred to as potential evapotranspiration (PET). As earlier
discussed (section 7.1), for a catchment water balance, we are interested in the actual et,
because that is the rate at which water is removed. The concept PET is useful as a tool if we
have some means to estimate actual et from knowledge of the potential rate as well as the
nature of the surface conditions. By definition, when the surface is wet, the ratio of et to PET will
be unity. Conversely, when the surface is completely dry, the ratio will go to zero.

Following from equation (7.d), a convenient approach is to replace the sensible heat flux term
with an expression of the ratio of sensible to latent heat flux, called the Bowen ratio (B=H/El).
This effectively removes H from the energy balance equation.

dQ
E l + H = E l (1 + B ) = Rn − G − ……………….(7.g)
dt

11
Which allows us to calculate the rate of evapotranspiration:

Rn − G − dQ / dt
et = ......................................( 7.h )
ρ w λv (1 + B )
The Bowen ratio is a function of the gradients of temperature and water vapour pressure in the
air above the surface. We can estimate these gradients from differences between the surface
values (Ts, es) and air values at some fixed height above the surface (Ta, ea). The sensible and
latent heat fluxes are proportional to these differences, such that the Bowen ratio is:

Ts − Ta
Bα ...........................................(7.i )
e s − ea
So we see that as a surface becomes warm, (Ts increases) and dry (es decreases), the Bowen
ratio tends to increase. Consequently, the sensible heat flux increases relative to the latent heat
flux. The Bowen ratio ranges from approximately 0.1 in very humid regions (B=0.1 over the
oceans) to values greater than 1 in arid regions. In deserts, B may be greater than 3 (Chang
2006). With estimates or measures of Rn , G , dQ / dt and the Bowen ratio, equation (7.i) provides

a simple and direct estimate of et.

7.3.3 Aerodynamic (Mass Transfer) Methods


a) Diffusion Methods
In this case evaporation can be estimated from the measurements of temperature, the humidity,
and the velocity at two points within the turbulent boundary layer together with the height of the
effective datum.

b) Eddy Correlation
This method relies on the analysis of the structure of turbulence to determine transport rates.
Although mean wind velocity will be parallel to the surface, there will be instantaneous
perpendicular components. Similarly, there will be fluctuations in air temperature and humidity
about mean values at a given height. Hence if fluctuations of velocity and e.g. humidity are
simultaneously measured, the net flux can be calculated.

12
7.3.4 Combination Methods

A difficulty with the aerodynamic and energy budget methods is that humidity measurements at
2 levels on a short time-scale are required, and these are difficult measurements to make.
A combination approach can be used in which the aerodynamic method is used to determine H
(using temperature measurements, not humidity) the energy balance is then used to determine
λ E. This overcomes the need for humidity measurements, but still includes the other
disadvantages of the aerodynamic method.

7.4 Estimation of Evapotranspiration


Currently, there are no widely used methods for estimating areal distribution of actual et,
although satellite-derived surface data offer promise of delivering useful estimates of
components of the surface energy balance from which actual et might be inferred (Moran and
Jackson 1991). In hydrology, computations using empirical evapotranspiration equations
typically are used to estimate actual et. For a detailed discussion of techniques to measure and
estimate et, see Brutsaert (1982) and Raghunath (1981). Some of the methods of estimating
evapotranspiration are;
(i) Tanks and Lysimeter experiments
(ii) Field experimental Plots
(iii) Installation of sunken (Colarado) pans
(iv) Evapotranspiration equations as developed by Lowry-Johnson, Penman, Penman Monteith
Combination Equation, Thornwaite, Blaney-Criddle etc.
(v) Evaporation Index Method, i.e from pan evaporation data as developed by Hargreaves and
Christiansen

7.4.1 Evaporation Equations

7.4.1.1 The Penman Combination Equation

Penman (1948) developed a formula for calculating open water evaporation based on
fundamental physical principles, with some empirical concepts incorporated, to enable standard
meteorological observations to be used. He began considering evaporation from open water
and gradually refined his formulae over three decades in order to better represent processes of
transpiration. Refinements in the representation of transpiration and of evaporation from
intercepted water have also been added by Monteith and by the British Meteorological Office. In
fact, the great strengths of Penman’s approach have been its adaptability and its pragmatic use
of available data. Its world-wide popularity is largely due to Penman’s provision of alternative

13
methods for calculating the essential parameters, based on his appreciation of lack of
instrumentation in many parts of the World.
In essence, penman combined the energy budget and mass transfer approaches. He used the
energy budget method to calculate the energy available for evaporation, but he simplified it by
ignoring the effects of storage and advection. The basic equations are modified and rearranged
to use meteorological constants and measurements or variables made regularly at
climatologically stations.

In a simplified energy balance equation:

H = Eo + Q (7.7)

Where H is the available heat, Eo is energy for evaporation and Q is energy for heating the air.

The values of Eo and Q can be defined by the aerodynamic equations Equation 7.10.

According to Dalton (1802), evaporation is estimated by;


E = f (u )(e sat − e act ) Where; f (u ) = Function of wind

e sat = Saturated vapour pressure at temperature

e act = actual vapour pressure

Therefore from eqn (7.7) ;


E 0 = f (u )(e s − e d ) (Dalton-1802) (7.8)

and

Q = γ f1 (u) (Ts – Ta) (7.9)


γ is the psychrometric constant (0.27 mm of mercury/oF) to keep units consistent. (also see
equation 7.23). It is generally assumed that f(u) = f1 (u). If the aerodynamic equation (Equation
7.8) is based on the air humidity using the air temperature Ta, then:

E a = f (u ) (e s − e a ) (7.10)

14
where ea is the vapour pressure at air temperature Ta, and thus (es – ea) is the saturation deficit
The temperature, Ta, is easily measured, whence ea is easily obtained, whereas es is difficult to
evaluate.

If ∆ represents the slope of the curve of saturated vapour pressure plotted against temperature,
then:
de e s − ed ea − ed
∆= ≈ ≈ (if gradients are small)
dT Ts − Td Ta − Td

Then from Equation 7.9:

Q = γ f(u) [(T s − T d ) − (T a − Td )]

 e s − ed e a − e d 
= γ f(u)  −
 ∆ ∆ 

γE o γE a
= − (7.11)
∆ ∆

Substituting for Q in the energy balance equation (Equation 7.7)

γE o γE a
Eo = H - +
∆ ∆

∆ Eo + γEo = ∆H + γEa

∆H γE a
Eo = +
∆+γ ∆+γ
∆H + γE a
Eo=
∆ +γ
Dividing thru by γ

15
∆ 
 H + E a 
Eo=
γ  (7.12)
∆ 
 + 1
γ 

This final equation is the basic Penman formula for open water evaporation. It requires values
of H and Ea as well as ∆ for its application.
If net radiation measurements are available, then H, the available heat may be obtained directly
in mm water equivalent. More often, H is calculated from incoming (RI) and outgoing (Ro)
radiation determined from sunshine records, temperature and humidity, using:

H = R1 (1-r) - Ro (7.13)

Where r is the albedo and equals 0.05 for water. R1 is a function of Ra, the solar radiation (fixed
by latitude and season) modulated by a function of the ratio, n/N, of measured to maximum
possible sunshine duration. Using r = 0.05 gives:
R1 (1-r) = 0.95 Rafa (n/N) (7.14)
Penman used fa (n/N) = 0.18 + 0.55n/N in the original work, but later studies have shown that
the function fa (n/N) depends on the clarity of the atmosphere and latitude (MAFF, 1967).
The term Ro in Equation 7.13 is given by:

(
Ro = σTa4 0.56 − 0.09 ed ) (0.10 + 0.90n / N ) (7.15)

Where σTa4 is the theoretical black body radiation at Ta which is then modified by functions of
the humidity of the air (ed) and the cloudiness (n/N). σ is Stefan-Boltzmann Constant (5.76 x
10-8 Wm-2K-4).

Next, Ea in Equation 7.16 is found using the coefficients derived by experiment for open water:
Ea = 0.35(0.5 + u2/100)(ea – ed) (7.16)
Finally a value of ∆ is found from the curve of saturated vapour pressure against temperature
corresponding to the air temperature, Ta.
The four measurements required to calculate the open water evaporation are thus:
 Ta mean air temperature for a week, 10 days or a month, oF or oC
 ed mean vapour pressure for the same period, mmHg
 n bright sunshine over the same period, h/day
 u2 mean wind speed at 2m above the surface, miles/day
 Ra and N are obtained from standard metrological tables (see Appendix).

16
Note:
With meteorological observations made in various units and the tendency to work now in SI
units, care is needed in converting measurements into the appropriate units for the formula.
The evaporation from open water Eo is finally in mm/day. The values obtained by the Penman
equation are comparable to the pan measurements, since it was derived to approximate
evaporation from a pan.

i) The East African Penman Formula


Specific developments of the empirical relations have been proposed for certain regions. For
example, in East Africa (McCulloch. 1965), equation 7.17 is presented:

Eo =

∆+γ
 
Ra (1 − r )  0.29Cosϕ + 0.52

n
N

 −

 ∆ + γ
 4 n
( )
σTa  0.10 + 0.90  0.56 − 0.08 e 
 N 
  
γ   h  u  
+ 0.261 + 1 + (es − e )
∆+γ   20000  100  

(7.17)

Where in addition to the previously defined terms,


Ra = maximum radiation on a clear day for a given latitude and time of the year
r = 0.05, for water
= 0.25 for green crops

h is the altitude in m, φ is the latitude.


e is in mbar and u in miles/day
A further modification by Woodhead suggests that the Incoming radiation term should be
n
calculated using Ra (0.23+0.53 ) for the same region.
N
In a study (Mangeni 2007) on the estimation of the evaporation of Lake Victoria, a combination
of the Penman (1948) and the energy balance yields an energy budget that relies on the
radiation balance and mass transfer terms over the lake, to make more realistic estimates of
evaporation over the lake. The Penman terms were evaluated according to Gieske and Yin and
Sharon energy balance and water and heat transfer expressions, using input data from island
stations and a few lakeshore stations. These resulted in annual evaporation of 1678 mm which
is considered more representative. This value is greater than the mean annual evaporation of
most recent estimates of 1537 mm per year, regarded as an underestimate.

ii) Calculation of Potential Evaporation

17
A value of the actual evaporation (Et) over a catchment can be obtained by first calculating the
potential evaporation plus transpiration (PE), i.e. assuming an unrestricted availability of water,
and then modifying the answer by accounting for the actual soil moisture content. This method
was developed by Penman and has been recommended for general use. There are several
formulae for calculating potential evaporation and their derivation has been largely due to the
need to assess irrigation demands. The Penman formula in particular, has served more widely
to help provide a numerical evaluation of the moisture content of a catchment and the
calculations of Et.
He related the potential evaporation from a vegetated land surface to the evaporation from open
water calculated:
PE = f.Eo (7.18)
where f is the seasonal conversion factor
He obtained the following values considered applicable in the climate of Western Europe:
Summer f = 0.8 (May to, August)
Winter f = 0.6 (November to February)
Equinoctial months f = 0.7 (March, April, September, October)

Thus, having calculated Eo for a particular month using Equation 7.12 and its defined
components, the potential evaporation is obtained by applying the appropriate factor, which is
locally applicable.
Penman gave three main reasons why the estimate of potential evapotranspiration cannot be
greater than open water surface evaporation;
a) The albedo of vegetation (10%-18%) is generally greater than that for water (2% to 7%),
hence less energy is absorbed and available for evapotranspiration than for evaporation
from the water surface.
b) At night when the leaf stomata are closed transpiration is virtually nil, whereas
evaporation may still occur from a water surface using the heat stored within.
c) Even when stomata are open during the day they offer resistance to the diffusion of
water vapour.

Note: Surface reflectivity to incoming radiation varies with the incident angle and surface
characteristics such as colour, soil moisture content and roughness. Below are examples of
surface reflectivity of different surfaces.
Water reflects 5% to 15%,

18
Snow reflects 70% to 90%.
Hardwoods 0.15 – 0.20
Shortgrass 0.15 – 0.25
Crops 0.15 – 0.25
Soils, Clay 0.20 – 0.35
Sandy 0.25 – 0.45

More recently (Finch 2003) noted that empirical factors applied in the Penman Potential
evaporation equation are likely to significantly overestimate open water evaporation and
proposes new empirical factors that have greater accuracy.

7.4.1.2 The Penman Monteith Combination Equation

This was a further development of the original Penman equation by (Monteith 1965) [and later
discussed by Monteith 1985 and Unsworth (1990)] in he incorporated biologically based canopy
and physically based aerodynamic resistance parameters into the Penman’s wind function for
estimating water vapour diffusion over vegetated surfaces. It represents the evaporating surface
as a single big leaf, with two parameters; one determined by the crop canopy architecture, while
the other depends on the biological behavior of the crop canopy surface and is related to both
crop specific parameters; light attenuation, leaf stomatal resistances and environmental factors.
The Penman-Monteith model is probably the most popular equation for estimating PE; however
the Penman-Monteith model requires climatic data that are not usually available e.g wind speed
and resistance parameters. This limits its applicability. The estimates from this method are
usually greater than pan evaporation, which can occur for short periods.
The FAO-56 Penman-Monteith equation has been adopted as the new standard method for
estimating reference crop evapotranspiration and is given as:

0.408∆ (Rn − G ) + U 2 (es − ea )γ


900
ETo = T + 273 (7.19)
∆ + γ (1 + 0.34U 2 )
where:
ETo= reference crop evapotranspiration (mm/day)
Rn= net radiation at the crop surface (MJ/m2/day)
G= soil heat flux density
T= mean daily air temperature at 2m height (0C)
U2= wind speed at 2m height (m/s)

19
es= saturation vapour pressure (kPa), given by

e (T max )+ e (T min )
es =
2 (7.20)
 17.27T 
e(T ) = 0.6108 exp  
 T + 237.3  (7.21)
where:
ea= actual vapour pressure
es-ea= saturation vapour pressure deficit (kPa), where:
ea = e(Tdew )
Tdew is the dewpoint temperature (for well-watered sites, Tdew can be approximated to the
minimum air temperature).
∆= slope vapour pressure curve (KPa/0C) given by:

  17.27T 
40980.6108 exp 
  T + 237.3 
∆= (7.22)
(T + 237.3)2
γ = Psychrometric constant (KPa/0C) given by:

cpP
γ = (7.23)
ελ v
where P is the atmospheric pressure (KPa/0C), cp is the specific heat of air at constant
pressure, ε is the ratio of the molecular weight of water vapour to that of dry air= 0.622, and
λv is the latent heat of vaporization. Thus, the psychrometric constant depends on atmospheric
pressure which also changes with altitude and varies slightly overtime in a given location. Using
typical values of;

cp = 0.24 cal g-1 0C-1


Atmospheric Pressure = 1.013MJkg-10C-1 = 1013mb
λv = 590 cal g-1, we find that
γ = 0.66 mb 0C- and this is the value commonly used.
However, the decrease of pressure with elevation should be accounted for when calculating γ
for applications at high elevations.

20
For Net longwave radiation (Rnl),

Rnl = δ 
2
(
 Tmax , K 4 + Tmin , K 4 
)
 0.34 − 0.14 ea 1.35
Rs
Rso

− 0.35  (7.24)
   
Where Rnl = net outgoing longwave radiation

In a study (Akinbile, 2007) tested five models on the suitability for determining the reference
evapotranspiration in the region in comparison to the Penman Monteith equation. These were
Jensen-Haise, Blaney- Morin-Nigeria, Thornthwaite, Modified Hargreaves and Penman. The
results showed that after Penman, the Modified Hargreaves gave the best covariabilty. This
suggests that in situations where there is very limited data, as in many parts of of sub-Saharan
Africa, the Modified Hargreaves can be used to obtain reasonable estimates of reference
evapotranspiration.

Example 7.1
Given the monthly average climatic data of April of a station at Atumatak in Moroto District,
Uganda, located at 20141N and at an elevation of 1260m, calculate the potential evaporation
using the Penman formula assuming a seasonal conversion factor of 0.7.
A. Data April
0
1. Mean air temperature 22.5 C
2. Mean relative humidity, % 60
3. Mean sunshine hours, n 12
4. Possible sunshine hours (N from Table 7.7) 12
5. Value of the ratio n/N 1
6. Wind speed at 2m height in miles/day (u2) = 172.8
7. Monthly average daily vapour pressure (ea) =
1.02kPa
8. Extra terrestial radiation in mm/day (Ra from App. Table
7.6) 37.13/2.45=15.155mm/day
9. Reflection coefficient (r) 0.05

B. Solving expression RI(1-r)


10. (1-r)=(1-0.05)= 0.95
11. (a+b.n/N)=(0.18+0. 55x1.0)= 0.73
12. RI(1-r) = Nos (8x10x11)= 10.510 mm/day
C. Solving expression δ Ta (0.56-0.092ed )(0.10+0.90n/N)
4 1/2

13. Vapour pressure i) ea(From Table 7.3)= 2.726kPa = 20.447mmHg


ii) ed (recorded)= 1.02kPa = 7.651mmHg
iii) ed1/2 = 2.766
14. σTa (= 5.67x 10 x (273+22.5)4 =
4 -8
-2
Ta must be converted to K 432.328 Wm

21
= 432.328 x (2.388 x 10-5 x 1440 min/day) 14.867mm/day
15. (0.56-0.09ed1/2)= (0.56-0.09x2.766)= 0.311
16. (0.10+0.90.n/N)= (0.10+0.90x1.0)= 1.0
17. Ro = (Nos. 14x15x16)= 4.624 mm/day
D. Solving for H = Item No. 12- item No. 17=(10.50-7.29)= 5.886mm/day
E. Solving for Ea=0.35(ea-ed)(0.5+u2/100)
=0.35(20.447-7.651)(0.5+172.8/100)= 9.978 mm/day
From Table 7.4
∆ = 0.165kPa/0C = 1.238mmHg/oC = 1.238/33.8 = 0.037mmHg/0F
∆/ γ = 0.037/0.27 = 0.137
0 . 137 x 5 . 886 + 9 . 978
Solving for E0 =
F. 1 + 0 . 137 9.485 mm/day
G. i) PET= Eo x K for April = 9.485x0.7 = 6.640 mm/day

When the seasonal conversion factor is not given, K can be taken as 1.0 so that PE = Eo
(Computed).

Actual Evapotranspiration
When water supply for vapourisation is deficient or soil moisture content is below the field
capacity, the vapourisation cannot proceed at the potential level. Thus, the actual
evapotranspiration (AE) is only a fraction of (PE) or AE = α (PE); Where α is affected not only by
soil moisture content but also by climate and species. It is the so called Crop coefficient in
agricultural Irrigation.

Example 7.2
Given the monthly average climatic data of April of a station at Atumatak in Moroto district
located at 20141N and at an elevation of 1260m, calculate the potential evaporation using FAO
Penman-Monteith:
Monthly average daily maximum temperature (Tmax) = 29.80C
Monthly average daily minimum temperature (Tmin) = 13.70C
Monthly average daily vapour pressure (ea) = 1.02kPa
Monthly average daily wind speed (u2) = 3.2 m/s
Monthly average daily sunshine duration (n) = 12hours/day
For April, Mean monthly average temperature (Tmonth, i) = 22.50C
For March, Mean average minimum temperature = (Tmonth, i-1) = 23.20C

22
Solution
i. Parameters
T max + T min 29 . 8 + 13 . 7
Tmean= = = 21.750C
2 2
Slope of vapour pressure-temperature curve ∆

  17.27T    17.27 x 21.75 


40980.6108 exp  40980.6108 exp 
  T + 237.3    21.75 + 237.3 
∆= = =0.16
(T + 237.3)2 (21.75 + 237.3)2
Altitude = 1260m

 293 − 0.0065H 
5.26

Atmospheric pressure P = 101.3 


 293

 293 − 0.0065 x1260 


5.26

= 101.3  = 87.3kPa
 293
cpP
Psychometric constant γ = = 0.665 X 10 −3 P = 0.665x10-3x87.3= 0.058 kPa/0C
ελ

ii. Vapour pressure deficit


 17.27T 
e s (T ) = 0.6108 exp  
 T + 237.3 
 17.27 x 29.8 
e s (Tmax = 29.8) = 0.6108 exp   = 4.20kPa
 29.8 + 237.3 

 17.27 x13.7 
e s (Tmin = 13.7 ) = 0.6108 exp   = 1.57kPa
13.7 + 237.3 
Given (ea) = 1.02kPa
es= (4.20+1.57)/2= 2.89kPa
Vapour pressure deficit (es-ea) = 1.87kPa

23
iii. Radiation
From Table 7.6 J = (for 15 April) 105
0 1 0
or 7.5 Latitude=2 14 N= (2+14/60)= 2.23 N
Ra= 37.13 MJm-2day-1
N (Table 7.7) Day length N= 12.0 hours
n/N=12/12= 1
Incoming Solar radiation Rs=(0.25+0.50n/N)Ra= 27.85 MJm-2day-1
Clear-sky solar radiation Rso =
[0.75+(2x10-5xAltitude)]Ra= 28.78 MJm-2day-1
Rs/Rso=27.85/28.78= 0.97
Net solar radiation Rns= (1 − r )Rs =0.77x27.85= 21.45

r = 0.23 for the hypothetical grass reference crop


0
From Table 7.8 Tmax= 29.8 C
σ Tmax , K 4 = 41.30 MJm-2day-1
0
From Table 7.8 Tmin= 13.7 C
σ Tmin , K 4 = 33.20 MJm-2day-1

( σ Tmax , K 4 + σ Tmin , K 4 )/2= 37.25 MJm-2day-1

0.34-0.14 e a = 0.20

1.35Rs/Rso-0.35= 0.96
Rnl = 37.61x0.20x0.96= 7.22 MJm-2day-1
Rn = 21.45-7.22= 14.23 MJm-2day-1
G= 0.14(22.5-20.5)= 0.21 MJm-2day-1
Rn-G= = 14.23-0.21= 14.02 MJm-2day-1
0.408(Rn-G)= 5.72 mm/day

0.408∆ (Rn − G ) + U 2 (es − ea )γ


900
ETo = T + 273
∆ + γ (1 + 0.34U 2 )

24
900
0.16 x5.72 + 3.2 x1.87 x0.058 x
ETo = 22.5 + 273 =7.02mm/day
0.16 + 0.058(1 + 0.34 x3.2 )

7.4.1.3 Blaney – Criddle Method


This also one of the most widely used methods in determination of evapotranspiration or
Consumptive use. The equation is given by;
U = ∑ kf = K ∑ f = KF ……………………………………………..7.25a

p (4.6t + 81.3)
f = …………….. …………………………………….7.25b
100
Where; U = seasonal consumptive use (cm)
t = mean monthly temperature (oC)
p = monthly percentage of hours of bright sunshine (of the year)
k = monthly consumptive use coefficient determined from experimental data.
f = monthly consumptive use factor
K, F = Seasonal values of consumptive use coefficient and factor respectively

∑ = refers to the summation for all the months of the growing season.

Example
Determine the evapotranspiration and irrigation requirement for wheat if the water application
efficiency is 65% and the consumptive use co-efficient for the growing season is 0.8 from the
following data:

Month Mean monthly Monthly percentage of Effective


temp; oC sunshine hours rainfall (cm)
November 18 7.20 2.6
December 15 7.15 2.8
January 13.5 7.30 3.5
February 14.5 7.10 2.0

25
Solution
Note: You may need to first refer to section 7.5

Month Mean monthly Monthly Effective Monthly consumptive


o
temp; C percentage of rainfall Pe (cm) use factor;
sunshine hours p (4.6t + 81.3)
(p) f =
100
November 18 7.20 2.6 11.82
December 15 7.15 2.8 10.74
January 13.5 7.30 3.5 10.48
February 14.5 7.10 2.0 10.50

TOTAL
∑ Pe=10.9 ∑ f =43.54

Seasonal Consumptive use, U = ∑ kf = K ∑ f


= 0.8 x 43.54
= 34.83 cm
U - ∑ Pe
Field Irrigation Requirement, F.I.R. = , where η = water application efficiency
η
34.83 - 10.90
Thus, F.I.R. = = 36.9 cm
0.65
7.4.2 Evaporation Pans
Currently, an indirect measurement of evaporation from open water is made by taking the
difference in storage of a body of water measured at two known times, which gives a measure
of the evaporated water over the time interval. If rain has fallen during the time period then the
rainfall quantity must be taken into account. In practice, this water budget method is used on
two widely differing spatial scales, by measurements at reservoirs and by measurements with
specially designed instruments maintained at meteorological stations.

i) Tanks and Pans

These are basically called evaporimeters. The advantage being that they can easily be
transferred and installed. The most widely used instrument nowadays is the American Class A
Pan (Fig. 7.5). This is circular with a diameter of 1.21 m and is 255 mm deep. It is set with the
base 150 mm above the ground surface on an open wooden frame so that the air circulates

26
freely round and under the pan. The water level in the pan is kept to about 50 mm below the
rim. The level is measured daily with a hook gauge and the difference between two readings
gives a daily value of evaporation. Alternatively, evaporation can be obtained by bringing the
water level in the pan back to a fixed level with a measured amount of water. Pans provide an
integrated measurement of the effects of radiation, wind, temperature and humidity.

Another instrument that has been accepted by many countries is the Russian tank. The USSR
GGI-3000 tank has a smaller surface area (0.3 m2, 0.618 m diameter) than the other
instruments, but has the depth of the British tank (0.60-0.685 m). It is cylindrical with a conical
base and is made of galvanized iron.

Pan Coefficient Cp
Evaporation pan data cannot be applied to free water surfaces directly but must be adjusted for
the differences inn physical and climatological factors. For example, a lake is larger and deeper
and may be exposed to different wind speed, as compared to a pan. Therefore, Evaporation
pans are not exact models of large reservoirs and have the following principle drawbacks:

1. They differ in the heat-storing capacity and heat transfer from the sides and bottom.
Also, the small volume of water in the metallic pan is greatly affected by temperature
fluctuations in the air or by solar radiations in contrast with large bodies of water with
little temperature fluctuations.

The sunken pan and floating pan aim to reduce this deficiency. As a result of this factor,
the evaporation from a pan depends to a certain extent on its size. While a pan of 3m
diameter is known to give a value which is about the same as from a neighbouring large
lake, a pan of size 1.0 m diameter indicates about 20% excess evaporation than that of
the 3m diameter pan.

2. The height of the rim in an evaporation pan affects the wind action over the surface.
Also, it casts a shadow of variable magnitude over the water surface.

3. The heat transfer characteristics of the pan material are different from that of the
reservoir.

27
In view of the above, the pan evaporation data have to be to be corrected to get the actual
evaporation from water surfaces of lakes and reservoirs under similar climatic and exposure
conditions. Thus a coefficient is introduced as:
Lake evaporation = Cp x Pan evaporation (7.31)

In which CP = Pan coefficient. The experimental values for Pan co-efficients (Cp) range from 0.6
to 0.8 depending on the wind speed and relative humidity at the pan site as well as on the type
of pan.

Evaporation Stations

It is usual to install evaporation pans in such locations where other meteorological data are also
simultaneously collected. The WMO recommends the minimum network of evaporimeter
stations as below:
1. Arid zones - One station for every 30,000km2
2. Humid temperate climates -One station for every 50,000km2
3. Cold regions - One station for every 100,000km2

28
Fig 7.5 Evaporimeters

29
Example:

Compute the daily evaporation from a Class A pan if the amounts of water added to bring the
level to the fixed point are as follows;
Day 1 2 3 4 5 6 7
Rainfall (mm) 14 6 12 8 0 5 6
Water added(mm) -5 3 0 0 7 4 3
(Removed)

(ii) What is the evaporation loss of water in this week from a lake (Surface area =
6.4x106 m2) in the vicinity, assuming a pan co-efficient of 0.75? (Ans: Lake Evaporation
during the week = 63mm X 0.75 = 47.25 mm)

7.4.3 Evaporation Index Method


Analysis of data on consumptive use indicate a high degree of correlation between pan
evaporation values and consumptive use. The relationship between Evapotranspiration (Et) and
Pan Evaporation (Ep) is usually expressed as;
Et = Kc. Ep,
where Kc is a co-efficient and is found to vary according to the stage of growth of the crop (See
Fig 7.6). The values of Kc for different crops at 5% increments of the crop growing season are
presented by G. H. Hargreaves.

Example

Assuming a growing season of 4 months December to March for Wheat, determine the
consumptive use of wheat in the month of January if the pan evaporation for the month is
9.5cm. Take the consumptive use co-efficient at 40% stage growth of the crop as 0.52.

Solution

Et = Kc Ep
The crop season is December to March, 1.e 120 days. By middle of January, the number of
days of growth is 47. i.e 47/120 = 0.4 = 40% stage growth of the crop has reached and Kc for
this stage is 0.52

Therefore; Et = 0.52 x 9.5 = 4.94 cm.


The daily consumptive use for the month of January = (4.94/31)*10 = 1.6mm/day

30
7.5 Crop Water Requirements
In designing irrigation schemes and assessing the quantity of water to be supplied, the engineer
must make estimates of crop water requirements. The Crop water requirement may be defined
as the quantity of water, regardless of its source, required by a crop or diversified pattern of
crops in a given period of time, for its normal growth under field conditions at a place. Water
requirement includes the losses due to Et or consumptive use (Cw) plus the losses during the
application of irrigation water (unavoidable losses) and the quantity of water required for special
operations such as land preparation, transplanting and leaching. It may be formulated as
follows:
WR = Et or Cw + Application losses + Special needs (7.32)
Water requirement is, therefore, a ‘demand’ and the ‘supply’ would consist of contributions from
any of the sources of water, the major source being the irrigation water (IR), effective rainfall
(ER) and soil profile contributions (S) including that from shallow water tables. That is;
WR = IR + ER + S (7.33)
The field irrigation requirement of a crop, therefore, refers to the water requirement of crops,
exclusive of effective rainfall and contribution from soil profile, and it may be given as:
IR = WR – (ER + S) (7.34)
The farm irrigation requirement depends on the irrigation needs of individual crops, their area
and the losses in the farm water distribution systems, mainly by way of seepage.
The needs vary for different crops and during the periods of growth. The several computational
methods use a reference crop evapotranspiration, which is the rate of evapotranspiration from
an extensive uniformly covered grass surface which is never short of water. This is the potential
evaporation (PE) of the Penman formula. Two other methods to derive the reference crop
evapotranspiration have been widely used namely i) the Blaney-Criddle formula developed and
applied successfully by American irrigation engineers, and
ii) the measurements of Eo from evaporation pans (Doorenbos and Pruitt, 1977).

There are two main components in the calculations:


(1) There is need to determine the reference crop evapotranspiration (ETo = PE) as in the
FAO-56 Penman Monteith equation

(a) Results of the Penman calculations in using monthly mean data provide ETo in
mm d-1 but these are subject to an adjustment factor depending on relative

31
humidity, incoming radiation and the ratio of daytime to night-time mean wind
speeds.
(b) The Blaney – Criddle formula requires only mean daily temperatures (ToC) over
each month.

ETo = P (0.46T + 8)mmd-1 (7.27)

With P the monthly percentage of hours of bright sunshine (of the year)

An adjustment factor is applied similarly for relative humidity, sunshine hours and
daytime wind speed estimates.

(c) From evaporation pan measurements.

ETo = kp E mmd-1 (mean daily value) (7.35)

Pan coefficients (kp) are available for the US Class A pan according to different ground
cover round the pan, mean relative humidity and the daily run of wind.

(2) The second stage in estimating crop water requirements is the selection of the crop
coefficient (kc) according to the cropping pattern during a production season and the
growth characteristics of the crop.

Then ETcrop = kc ETo calculated for each of the 30 or 10 day periods through the growing
season, depending on the chosen budgeting period for the application of water to
supplement any rainfall.

Irrigation is of prime importance in hot climates and especially as there is need to improve on
food security and with the challenges of global warming. In evaluating the total water
requirements for an irrigation scheme many more factors need consideration. The ETcrop is
variable over time and area and can be affected by changing local conditions. The quality of the
soil and method of application together with agricultural practices, all have to be assessed in
calculating the total water needs.

32
Crop Coefficient

The crop coefficient is dependent upon the development stage of a crop (Fig. 7.6) as follows:

The four stages of crop development are described herein as:

i) Initial stage germination and early growth when the soil surface
is not or is hardly covered by the groundcover (<10%)

ii) Crop development stage from end of initial stage to attainment of effective
full ground cover (ground cover ≈ 70 - 80%). Start of mid-
season stage can be recognized in the field when crop has
attained 70 to 80% ground cover which, however, does not
mean that the crop has reached its mature height.
Effective full groundcover refers to cover when kc is
approaching a maximum.

ii) Mid - season stage : from attainment of effective full groundcover to time
of start of maturing as indicated by discoloring of leaves
(beans) or leaves falling off (cotton). For some crops this
may extend to very near harvest (sugar beats) unless
irrigation is not applied at late season and reduction in ET
crop is induced to increase yield and/or quality (sugar
cane, cotton, some grains); normally well past the
flowering stage of annual crops.

Iv) Late season stage : from end of mid-season stage until full
maturity or harvest.

33
Fig 7.6 Crop Coefficient Curve

34
Factors affecting ETcrop
Actual ETcrop will depend on local factors which are not covered in the methods mentioned.
These factors include:
i) Climate will vary from year to year and for each period within the year. This brings about
the time variation aspect. In selecting ETcrop for project planning, knowledge should be
obtained on level and frequency at which high demands for water can be expected,
particularly in the months of peak water use.
ii) Climatic data are sometimes used from stations located some distance away from the
area under study. This is only permissible in areas where the same weather extends for
long distances.
iii) Altitude also affects the ETcrop – as radiation at high altitudes may be different to that in
low lying areas due to temperature, humidity and wind at the different altitudes.
iv) Soil Water is an important factor that includes level of available soil water, the ground
water and the salinity.
v) Methods of irrigation - techniques of irrigation if properly designed and administered can
affect the amount of evapotranspiration because different methods use different
methods of application of water.
vi) Cultural practices of the area in which the irrigation is being applied. For example the
fertiliser used is greatly dependent on method and frequency of irrigation, plant
population as obviously the more the plant cover the more the evapotranspiration,
mulching practices, use of artificial wind breakers and use of anti-transpirants in plant
foliage properties to reduce evapotranspiration.

A study (Rugumayo 2003) was carried out to determine the reliability of rainfall in relation to
crop water requirements, for different crops in selected climatic regions in Uganda. Available
rainfall data from these regions were examined for consistency using the double mass curve
and infilled using Markov generation methods. The data was then subjected to statistical tests to
determine the probability distributions that best fit them. Probability distributions were selected
from among the Log-Normal, Pearson Type III, Log-Pearson Type III and the Gumbel Extreme
Value Type I distributions. Two methods were applied in determining the most suitable
distribution, namely, the Chi-square goodness of fit test and regression analysis of the
probability plots. Representative crops from the districts were then selected and their crop water
requirements determined using the FAO Penman Monteith equation. These were compared to
the rainfall to determine the effectiveness of the rainfall in meeting crop water requirements. The

35
crop water requirements were adjusted with respect to the effective rainfall to find a planting
date that minimizes the additional water requirement. Crops that required additional water were
identified and the yield reduction due to moisture stresses determined. Irrigation schedules were
then developed for the crops that required additional water.

7.6 Effective Rainfall


It can generally be defined as useful or utilizable rainfall. It is usually interpreted differently by
specialists in different fields. From the point of view of water requirements for crops, annual or
seasonal effective rainfall can be defined as that part of the total annual or seasonal rainfall
which is useful directly or indirectly for crop production on a site where it falls. It excludes water
lost by surface run off, deep percolation, moisture remaining in the soil after crops have been
harvested. This concept is used for planning and operation of irrigation projects. Measurement
of effective rainfall is done by measuring rainfall, irrigation, losses by surface runoff, percolation
beyond the root zone and soil moisture uptake by the crop for evapotranspiration.
Factors that affect effective rainfall
a. Rainfall characteristics
b. Land slope
c. Soil characteristics
d. Ground water characteristics
e. Management practices
f. Crop characteristics
g. Carry-over soil moisture
h. Ground water contribution
i. Surface and subsurface in-flows and out-flows
j. Deep percolation
k. Measurement of effective rainfall
l. Irrigation requirements

36
Appendix
Table 7.1

Table 7.2

37
Table 7.3

38
Table 7.4

39
Table 7.5

40
Table 7.5 (continued)

41
Table 7.6
Daily extraterrestrial radiation (Ra) for different latitudes for the 15th day of the month1

G sc d r [ω s sin (ϕ )sin (δ ) + cos(φ )sin (ω s )]


24(60)
Ra=
π
(values in MJM-2day-1)2

Table 7.7

42
Mean daylight hours (N) for different latitudes for the 15th of the month1
24
N= ωs
π

43
Table 7.7 (continued)
Mean daylight hours (N) for different latitudes for the 15th of the month1
24
N= ωs
π

44
Table 7.8

45
References
1. E.M Shaw, Hydrology in Practice, Chapman and Hall, 1994, London, UK.
2. J.O Ayoade, Tropical Hydrology and Water Resources, Macmillan, 1998, London, UK.
3. D. Butler, J.W. Davies, Urban Drainage, E&FN Spon, 2000, London, UK.
4. A. M. Michael, Irrigation Theory and Practice, Vikas Publishing House PVT Ltd, 2003, New
Delhi, India
5. Guidelines for Predicting Crop Water Requirements, FAO Irrigation and Drainage Paper 24,
FAO 1977
6. W. Viessman, Jr. G.L. Lewis, Introduction to Hydrplogy, Fourth Edition, Harper Collins
College Publishers, 1996, Florida USA.
7. E. M. Wilson, Engineering Hydrology, Fourth Edition, Macmillan 1990, London, UK
8. K. Subramanya, Engineering Hydrology, Second Edition, Tata McGraw- Hill, New Dehli,
India.
9. Brutsaert, W 1982; Evapotranspiration into the atmosphere. Dordrecht, Holland: D.Reidel.
10. Crop Evapotranspiration, Guidelines for Computing Crop Water Requirements, FAO
Irrigation and Drainage Paper 56, 1998, Rome, Italy
11. J.W.Finch Empirical Factors for Estimating Open Water Evaporation from Potential
Evaporation; Water and Environmental Journal Vol. 17 No. 2003, London, UK
12. World Meteorological Organisation: Guide to Hydrological Practice, Fifth Edition, 1994,
Geneva, Switzerland
13. H.L.Penman, Natural Evaporation from Open Water, Bare Soil and Grass,
Proc.Roy.Soc.193, 1948, London, UK
14. J.S.G. McCulloch, Tables for the Rapid Computation of the Penman Estimate of
Evaporation, East African Agriculture and Forestry Journal No. 30, 286-95,1965, London,
UK.
15. C.W Thornthwaite, An Approach Towards a Rapid Classification of Climate, The
Geographical Review Vol. 38No. pp. 55-94
16. H.L.Penman Evaporation: an Introductory Survey, Netherlands Journal of Agricultural
Science, Vol 4pp 9-29,1956, Amsterdam, Netherlands.
17. B. Mangeni, G. N. Katashaya, Estimation of Lake Victoria Evaporation, Proceedings
Conference on Collaborative Technical Research, 2007 Kampala, Uganda.
18. C.O. Akinbile, Reliability Estimation of Evapotranspiration Equations under Inaccurate Data
conditions in Sub- Saharan Africa, , Proceedings Conference on Collaborative Technical
Research, 2007 Kampala, Uganda.

46
19. Moran, M, S., and R. D. Jackson (1991); Assessing the spatial distribution of
evapotranspiration using remotely sensed inputs. Journal of Environmental Quality 20: 725-
737.
20. A. Rugumayo, N. Kiiza, J. Shiima,2003; Rainfall Reliability for Crop Production, A case
study in Uganda, Proceedings of the Diffuse Pollution Conference,2003 Dublin, Republic of
Ireland.

Questions
1. a) Distinguish between the two forms of evaporation Eo and Et. What is
meant by Potential Evaporation?
2 a) Explain the assumptions and derive the Penman's Formula for the evaporation from an
open water surface. .
3. Below is information about an area in North Africa at latitude 32o N for the period 3 –
12th August 1995.
• Mean air temperature was 17.4 о C.

• Mean relative humidity was 82.4%.

• Mean number of sunshine hours, n = 2.87.

• Wind speed (u) at 2m height was 93 miles/day.

• Reflection coefficient, r = 0.05.

• Recorded vapor pressure at dew point, ed = 11.2 mmHg.

i) Use Penman’s formula to determine the potential evapotranspiration from an open water
surface in the area. (1m bar = 0.75mmHg, γ = 0.49 mmHg / о C)

ii) Use Penman’s formula to determine the evaporation for part of the area with vegetal cover
during that period. (where seasonal conversion factor f = 0.8 for August)

4) Describe the plant-water relationship as far as movement of water and its use by the
plants is concerned.
5) a) What is effective rainfall and how does it apply to crop requirements in irrigation?
b) Discuss the factors that affect the effective rainfall
6) Discuss the different methods used to determine the effect of climate on crop
water requirements. What factors affect evapotranspiration?
7). Given the monthly average climatic data of November of a station in central Uganda, located
at 10010011N and at an elevation of 1650m. Assume a seasonal conversion factor of 0.6.

47
Calculate the potential evaporation using:
a) Penman formula
b) Penman Monteith

48

Potrebbero piacerti anche