Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

CVEN9625 Fundamentals of Water Engineering

EVAPORATION AND EVAPOTRANSPIRATION

1 Introduction
Water is removed from the surface of the earth to the atmosphere by two distinct mechanisms – evaporation
and transpiration. Both describe a process whereby liquid water is transformed to a gas (water vapour). This
requires large amounts of energy. Therefore driving force behind all evaporation is the quantity of energy
received from the sun. This is why we have covered the energy balance of the earth in detail in the previous
sections.

Evaporation can (somewhat obviously) only occur where and when liquid water is available. It also requires
that the atmosphere is not saturated so that the water vapour has somewhere to go once it leaves the surface.
This chapter discusses the mechanisms for evaporation and evapotranspiration and methods for calculating its
contribution to the water cycle.

The importance of evaporation can be seen from the data in Table 1 which lists monthly average rainfall and
evaporation for Sydney. The two fluxes are very similar, indicating that runoff and infiltration could be second
order processes.
Table 1 Mean monthly distribution of rainfall and pan evaporation for Sydney (Australian Bureau of Meteorology, Stn 066062)

Mean Rainfall Mean Pan


Month
(mm) Evaporation (mm)
January 101.1 142.6
February 118.0 109.2
March 129.7 96.1
April 127.1 78.0
May 119.9 58.9
June 132.0 36.0
July 97.4 46.5
August 80.7 58.9
September 68.3 75.0
October 76.9 102.3
November 83.9 129
December 77.6 136.4
Annual 1211.8 1058.5

Average annual precipitation and evaporation data for Australia is shown in Figure 1 and Figure 2 sourced from
the Australian Bureau of Meteorology (http://www.bom.gov.au/climate/averages/maps.shtml). It can be seen
that for many parts of Australia evaporation is much larger than the rainfall.

The total evaporation from continental areas around the world is approximately 70% of total precipitation
over the continents. In Australia the ratio is much larger with evaporation accounting for approximately 90% of
the total rainfall that occurs over the continent.

Based on notes by Ashish Sharma, Ian Acworth Page 1


CVEN9625 Fundamentals of Water Engineering

Evaporation is an important part of the water balance and has large impacts on many water resources
systems. Evaporation losses from reservoirs are a substantial percentage of the total storage capacity
(generally around 20% yield) and in some cases can exceed 50%. Evaporation and evapotranspiration are also
important for agriculture. It is therefore vital that we correctly measure or estimate evaporation.

Figure 1 Average annual rainfall for Australia for the period 1961-1990 (Australian Bureau of Meteorology Product Code IDCJCM004)

Figure 2 Average annual pan evaporation for Australia for the period 1975-2005 (Australian Bureau of Meteorology Product Code
IDCJCM0006)

Based on notes by Ashish Sharma, Ian Acworth Page 2


CVEN9625 Fundamentals of Water Engineering

2 Important definitions
There are a number of key terms when thinking about evaporation and evapotranspiration.

Evaporation: the amount of water that passes or could pass into the atmosphere across a soil/air or water/air
interface

Transpiration: the process by which water is removed from vegetation into the atmosphere by evaporation
from the plant stomates. Alternately, transpiration is the transport of that water within the plant and its
subsequent release as a vapour into the atmosphere.

Evapotranspiration: the combined process of evaporation and transpiration. It describes the amount of water
that passes into the atmosphere across the plant/air interface. It is often used interchangeably with
evaporation. Commonly 'evaporation' refers to an open water surface or bare soil and 'evapotranspiration' is
used when referring to soil surfaces with plants.

Potential evaporation/evapotranspiration (ET0): the maximum amount of water that can evaporation or
transpire from a surface when water availability is not limiting (i.e. a well-watered surface or an open water
body). Potential evaporation is limited by the amount of solar radiation that is available and the capacity of the
air to receive more water.

Actual evaporation/evapotranspiration (ETa): the actual amount of water that is evaporated into the air. It is
limited by the amount of water available in the soil for the evaporation rather than the moisture holding
capacity of the air. Actual evaporation is always equal to or less than potential evaporation.

Reference crop evapotranspiration (ETrc): the rate of evapotranspiration from an idealised grass crop with an
assumed crop height (0.12 m), a fixed canopy resistance (70 s/m) and albedo (0.23).

Crop coefficient (kc): the ratio of evapotranspiration of any plant/crop compared to the reference crop defined
above.

3 Physics of evaporation
3.1 Introduction
The evaporation process is the result of an exchange of molecules between water and the atmosphere. With
an increase in the water temperature, the kinetic energy of the water molecule increases. This enables some
of them to escape from the surface. When in the vapour phase, each molecule is separate from the others by a
large distance, and hence the hydrogen bonding properties of the molecules are all but absent. Some of the
escaped molecules cool down and try to re-enter the water – this process is termed condensation. Evaporation
is the difference between the number of molecules leaving and those re-entering the water body.

There is a very thin layer of saturated water just above the water surface. This is formed due to the escape of
water molecules form the water surface and also the re-entry of some molecules. When molecules escape this
layer to the air above, space is crated for more evaporation from the water surface. This concept is
represented by Dalton's law:

E = C (es − ea ) Eqn 1

Where E is the evaporation, C is a coefficient and es is the saturation vapour pressure (at the current air
temperature) and ea is the saturation vapour pressure at the dew point temperature.

Remember that the saturation vapour pressure at the dew point temperature (ed) is the same as the actual
vapour pressure at the present air temperature (e). This means that in Equation 1 it is the difference between
the saturation vapour pressure and the actual vapour pressure that drives evaporation. As the air becomes
more saturated, ea (or e) equals es and the evaporation tends to zero. As the humidity in soils is often close to
100% (i.e. es equals ea) there is little evaporation from below the soil surface.

Based on notes by Ashish Sharma, Ian Acworth Page 3


CVEN9625 Fundamentals of Water Engineering

Saturation vapour pressure is a function of the temperature. It is low at low temperatures and increases at an
exponential rate from there as shown in Figure 3. Hence warm air can hold a lot more water than cold air. An
approximate relationship for the saturation vapour pressure is:

 17.27T 
es = 0.6108 exp  Eqn 2
 237.3 + T 
Where T is the air temperature in °C and es is the saturation vapour pressure in kPa.
Saturated Vapour Pressure (kPa)
7
6
5
4
3
2
1

-10 0 10 20 30 40
Temperature (deg C)

Figure 3 Saturation vapour pressure relationship with air temperature

3.2 Factors affecting evaporation


Vapour pressure/humidity – when water has a lower temperature than air, evaporation will be low since the
vapour pressure for cold water is low. When the air is warmer, evaporation continues until both vapour
pressures are equal. When air is cooler and the two vapour pressures are the same, the air becomes
supersaturated and a fog develops.

Temperature – an increase of temperature increases the kinetic energy of the water molecules. They then
have a better chance to escape and vapourise. Hence evaporation increases with temperature. However the
impact of other factors needs to be considered.

Atmospheric Pressure – since there is a smaller density of air molecules at a lower air peruse, there is less
chance of water molecules colliding with air molecules. So for example with an increase in altitude and a
corresponding decrease in air pressure, evaporation would increase if all other factors were the same.
However generally temperatures also decrease with altitude (lapse rate) so the relationship is not
straightforward.

Wind – Since a thin layer of saturate vapour is formed at the water surface, an increase in wind speed enables
the wind to remove this layer and to expose more water to unsaturated air. Therefore an increase in wind can
be generally expected to result in increased evaporation. There is a limit where further increases in wind speed
have little effect. This is more pronounced for larger water bodies than smaller ones.

Water quality – Dissolved solids tend to decrease the vapour pressure of water and decrease evaporation.
Seawater has lower vapour pressures than freshwater and lower evaporation is observed.

Based on notes by Ashish Sharma, Ian Acworth Page 4


CVEN9625 Fundamentals of Water Engineering

Water depth and soil type – Shallow surfaces tend to evaporate more quickly than deeper ones because they
are heated faster. For example, soil moisture and intercepted precipitation are evaporated quite fast. Darker
colour soils absorb more heat than lighter ones and hence have higher evaporation rates.

3.3 Applications of evaporation in hydrology


Evaporation is important for the design and operation of water storage reservoirs and for soil moisture. It
there has an impact on streamflows and catchment yields. Evaporation is less important during storm events,
firstly because the actual vapour pressure is close to saturation during precipitation and secondly because
storms do not usually have a very long duration.

Water resources managers can change the way that they operate regulated river systems to ensure that the
evaporation losses are minimised. For example it is better to release water from a dam in larger quantities less
frequently than to constantly release smaller amounts. This is because the water depths in the river will be
shallower when smaller amounts are released so the surface area to volume ratio will be higher and more
evaporation will result. An example is the management of Menindee Lakes in western New South Wales where
operation of the lakes is being studied to minimise evaporation losses http://www.water.nsw.gov.au/Water-
management/Water-recovery/Darling-Savings/Darling-water-saving

In some cases, evaporation can be suppressed by placing a thin film of certain chemical (e.g. cetyl alcohol) that
spread over the water surface and can reduce evaporation by as much as 70%. However the chemical layer can
be disrupted by wind and dust and can break up. This option is therefore only practicable for small dams
where wind effects are minimal.

Groundwater storage dams have also been found to be effective in some arid areas whereby the dam is filled
with sand or other relatively porous material. Water is stored in the pore spaces and evaporation is reduced.

Knowledge of evaporative processes has also been used to dispose of contaminated water by placing it is in
large evaporative ponds. This stops the contaminated water from running off or entering groundwater. The
ponds are designed to be shallow to increase the evaporation rate. Examples include brine from desalination
plants, waste water treatment plants or mine tailing water.

4 Evaporation measurement
4.1 Evaporation pan
The Class A evaporation pan is probably the most widely used instrument around the wold to measure
potential evaporation. The Class A pan is 120.7 cm in diameter and 25 cm deep and is constructed from
galvanised metal. The plan is placed in an open area and fenced to spot animals drinking from it. The water
level in the pan is maintained at a constant depth by adding or subtracting water from the pan each day. The
evaporation is calculated by considering a simple water balance by using the change in depth of the water in
the pan and the rainfall that has occurred in the previous 24 hours. The surface of the pan can either be left
open or a bird grill added. When grills were added to pans around Australia, evaporation was decreased on
average by around 7%. Long term records have been homogenised to account for this error. A Class A pan is
shown in Figure 4.

Based on notes by Ashish Sharma, Ian Acworth Page 5


CVEN9625 Fundamentals of Water Engineering

Figure 4 Class A evaporation pan in Townsville (http://www.bom.gov.au/qld/townsville/images/Evap_Pan_650.jpg)

The pan heats up more rapidly than the ground around it and there are also the side walls of the pan which
can receive some solar radiation. Therefore evaporation from a pan will be higher than from the environment.
A correction factor is therefore normally used to convert the pan evaporation measurement into true potential
evaporation. This pan factor is normally between 0.6 to 0.8 and depends on the soil type, surrounding
vegetation and climatic conditions. The pan coefficient can be calibrated for sites where enough data exists to
also directly calculated open water body evaporation using the Penman equation. In the absence of a locally
calibrated value, a table of pan coefficients is provided by Allen et al. [1998]. Using this table and average wind
speed (3.6 m/s) and relative humidity (65%) for Sydney a pan coefficient of 0.7 would be chosen (assuming 10
m of short green grass adjacent to the pan).

http://www.fao.org/docrep/X0490E/x0490e08.htm#pan%20evaporation%20method

The equation to use the pan coefficient (kpan) is:

ET0 = k pan × E pan Eqn 3

4.2 Lysimeter
A lysimeter is a tank of soil which is planted with vegetation and is hydrologically sealed so that the water
leakage from the system is negligible. It is used to measure evapotranspiration in the field and for studying
soil-water-plan relationships under natural conditions. The lysimeter should be representative of the
surrounding natural soil profile and vegetation types. The rate of evapotranspiration from this instrument is
obtained by undertaking a soil water budget. The precipitation on the lysimeter, the drainage through its
bottom, and the changes in soil moisture within the lysimeter are all measured. The amount of
evapotranspiration is the amount necessary to complete the water balance.

4.3 Eddy covariance measurement


If the energy fluxes at a site can be measured then evaporation can be calculated directly. The vertical
fluctuations of the wind and water vapour are measured and then their correlations calculated over some
averaging period (around 15 minutes to an hour). It is only in the last 10 to 15 years that suitable
instrumentation has become commercially available. However the instrumentation is expensive and requires
special skill to operated and therefore this method is only used in research experiments. It is the preferred

Based on notes by Ashish Sharma, Ian Acworth Page 6


CVEN9625 Fundamentals of Water Engineering

micrometeorological technique on the grounds that it is a direct measurement with minimum theoretical
assumptions. A map showing the locations of eddy covariance stations in Australia is in Figure 5.

Figure 5 Network of meteorological flux stations in Australia and New Zealand (http://www.ozflux.org.au/monitoringsites/index.html)

5 Evaporation calculations
As can be seen from the methods above the measurement of evaporation is labour intensive and expensive.
Therefore in most cases evaporation is calculated by considering the physical relationship between different
climatic variables and the evaporation rate. There are a number of different methods for calculating
evaporation/evapotranspiration and a comprehensive review of the different methods is provided by
McMahon et al. [2013]. In general the methods can be classified as:

• temperature-based methods
• radiation-based methods
• combination methods (resistance plus energy)

If all the required climatic data are available then the Penman Monteith method (a combination approach) is
recommended as the most accurate approach. Details of this method are provided in the next section.

5.1 Energy balance to drive evaporation


As discussed above evaporation is driven by energy allowing water molecules to escape from the water
surface. Therefore the general principle of calculating evaporation is to use consider the energy budget. Figure
6 shows the components of the whole energy balance for daytime conditions (i.e. when there is incoming solar
radiation).

Based on notes by Ashish Sharma, Ian Acworth Page 7


CVEN9625 Fundamentals of Water Engineering

Figure 6 Energy fluxes in a control volume for a plant canopy [Shuttleworth, 1993]

The available energy A is the energy balance of these components as shown in Equation 4:

𝐴𝐴 = 𝑅𝑅𝑛𝑛 − 𝐺𝐺 − 𝑆𝑆 − 𝑃𝑃 − 𝐴𝐴𝑑𝑑 Eqn 4


Where

• A is Available Energy
• Rn is Net Incoming Radiation (i.e. considering the solar and longwave radiation components and
directions)
• G is the outgoing heat conduction into the soil
• S is the energy temporarily stored within the soil volume. It is proportional to temperature changes in
the vegetation, air and shallow soil layer and to changes in atmospheric humidity.
• P is the energy absorbed by biochemical processes in the plants, typically taken as 2% of net radiation
• Ad is the energy associated with horizontal air movement, only significant in oasis situations

Under most conditions the terms S, P and Ad are neglected. The temporary soil volume energy (S) needs to be
considered when the energy balance is over a forest. Over the course of a day G is approximately equal to zero
so can also generally be neglected if daily evaporation estimates are required. Therefore the available energy
can be approximated as the net radiation.

As shown in Figure 6 and Equation 5, the available energy A can be partitioned into two components – sensible
heat H and latent energy λE (i.e. the outgoing energy in the form of evaporation)

𝐴𝐴 = 𝐻𝐻 + 𝜆𝜆𝜆𝜆 Eqn 5
Thus if there is limited water available for evaporation, the sensible heat partition will become larger and the
air temperatures will be higher. The ratio between sensible heat and latent heat is called the Bowen Ratio and
can be used to summarise the aridity of a location.

𝛽𝛽 = 𝐻𝐻/𝜆𝜆𝜆𝜆 Eqn 6
Table 0-2 lists Bowen ratios for a number of different climatic conditions.

Based on notes by Ashish Sharma, Ian Acworth Page 8


CVEN9625 Fundamentals of Water Engineering

Table 2 Typical values of the Bowen ration [Ladson, 2008]

Conditions Bowen ratio


Arid conditions (hot deserts) 10
Semi-arid regions 2-6
Temperate forests and grass lands 0.4-0.8
Tropical rain forests 0.2
Tropical oceans 0.1
Well watered short vegetation with no wind and low temperatures ~0
(i.e. close to zero sensible heat flux)
Well watered vegetation with low humidity. In this case the leaf <0
temperature can be less than the air temperature because of
evaporative cooling so the sensible heat is providing additional
energy for evaporation i.e. the Bowen ratio can be negative

5.2 Surface and aerodynamic resistance


Surface (bulk) resistance
Water flows from the soil to roots, through xylem, through mesophyli cells and cell walls and finally evaporates
into substomatal cavities. It then diffuses out of the stomates, through the leaf and canopy layers, and is mixed
with the bulk atmosphere. Water potential is highest in the soil and deceases along the transpiration path. This
potential gradient provides the driving force for water transport form the soil to the atmosphere. Rough
estimates of plant resistance can be made by considering typical potentials and transpiration rates.

Nomenclature in different textbooks describing the resistance terms varies and should be carefully checked.

The stomatal resistance of the whole canopy (rs) is smaller when more leaves are present since there are more
stomata through which the vapours can be emitted. The movement of water vapour from inside plant leaves
to the air is governed by small apertures know as stomata. The air insides the stomatal cavity is nearly
saturated while the outside air is usually less saturated. The movement of vapour is controlled by the plant
through these cavities. The plant can open and close the cavities depending on the atmospheric moisture
demand (es – ea) and the amount of moisture in the soil.

For a short green grass crop not short of water (reference crop discussed below) the total surface resistance rs
-1
is calculated to be 70 s m . As canopy cover deceases and water availability in the soil decrease the value of rs
increases to several thousand. For a wet canopy when soil moisture is not limiting, rc approaches zero.

One useful approximation for rs is to relate it to the Leaf Area Index (LAI):

200
rc = Eqn 7
LAI

Aerodynamic resistance
Another important diffusion process is turbulent diffusion. This refers to the turbulence caused by the
retardation of the horizontal movement of wind over a crop (or water) surface. This is a more important
process for exchange of air from close to the ground to higher levels of the atmosphere than molecular
diffusion discussed above.

The resistance due to turbulent diffusion is called the aerodynamic resistance (ra) and is inversely proportional
to wind speed:

Based on notes by Ashish Sharma, Ian Acworth Page 9


CVEN9625 Fundamentals of Water Engineering

 z − 0.67 h   z h − 0.67 h 
ln m  ln 
ra = 
0.123h   0.0123h 
Eqn 8
k 2u z

Where:
-1
ra is the aerodynamic resistance (s m )

zm and zh are the heights of the wind and humidity measurements respectively (m)

h is the crop height (m)

k is von Karman's constant of 0.41


-1
uz is the wind speed at height z (m s )

It can be seen from Equation 8 that as wind speed increases the aerodynamic resistance decreases and hence
evaporation will be higher. Less clear is that as the crop height increases, the numerator of Equation 7
decreases and the aerodynamic resistance also becomes smaller. Thus the aerodynamic exchange for taller
crops is more efficient than for shorter crops [Shuttleworth, 1993]. Water caught on the leaves of taller crops
during rain or irrigation will evaporation more quickly than shorter crops which has implications for
agricultural land management.

5.3 Available energy


Shortwave (solar) radiation
The extraterrestrial solar radiation is the radiation received at the top of the earth's atmosphere on a
horizontal surface. It changes throughout the year due to changes in the position of the sun and the length of
the day. It is therefore a function of the latitude, date and time of day. These values can be substituted into
-2 -1
the following equation to calculate the extraterrestrial solar radiation Ra (MJ m day ).

Ra =
118.1
π
(
d r ω s sin φ sin d + cos φ cos d sin ω s ) Eqn 9

where ωs represents the sunset hour angle:

ω s = arccos(− tan φ tan δ ) Eqn 10

and φ is the latitude for the site (negative for Southern Hemisphere) with d the solar declination (in radians),
given as:

 2π 
δ = 0.4093 sin J − 1.405  Eqn 11
 365 

and J is the Julian day number (day number from start of year).

The relative distance between the earth and sun is calculated as:

 2π 
d r = 1 + 0.033 cos J Eqn 12
 365 

Not all the energy at the top of the atmosphere reaches the earth's surface and therefore solar radiation (Rs)
at the surface will be less than extraterrestrial solar radiation. On a cloudless day clear sky solar radiation (Rso)
is approximately 75% of the extraterrestrial radiation. When there are clouds the solar radiation will be even
lower.

Solar radiation (Rs) can be calculated using the Angstrom formula:

Based on notes by Ashish Sharma, Ian Acworth Page 10


CVEN9625 Fundamentals of Water Engineering

 n
Rs = Ra  0.25 + 0.5  Eqn 13
 N

Where n is the actual duration of sunshine (hours) and N is the maximum possible duration of sunshine or
daylight hours (hours) calculated as:

24
N= ωs Eqn 14
π
The constants in the Angstrom formula can vary depending on location but the above values are
recommended by Allen et al. [1998] in the absence of local data.

The net solar radiation (Rns) is the balance between the incoming and reflected solar radiation which is
controlled by the albedo (α). An albedo of 0.23 is assumed for the reference crop (discussed below).

Rns = Rs (1 − α ) Eqn 15

Longwave (terrestrial) radiation


The longwave energy is described by the Stefan-Boltzmann law which states that the energy emission is
proportional to the absolute temperature of the surface raised to the fourth power. Clouds, water vapour,
carbon dioxide and dust can absorb the emitted longwave radiation and re-emit towards earth. Therefore net
outgoing longwave radiation will be smaller when there is higher cloudiness or humidity. This relationship is
shown in the following equation:

T4 +T4 
Rnl = s  max, K min, K
 2
(
 0.34 − 0.14 e
 a ) 1.35 RR
s 
− 0.35  Eqn 16
   so 

Where:
-2 -1
Rnl is the net outgoing longwave radiation (MJ m day )
-8 -2 -4
σ is the Stefan Boltzmann constant = 4.903 x 10-9 MJ m-2 K-4 Day ( = 5.67037 x 10 W m K )

Tmax,K and Tmin,K are the maximum and minimum daily air temperature (K)

ea is the actual vapour pressure (kPa)

and Rso is found using:

(
Rso = Ra 0.75 + 2 ×10−5 z ) Eqn 17

Where z is the station elevation (m above sea level). Once again the constants in this equation can be locally
calibrated. More details are provided in Allen et al. [1998].

Net radiation
Net radiation is simply the difference between incoming net shortwave radiation and outgoing net longwave
radiation:

Rn = Rns − Rnl Eqn 18

Other heat fluxes (if significant) are subtracted from the net radiation in Equation 18 to arrive at the available
-2 -1
energy (Equation 4). Note that the above estimate is in MJ m day which can be converted to mm units by
dividing it by the latent heat of vaporisation of water. The following conversion may be used to convert energy
to other units:
-2 -1 -2 -1
1 MJ m day = 11.57 W m = 0.408 mm day (at 20°C) Eqn 19

Based on notes by Ashish Sharma, Ian Acworth Page 11


CVEN9625 Fundamentals of Water Engineering

5.4 Penman-Monteith equation


Penman [1948] combined the energy balance with the mass transfer method and derived an equation to
compute the evaporation from an open water surface from standard climatological records of sunshine,
temperature, humidity and wind speed. This so-called combination method was further developed by many
researchers and extended to cropped surfaces by introducing resistance factors.

As was shown in Equation 1, evaporation is controlled by the difference between the saturated vapour
pressure es and actual vapour pressure ea (equivalent to saturation vapour pressure at the dew point
temperature). The vapour pressure deficit is normally denoted as D such that:

D = e s − ea Eqn 20

The Penman-Monteith approach is called a combination approach because it calculates evaporation as the
weighted combination of the available energy and the vapour pressure deficit. The general form for the
equation is therefore:

DA + rc p D
ra
λE = Eqn 21
D + γ (1 + rs ra )

Where:
-2 -1
λE is the latent heat flux of evaporation (kJ m s )
-1
E is the evaporation rate (m s )
-1
λ is the latent heat of vapourisation (MJ kg )

Δ is the slope of the saturated vapour pressure – temperature curve which was shown in Figure 3
-2 -1
A is the available energy (kJ m s )

D is the vapour pressure deficit (kPa)


-3
ρ is the density of air (kg m )
-1 -1
cp is the specific heat of moist air (kJ kg °C ) and is equal to 1.013
-1
rs is the surface resistance (s m )
-1
ra is the aerodynamic resistance (s m )
-1
γ is the psychometric constant (kPa °C )

The slope of the saturation vapour pressure relationship with respect to temperature is:

4098es
∆= Eqn 22
(237.3 + T )2
The latent heat of vapourisation (λ) can be calculated using Equation 21 if the surface temperature of the
water surface (Ts) in °C is known

λ = 2.501 − 0.002361Ts Eqn 23

Finally the psychometric constant (γ) is defined as:

P
γ = 0.00163 Eqn 24
λ
Where P is the atmospheric pressure (kPa). In the absence of data on atmospheric pressure an estimate can be
made using the site elevation (z) in units of metres:

Based on notes by Ashish Sharma, Ian Acworth Page 12


CVEN9625 Fundamentals of Water Engineering

5.26
 293 − 0.0065 z 
P = 101.3  Eqn 25
 293 
The combination approach can be seen more clearly if Equation 21 is split into two components:

ET0 = ETrad + ETaero Eqn 26

Where ET0 is the potential evapotranspiration and ETrad is the contribution from radiation energy input (i.e.
available energy) and ETaero is the contribution from the aerodynamic component (driven by the vapour
pressure deficit and advection from wind).

Alternatively we can write the equation as:

ET0 = FA A + FD D Eqn 27

In this form the weighting factors FA and FD depend on whether evapotranspiration or open water body
evaporation is being calculated. Firstly we will look at the estimate for the reference crop evapotranspiration,
and then with open water body evaporation.

5.5 Penman-Monteith reference crop evapotranspiration


As described in the section on resistance above, aerodynamic resistance will vary according to the plant type.
Therefore to standardise the estimates from the Penman-Monteith equation, a reference crop has been
rc -1
defined by Allen et al. [1998] which has a surface resistance rc = 70 s m . The reference crop is defined as a
hypothetical crop with a height of 0.12 m and an albedo of 0.23. The reference surface is assumed to be of
green grass of uniform height which is actively growing. Importantly the crop is completely shading the ground
and has adequate water so that it is forms potential evapotranspiration conditions. The requirements that the
grass surface should be extensive and uniform result from the assumption that all fluxes are one-dimensional
upwards [Allen et al., 1998].

Using Equation 21 and standard meteorological observations and the information on the reference crop, the
reference crop evapotranspiration is estimated as:

900
0.408 D A + γ u2 D
ETrc = T + 273 Eqn 28
D + γ (1 + 0.34 u 2 )
-1
u2 is the wind speed at 2 m height (m s )

T is air temperature at 2 m height (°C)


-1
ETrc is reference crop evapotranspiration (mm day )
-2 -1 -1
The units for A should be MJ m day and D in kPa to give the evapotranspiration in units of mm day . Note
-1
that the constant of 900 has units of kJ kg K.

In practice actual vapour pressure may not be available (if dew point temperature has not been recorded) and
therefore it may need to be calculated from relative humidity measurements. Because the saturated vapour
pressure curve is non-linear, average saturated vapour pressure cannot be calculated using average
temperature. Therefore average saturated vapour pressure needs to be calculated using the minimum and
maximum temperatures. Allen et al. [1998] recommends the following procedures to estimate daily average
saturated and actual vapour pressure (es and ea respectively)

(
es = 0.5 × e o (Tmin ) + e o (Tmax ) ) Eqn 29
o
Where e (T) is the saturated vapour pressure calculated at a specific temperature (T) using Equation 2 and Tmin
and Tmax are the daily minimum and maximum temperatures for which the vapour pressures are calculated.
If maximum and minimum relative humidity data is available then the actual vapour pressure (ea) is calculated
as:

Based on notes by Ashish Sharma, Ian Acworth Page 13


CVEN9625 Fundamentals of Water Engineering

(
ea = 0.5 × e o (Tmin ) × RH max + e o (Tmax ) × RH min ) Eqn 30

Where RHmax and RHmin are the maximum and minimum relative humidites (in %) for the day. The idea is that
the maximum relative humidity generally occurs in the morning when temperatures are lowest and the lowest
relative humidity occurs in the afternoon when temperatures are highest.
Refer to Allen et al. [1998] for details of other methods to calculate actual vapour pressure in the absence of
minimum and/or maximum relative humidity measurements.
If wind measurements at a height of 2m are not available, the following equation may be used to convert
measurements from a height zm to corresponding values for a height of 2m:

ln(2 / 0.0023)
u2 = u z Eqn 31
ln( z m / 0.0023)

where uz is the wind speed measured at a height of zm. Commonly wind speed measurements are made at a
height of 10 m.

5.6 Penman open water body evaporation


For open water body evaporation the surface resistance can be neglected (rs = 0) in Equation 21 and thus the
form for the equation using standard meteorological variables is:

∆ A + 6.43 γ ( 1 + 0.53u2 ) ∆
ET0 = Eqn 32
(∆ + γ )λ
-2 -1
As for the reference crop, the units for A should be MJ m day and D in kPa to give the evaporation in units of
-1
mm day

5.7 Other methods


Radiation based equations
The Priestley Taylor equation [Priestley and Taylor, 1972] is a simpler relationship between reference crop
evaporation and the available energy, leaving out the vapour pressure deficit part of the Penman Monteith
equation, on the basis that the first term usually exceeds the second by a factor of four [Shuttleworth, 1993].
This is given as:

∆A
ETrc = α Eqn 33
∆ +γ
where α has been empirically estimated as 1.74 for arid climates with relative humidity less than 60% in the
month with peak evaporation and 1.26 for humid climates.

Empirical equations
There are a number of empirically based equations, particularly based on temperature, that are widely
referenced or have been commonly used in the past [McMahon et al., 2013]. The physical basis for estimating
evaporation using temperature alone is that both radiation and vapour pressure deficit are likely to have some
relationship with temperature. In general the only justification of using estimation equations of this type is
that temperature is the only available variable that has been measured. In this case it is unwise to make
evaporation estimates for less than a monthly averaging period [Shuttleworth, 1993]. McMahon et al. [2013]
also recommend the use of physically based equations (such as the Penman-Montheith method) should be
preferred compared to the empirical relationships – particularly for areas where the empirical coefficients
were not derived.

Based on notes by Ashish Sharma, Ian Acworth Page 14


CVEN9625 Fundamentals of Water Engineering

Hargreaves

The Hargreaves equation [McMahon et al., 2013] provides an empirically derived estimate of reference crop
evaporation. This is a temperature and radiation based method. It should only be used with weekly or monthly
data. It can be stated as:

ETrc = 0.0135C HS
Ra
(Tmax − Tmin )0.5 (T + 17.8) Eqn 34
λ
-2 -1
where Ra is the extraterrestrial solar radiation (MJ m day ), Tmin, Tmax and T are the minimum, maximum and
-1
average daily air temperatures respectively in °C. ETrc is in units of mm day . CHS is a modification to the
empirical coefficient to reduce the error in the solar radiation estimate. CHS is estimated as:

C HS = 0.00185(Tmax − Tmin )2 − 0.0433(Tmax − Tmin ) + 0.4023 Eqn 35

Blaney-Criddle

Another temperature based evaporation equation is the Blaney-Criddle equation [McMahon et al., 2013] and
-1
provides estimates of reference crop evapotranspiration (mm day ). It requires measurements of sunshine
hours, wind speed and relative humidity compared to the Hargreaves or Thornthwaite methods which only
require temperature. The equation is:

ETrc = a BC + bBC p y (0.46T + 8.13) Eqn 36

with
n
a BC = 0.0043RH min − ( ) − 1.41
N

n n
bBC = 0.8192 − 0.0041RH min + 1.0705 + 0.0656u 2 − 0.0060 RH min − 0.0006 RH min u 2
N N

where py is the percentage of actual daily daytime hours for the day compared to the day-light hours for the
entire year, T is the mean air temperature in °C, n/N is the measured sunshine hours divided by the possible
daily sunshine hours, RHmin is the minimum daily relative humidity (%) and u2 is the average daily wind speed at
-1
2m height (m s ). The empirical coefficients were based on lysimeter data located in the dry western USA
where advection effects are strong. The complexity of this equation is primarily due to a number of
modifications over the years. It is also possible to adjust the equation for elevation in arid and semi-arid
regions.

Thornthwaite

The Thornthwaite method [Shaw, 1994] provides estimates of potential evapotranspiration using only mean
monthly temperature data. The estimates are based on climatological average temperatures and therefore
provide a climatological estimate of evaporation rather than true evaporation for any particular day or month.

a
 10T 
ETo = 16  Eqn 37
 I 
Where I is a heat index computed using all monthly average temperatures as:

1.514
12
 Tj 
I= ∑ 
 5
j =1 



Eqn 38

And a is:

Based on notes by Ashish Sharma, Ian Acworth Page 15


CVEN9625 Fundamentals of Water Engineering

a = 6.75 × 10 −7 I 3 − 7.71× 10 −5 I 2 + 1.792 × 10 −2 I + 0.49239 Eqn 39

5.8 Estimation of potential evapotranspiration for other crops


Since the reference crop is rarely of interest, a conversion of the reference crop evaporation rate is needed.
This conversion is through the crop coefficient (Kc) which provides a conversion factor for a different crop, or
under non-ideal soil moisture conditions. The crop coefficient is a quite complex to quantify since it depends
on both the surface and aerodynamic resistances in proportions that depend on the other conditions
associated with individual crop types.

It is usually easier to estimate this coefficient for an irrigated crop (since then there is no need to worry about
the limiting presence of moisture in the soil). The crop coefficient for an irrigated surface is called the potential
crop coefficient (Kco) and can be estimated with more reliability than the crop coefficient Kc.

ET0 = K co ETrc Eqn 40

Considerable research effort has been devoted to estimate Kco as a function of time. It is usually observed to
vary during the growing season as shown in Figure 7.

Figure 7 Typical variation of the potential crop coefficient during the growing season (Shuttleworth, 1993)

The temporal variation of the potential crop coefficient is simplified by dividing the crop growth into four
stages. Potential crop coefficients for each stage are then estimated. The procedure is:

• Establish the planting date, and use Table 0-3 to estimate cut-off dates (D2, D3, D4, and D5) for the
other stages.
• Estimate the constant value of Kco for stage 1 based on Figure 8 as a function of reference crop
evaporation and irrigation efficiency.
• For a given crop type, the constant Kco value in stage 3, and at end of stage 4, are given in Table 3.

Based on notes by Ashish Sharma, Ian Acworth Page 16


CVEN9625 Fundamentals of Water Engineering

• The full seasonal variation in Kco is then obtained by interpolating between the values obtained in the
previous steps.

Figure 8 Values for KCO for stage 1 of the growing season (Shuttleworth, 1993)

An example is shown in Figure 8 where the reference crop evaporation rate (Erc) was found to be 8.4 mm/day
and the irrigation interval is 7 days. In this case Kco will be 0.35.
Table 3 Growing season estimates and values of Kco for Stage 3 and Stage 4 (Shuttleworth, 1993)

Based on notes by Ashish Sharma, Ian Acworth Page 17


CVEN9625 Fundamentals of Water Engineering

When the crop does not have a definite growing season, such as irrigated grass-like crops which are
successively cut throughout the season or well-irrigated and well-fertilised pastures (grass, or alfafa) then the
value in Table 0-4 are used and are assumed to be constant throughout the year.

Table 4 Values of Kco for successively cut fodder crops and grazed pastures which are well-watered and fertilised (Shuttleworth, 1993)

RH > 70% RH < 30%


-1 -1
Crop Wind (ms ) Wind (ms )
0-5 5+ 0-5 5+
Fodder (Alfalfa) 0.85 1.05 0.95 1.05
Fodder (Grasses) 0.80 1.00 0.90 1.00
Fodder (Clover/grass-legumes) 1.00 1.10 1.05 1.10
Pasture (grass, grass-legumes,
0.95 1.05 1.00 1.05
alfalfa)

5.9 Calculating actual evapotranspiration


The water status of the soil is very important in estimating the evapotranspiration for a crop that is not
regularly irrigated. The value of the crop coefficient is reduced from Kco by a factor related to the volumetric
soil water content as shown below, where f(θ) is the soil moisture extraction function.

ETa = K c ETrc = f (θ ) K co ETrc Eqn 41

A typical relationship for the soil moisture extraction function is shown in Figure 9.

Figure 9 Typical variation of the soil moisture extraction function f(θ)

As shown in Figure 9, soil saturated by irrigation has a moisture content θs (saturation moisture content), and
drains until the water that remains is held by surface tension that is in equilibrium with the gravitational forces
that pull it down. It is then said to be at field capacity (θf). The drying process goes on till the moisture content
falls down to θd, which is 50-80% of the field capacity moisture content, at which point the soil hydraulic
conductivity and transpiration rate start to fall. They continue to fall till the wilting point (θw). There are a
number of different forms for f(θ) as shown in Table 0-5 but the most common is to assume the shape shown
in Figure 9.

Wilting point (θw) is the minimal point of soil moisture the plant requires not to wilt. If moisture decreases to
this or any lower point, the plant wilts and can no longer recover its turgidity. Field capacity (θf) is the amount
of soil moisture or water content held in soil after excess water has drained away.

Based on notes by Ashish Sharma, Ian Acworth Page 18


CVEN9625 Fundamentals of Water Engineering

Therefore to calculate the actual evapotranspiration, all that is required is to calculate the actual soil moisture
θ and relate it to the wilting point and field capacity. The actual soil moisture can be found by carrying out a
running water balance taking into account infiltration that may occur due to rainfall and irrigation, and the loss
of water as the actual evapotranspiration for the previous time step. Since the amount of water assessable to
the plant depends on the rooting depth (which can vary over the year particularly if it is an annual crop), the
best models for estimation of soil water extraction discretise the soil into several small layers, each of which
acts as a moisture store arranged vertically over each other. The precise form of f(θ) is not too important for
many hydrological applications because the cumulative error is reset to zero each time the soil completely
dries or is completely full of moisture after sufficient rainfall.
Table 5 Recommended forms of soil moisture stress function f(θ) (Shuttleworth, 1993)

Fig 9

6 References
Allen, R. G., L. S. Pereira, D. Raes, and M. Smith (1998), Crop evapotranspiration - Guidelines for computing
crop water requirements, FAO - Food and Agriculture Organization of the United Names, Rome.
Ladson, A. R. (2008), Hydrology: an Australian introduction, Oxford university press.
McMahon, T., M. Peel, L. Lowe, R. Srikanthan, and T. McVicar (2013), Estimating actual, potential, reference
crop and pan evaporation using standard meteorological data: a pragmatic synthesis, Hydrology and Earth
System Sciences, 17(4), 1331-1363.
Penman, H. L. (1948), Natural Evaporation from Open Water, Bare Soil and Grass, Proceeding of the Royal
Society of London, Series A, Mathematical and Physical Sciences, 193(1032), 120-145.
Priestley, C. H., and R. J. Taylor (1972), Assessment of Surface Heat-Flux and Evaporation Using Large-Scale
Parameters, Monthly Weather Review, 100(2), 81-92.
Shaw, E. M. (1994), Hydrology in Practice, Chapman & Hall, London.
Shuttleworth, W. J. (1993), Evaporation, in Handbook of Hydrology, edited by D. R. Maidment, McGraw-Hill
Inc, New York.

Based on notes by Ashish Sharma, Ian Acworth Page 19

You might also like