Anda di halaman 1dari 114

INDUSTRIETECHNIK

SRI LANKA INSTITUTE of ADVANCED TECHNOLOGICAL EDUCATION

ATI Labuduwa

Civil Engineering
Trainee Manual

Training Unit

Hydrology Engineering
Theory

No: CE 037

Training Unit
Hydrology Engineering
Theoretical Part
No.: CE 037

Edition:

2009
All Rights Reserved

Editor:

MCE Industrietechnik Linz GmbH & Co


Education and Training Systems, DM-1
Lunzerstrasse 64 P.O.Box 36, A 4031 Linz / Austria
Tel. (+ 43 / 732) 6987 3475
Fax (+ 43 / 732) 6980 4271
Website: www.mcelinz.com
1
1/113

HYDROLOGY ENGINEERING
LIST OF CONTENT

CONTENTS

Page

PREFACE .............................................................................................................................5
1

INTRODUCTION ...........................................................................................................6
1.1

Hydrologic Cycle ...................................................................................................6

1.2

Application of Hydrology in Practice......................................................................8

1.3

Questions ..............................................................................................................9

HYDROLOGICAL PROCESSES .................................................................................10


2.1

Precepitation and its Procesess ..........................................................................10

2.2

Rainfall Characteristics........................................................................................15

2.3

Measurement of Precipitation..............................................................................16

2.3.1

Point Precipitation Measurement ....................................................................16

2.3.2

Estimation of Areal Precipitation .....................................................................18

2.4
2.4.1
2.5

Factors Influencing the Rate of Evaporation ...................................................23


Evapo- Transpiration and its Estimation..............................................................24

2.5.1

Equations ........................................................................................................24

2.5.2

Penman-Monteith Equation.............................................................................26

2.5.3

FAO Penman-Monteith Equation ....................................................................32

2.6
3

Evaporation and it its Estimation .........................................................................22

Possible Questions..............................................................................................36

STREAM FLOW MEASUREMENT..............................................................................38


3.1

Introduction..........................................................................................................38

3.2

Measurement of Stage and Velocity ...................................................................39

3.3

Stage-Discharge Relation: Rating Curve ............................................................39

3.3.1

Relation of Stream Stage to Streamflow is Always Changing .........................41

3.3.2

Method of Estimation of Flow by Direct Measurement ....................................42

3.4

Questions ............................................................................................................44

2
2/113

HYDROGRAPH ANALYSIS ........................................................................................46


4.1

Introduction..........................................................................................................46

4.2

Hydrograph and its Components.........................................................................46

4.2.1
4.3

Flow Duration Curve............................................................................................48

4.4

Effective Rainfall..................................................................................................50

4.5

Unit Hydrograph ..................................................................................................52

4.5.1

Principles and Definitions ................................................................................52

4.5.2

Unit Hydrograph Derivation .............................................................................57

4.5.3

The "S" Hydrograph.........................................................................................59

4.5.4

The Instantaneous Unit Hydrograph (IUH) ......................................................61

4.5.4.1

Definition and Properties .........................................................................61

4.5.4.2

Instantaneous Unit Hydrograph Resulting From Conceptual Models......63

4.5.4.3

The Instantaneous Unit Hydrograph of Nash ..........................................64

4.6
4.6.1
4.7

The Geomorphologic Unit Hydrograph................................................................68


Principles and Methods of Determination........................................................68
Regionalisation of the Unit Hydrograph (Synthetic Unit Hydrograph) .................71

4.7.1

The Synthetic SCS Unit Hydrograph ...............................................................71

4.7.2

Snyder's Synthetic Unit Hydrograph................................................................74

4.8
5

Factors Affecting Shape of Hydrograph ..........................................................48

Workout Examples and Exercises.......................................................................77

HYDROLOGIC ROUTING ...........................................................................................82


5.1

Introduction..........................................................................................................82

5.2

Reservoir Routing................................................................................................83

5.3

Channel Routing..................................................................................................85

5.4

One-Dimensional Hydraulic Routing ...................................................................88

FREGUENCY ANALYSIS OF HYDROLOGIC EVENTS .............................................93


6.1

Prediction Methods of Extreme Events and Confidence Limit ............................93

6.1.1

Gumbels Equation for Practical Use...............................................................93

6.1.2

Log-Pearson Type III Distribution....................................................................94

6.2

Hydrologic data series.........................................................................................97

6.3

Return Period and Exceedence Probability:........................................................98

3
3/113

INTRODUCTION TO GROUND WATER HYDROLOGY ............................................99


7.1

Forms of Subsurface Water ................................................................................99

7.2

Aquifers and Confining Bed...............................................................................100

7.3

Properties of a Aquifers.....................................................................................101

7.4

Basic Equation of Ground Water Flow ..............................................................104

7.5

Transmissivity (T) ..............................................................................................107

7.6

Storage Coefficient (S) ......................................................................................108

7.7

Cone of Depression...........................................................................................108

7.8

Workout Examples ............................................................................................109

7.9

Exercises...........................................................................................................112

References........................................................................................................................113

4
4/113

HYDROLOGY ENGINEERING

PREFACE
These lecture materials have been complied from available books, internet sources, the
existing lecture notes of the contributor. Teaching experiences in present institute is the
basis of preparing the material. It is believed that the user of this note should have basic
knowledge in higher school levels in science, or who will be instructing should have very
good knowledge in Water Resources Engineering related subjects such as Surface and
sub surface Hydraulics, Hydrology, Hydrogeology and ground water Engineering. The
contents of this notes is mainly comprises of the aforesaid subjects. Not like an exactly as
book form, attempt have been made to cover salient topics related to the subject named:
Hydrology Engineering. The students are suggested to consult the references given at the
end of this compilation, if necessary.
Conventionally, the person who teaches, he will have to prepare the deliberations of slide,
transparencies as per his schedule of classes. It is important to note that instructor who will
teach this material will have to prepare their slides or presentation material according his
convenience.
This instruction material may have required some unintentional typological or other
corrections, left the scope of future improvement.

5
5/113

HYDROLOGY ENGINEERING

INTRODUCTION

Hydrology (from Greek: Y, hydr, "water"; and , logos, "study") is the study of the
movement, distribution, and quality of water throughout the Earth, and thus addresses both
the hydrologic cycle and water resources. A practitioner of hydrology is a hydrologist,
working within the fields of earth or environmental science, physical geography or civil and
environmental engineering.
Domains of hydrology include hydrometeorology, surface hydrology, hydrogeology,
drainage basin management and water quality, where water plays the central role.
Hydrological research is useful as it allows us to better understand the world in which we
live, and also provides insight for environmental engineering, policy and planning.

1.1

Hydrologic Cycle

Water is essential to life. Without it, the biosphere that exists on the surface of the earth
wouldn't be possible. Nicknamed the "water" planet, Earth is covered by one of our most
precious resources. However, almost 93% is locked in the oceans, toxic to humans and
many plants and animals.
How do we obtain fresh water resources then?
Where does drinkable water come from?
To understand, we need to turn to the Hydrologic Cycle.

6
6/113

Fig. 1 The Hydrologic Cycle


Water has many unique properties that allow it to be such a universal material. One special
characteristic of water is its ability to change state very easily under Earth conditions. It can
be found readily on the planet in all of its three forms, solid, liquid, and gas. These forms
also play a great part in the hydrologic cycle. Now, exactly what is the hydrologic cycle?
The hydrologic cycle takes place in the hydrosphere, this is the region containing all the
water in the atmosphere and on the surface of the earth. The cycle is the movement of
water through this hydrosphere.
Now the entire process is very simple, divided in to five parts

Condensation

Infiltration

Runoff

Evaporation

Precipitation

The process begins with condensation, when water vapor condenses in the atmosphere to
form clouds. Condensation occurs when the temperature of the air or earth changes. Water
changes states when temperatures fluctuate. So when the air cools enough, water vapor
has to condense on particles in the air to form clouds. This process is very noticeable on
plants as they dew in the morning.

7
7/113

As clouds form, winds move them across the globe, spreading out the water vapor. When
eventually the clouds can't hold the moisture, they release it in the form of precipitation,
which can be rain, hail, snow etc.
The next three stages: infiltration, runoff, and evaporation occur simultaneously. Infiltration
occurs when precipitation seeps into the ground. This depends a lot on the permeability of
the ground.
Permeability is the measure of how easily something flows through a substance. The more
permeable, the more precipitation seeps into the ground. If precipitation occurs faster than
it can infiltrate the ground, it becomes runoff. Runoff remains on the surface and flows into
streams, rivers, and eventually large bodies such as lakes or the ocean. Infiltrated
groundwater moves similarly as it recharges rivers and heads towards large body of water.
As both of these processes are happening, the power of the sun is driving this cycle by
causing evaporation. Evaporation is the change of liquid water to a vapor. Sunlight aids
this process as it raises the temperature of liquid water in oceans and lakes. As the liquid
heats, molecule are released and change into a gas. Warm air rises up into the
atmosphere and becomes the vapor involved in condensation.
Considering so little of the water on earth is drinkable to people, it is amazing the supply
has survived as long as it has. The hydrologic cycle continues to move water and keep
sources fresh. It is estimated that 100 million billion gallons a year are cycled through this
process. Without this process life on Earth would be impossible. We need it to sustain us
and for all of our life processes to function. Without water, life would not be possible on
Earth.

1.2

Application of Hydrology in Practice

Knowledge of hydrology is required in the following engineering purposes:

Precipitation and evaporation estimation for Agricultural use

Rainfall run off analysis for flood forecasting

River flow estimation

8
8/113

1.3

Flood and draught estimation

Flood frequency analysis

Climate change analysis

Hydrological modelling such as rainfall runoff modelling and ground water modelling

Questions

Q1. What is hydrologic cycle? Sketch and briefly outline its various components.
Q2. Write the entire process of the hydrologic cycle.
Q3. Write down various application of hydrology in practice.

9
9/113

2.1

HYDROLOGICAL PROCESSES

Precipitation and its Processes

In meteorology, precipitation (also known as one of the classes of hydrometeors, which are
atmospheric water phenomena) is any product of the condensation of atmospheric water
vapour that is deposited on the earth's surface. It occurs when the atmosphere, a large
gaseous solution, becomes saturated with water vapour and the water condenses and falls
out of solution (i.e., precipitates). Two processes, possibly acting together, can lead to air
becoming saturated: cooling the air or adding water vapour to the air. Virga is precipitation
that begins falling to the earth but evaporates before reaching the surface; it is one of the
ways air can become saturated. Precipitation forms via collision with other rain drops or ice
crystals within a cloud.
Rain drops range in size from oblate, pancake-like shapes for larger drops, to small
spheres for smaller drops. Rain drops are not shaped like tear drops, but rather like
flattened pancakes. Precipitation that reaches the surface of the earth can occur in many
different forms, including rain, freezing rain, drizzle, ice needles, snow, ice pellets or sleet,
graupel and hail. While snow and ice pellets require temperatures to be near or below
freezing, hail can occur during much warmer temperature regimes due to the process of its
formation. Precipitation may occur on other celestial bodies, e.g. when it gets cold, Mars
has precipitation which most likely takes the form of ice needles, rather than rain or snow.
Moisture overriding weather fronts is a major method of precipitation production. If enough
moisture and upward motion is present, precipitation falls from convective clouds such as
cumulonimbus and organize into narrow rainbands. Precipitation can also form due to
forced ascent up the windward side of a mountain or mountain range. On the leeward side
of mountains, desert climates can exist due to the dry air caused by compressional
heating. The movement of the monsoon trough, or Intertropical convergence zone, brings
rainy seasons to savannah climes. Precipitation is a major component of the water cycle,
and is responsible for depositing most of the fresh water on the planet. Approximately
505,000 km3 (121,000 cu mi) of water falls as precipitation each year, 398,000 km3 (95,000
cu mi) of it over the oceans. Given the Earth's surface area, that means the globallyaveraged annual precipitation is about 1 metre (39 in), and the average annual
precipitation over oceans is about 1.1 metres (43 in).
10
10/113

Precipitation can be divided into three categories, based on whether it falls as liquid water,
liquid water that freezes on contact with the surface, or ice. Mixtures of different types of
precipitation, including types in different categories, can fall simultaneously.
How air becomes saturated?
Air contains water vapour, measured in grams of water per kilogram of dry air (g/kg), but
most commonly reported as a relative humidity. How much water vapour a parcel of air can
contain before it becomes saturated (100% relative humidity) depends on its temperature.
Warmer air can contain more water vapour than cooler air before becoming saturated.
Therefore, one way to saturate a parcel of air is to cool it. The dew point is the temperature
to which a parcel must be cooled in order to become saturated. Water vapor normally
begins to condense on condensation nuclei such as dust, ice, and salt in order to form
clouds. An elevated portion of a frontal zone forces broad areas of lift, which form clouds
decks such as altostratus or cirrostratus. Stratus is a stable cloud deck which tends to form
when a cool, stable air mass is trapped underneath a warm air mass. It can also form due
to the lifting of advection fog during breezy conditions.
There are four main mechanisms for cooling the air to its dew point: adiabatic cooling,
conductive cooling, radiational cooling, and evaporative cooling. Adiabatic cooling occurs
when air rises and expands. The air can rise due to convection, large-scale atmospheric
motions, or a physical barrier such as a mountain (orographic lift). Conductive cooling
occurs when the air comes into contact with a colder surface, usually by being blown from
one surface to another, for example from a liquid water surface to colder land. Radiational
cooling occurs due to the emission of infrared radiation, either by the air or by the surface
underneath. Evaporative cooling occurs when moisture is added to the air through
evaporation, which forces the air temperature to cool to its wet bulb temperature, or until it
reaches saturation.

Adding moisture to the air


The main ways water vapour is added to the air are:

Wind convergence into areas of upward motion

Precipitation or virga falling from above

Daytime heating evaporating water from the surface of oceans, water bodies or wet
land

Transpiration from plants


11
11/113

Cool, dry air moving over warmer water

Lifting air over mountains

Formation

Condensation and coalescence are important parts of the water cycle

Coalescence and the Bergeron process


Coalescence occurs when water droplets fuse to create larger water droplets, or when
water droplets freeze onto an ice crystal, which is known as the Bergeron process. Air
resistance typically causes the water droplets in a cloud to remain stationary. When air
turbulence occurs, water droplets collide, producing larger droplets. As these larger water
droplets descend, coalescence continues, so that drops become heavy enough to
overcome air resistance and fall as rain. Coalescence generally happens most often in
clouds above freezing. In clouds below freezing, when ice crystals gain enough mass they
begin to fall. This generally requires more mass than coalescence when occurring between
the crystal and neighbouring water droplets. This process is temperature dependent, as
super-cooled water droplets only exist in a cloud that is below freezing. In addition,
because of the great temperature difference between cloud and ground level, these ice
crystals may melt as they fall and become rain.

Raindrop characteristics
Raindrops have sizes ranging from 0.1 millimetres (0.0039 in) to 9 millimetres (0.35 in)
mean diameter, above which they tend to break up. Smaller drops are called cloud
droplets, and their shape is spherical. As a raindrop increases in size, its shape becomes
more oblate, with its largest cross-section facing the oncoming airflow. Contrary to the
cartoon pictures of raindrops, their shape does not resemble a teardrop. Intensity and
duration of rainfall are usually inversely related, i.e., high intensity storms are likely to be of
short duration and low intensity storms can have a long duration. Rain drops associated
with melting hail tend to be larger than other rain drops.

12
12/113

Causes

Frontal activity
Stratiform or dynamic precipitation occurs as a consequence of slow ascent of air in
synoptic systems (on the order of cm/s), such as over surface cold fronts, and over and
ahead of warm fronts. Similar ascent is seen around tropical cyclones outside of the
eyewall, and in comma head precipitation patterns around mid-latitude cyclones. A wide
variety of weather can be found along an occluded front, with thunderstorms possible, but
usually their passage is associated with a drying of the air mass. Occluded fronts usually
form around mature low-pressure areas. Precipitation may occur on other celestial bodies
other than Earth. When it gets cold, Mars has precipitation which most likely takes the form
of ice needles, rather than rain or snow.

Convection

Fig. 2 Convective Precipitation


Convective rain, or showery precipitation, occurs from convective clouds, e.g.,
cumulonimbus or cumulus congestus. It falls as showers with rapidly changing intensity.
Convective precipitation falls over a certain area for a relatively short time, as convective
clouds have limited horizontal extent. Most precipitation in the tropics appears to be
convective; however, it has been suggested that stratiform precipitation also occurs.
Graupel and hail indicate convection. In mid-latitudes, convective precipitation in
intermittent and often associated with baroclinic boundaries such as cold fronts, squall
lines, and warm fronts.

13
13/113

Orographic effects

Fig. 3 Orographic Precipitation


Orographic precipitation occurs on the windward side of mountains and is caused by the
rising air motion of a large-scale flow of moist air across the mountain ridge, resulting in
adiabatic cooling and condensation. In mountainous parts of the world subjected to
relatively consistent winds (for example, the trade winds), a more moist climate usually
prevails on the windward side of a mountain than on the leeward or downwind side.
Moisture is removed by orographic lift, leaving drier air (see katabatic wind) on the
descending and generally warming, leeward side where a rain shadow is observed.
Within the tropics
The wet, or rainy, season is the time of year, covering one or more months, when most of
the average annual rainfall in a region falls. The term green season is also sometimes
used as a euphemism by tourist authorities. Areas with wet seasons are dispersed across
portions of the tropics and subtropics. Savannah climates and areas with monsoon
regimes have wet summers and dry winters. Tropical rainforests technically do not have
dry or wet seasons, since their rainfall is equally distributed through the year. Some areas
with pronounced rainy seasons will see a break in rainfall mid-season when the intertropical convergence zone or monsoon trough move poleward of their location during the
middle of the warm season. When the wet season occurs during the warm season, or
summer, rain falls mainly during the late afternoon and early evening hours. The wet
season is a time when air quality improves, freshwater quality improves, and vegetation
grows significantly. Soil nutrients diminish and erosion increases. Animals have adaptation
and survival strategies for the wetter regime. Unfortunately, the previous dry season leads
to food shortages into the wet season, as the crops have yet to mature. Developing
countries have noted that their populations show seasonal weight fluctuations due to food
shortages seen before the first harvest, which occurs late in the wet season.

14
14/113

Tropical cyclones, a source of very heavy rainfall, consist of large air masses several
hundred miles across with low pressure at the centre and with winds blowing inward
towards the centre in either a clockwise direction (southern hemisphere) or counterclockwise (northern hemisphere). Although cyclones can take an enormous toll in lives and
personal property, they may be important factors in the precipitation regimes of places they
impact, as they may bring much-needed precipitation to otherwise dry regions. Areas in
their path can receive a year's worth of rainfall from a tropical cyclone passage.

2.2

Rainfall Characteristics

Rainfall characteristics affect the amount of runoff which occurs, the severity of erosion
possible in various parts of the country, and our dependency on irrigation for crop growth.
Specific important characteristics of rainfall are:

Size and Shape.

Rainfall occurs when moisture in the atmosphere condenses into drops. Raindrops occur
about in any shape up to approximately 9 mm mean diameter after which they tend to
break up. However, they do tend, if turbulence does not interfere, toward an
aerodynamically stable shape (tear-drop) because this affords the least surface resistance
to movement.

Intensity and Duration.

These are usually inversely related, i.e., high intensity storms are likely to be of short
duration and low intensity storms can have a long duration.

Intensity and Area.

We can expect a less intense rainfall (amount also) over a large area than we can over a
small area.

Intensity and Drop Size.

High intensity storms have a larger drop size than low intensity storms.
15
15/113

Drop size and Terminal Velocity.

The terminal velocity of raindrops increases as the drop size increases up to about 35 feet
per second. Storms with large drop sizes have a high erosion potential.

Rainfall Distribution and Supply.

These are difficult to predict for a given season, but averages based on long-term records
tells us much about the kind of water management necessary for an area.

Rainfall Amount/Intensity

The most common historical rainfall data are daily total amounts at selected locations.
Such data have usually been obtained from a "standard raingage" (tube) which gives the
depth of rainfall accumulated between observations. More useful are data from a recording
raingage which gives a record of accumulation as a function of time. Hence, intensities can
be determined from recording gages. Traditional raingages use a mechanical weighing
mechanism, while modern electronic units tend to digitally record the time when each 1/4
mm accumulates.

2.3

2.3.1

Measurement of Precipitation

Point Precipitation Measurement

The standard way of measuring rainfall or snowfall is the standard rain gauge, which can
be found in 100-mm (4-in) plastic and 200-mm (8-in) metal varieties. The inner cylinder is
filled by 25 mm (1 in) of rain, with overflow flowing into the outer cylinder. Plastic gages will
have markings on the inner cylinder down to 0.25 mm (0.01 in) resolution, which metal
gages will require use of a stick designed with the appropriate 0.25 mm (0.01 in) markings.
After the inner cylinder is filled, the amount inside it is discarded, then filled with the
remaining rainfall in the outer cylinder until all the fluid in the outer cylinder is gone, adding
to the overall total until the outer cylinder is empty. These gages are winterized by
removing the funnel and inner cylinder and allowing the snow/freezing rain to collect inside
the outer cylinder. Some add anti-freeze to their gage so they do not have to melt the snow
16
16/113

or ice that falls into the gage. Once the snowfall/ice is finished accumulating, or as you
approach 300 mm (12 in), one can either bring it inside to melt, or use luke warm water to
fill the inner cylinder with in order to melt the frozen precipitation in the outer cylinder,
keeping track of the warm fluid added, which is subsequently subtracted from the overall
total once all the ice/snow is melted.

Fig. 4 Standard Rain Gauge


Other types of gages include the popular wedge gage (the cheapest rain gage and most
fragile), the tipping bucket rain gage, and the weighing rain gage. The wedge and tipping
bucket gages will have problems with snow. Attempts to compensate for snow/ice by
warming the tipping bucket meet with limited success, since snow may sublimate if the
gage is kept much above freezing. Weighing gages with antifreeze should do fine with
snow, but again, the funnel needs to be removed before the event begins. For those
looking to measure rainfall the most inexpensively, a can that is cylindrical with straight
sides will act as a rain gage if left out in the open, but its accuracy will depend on what
ruler you use to measure the rain with. Any of the above rain gages can be made at home,
with enough know-how.
Once someone has a device to measure precipitation, various networks exist across the
United States and elsewhere where rainfall measurements can be submitted through the
internet, such as CoCoRAHS or GLOBE. If a network is not available in the area where
one lives, the nearest local weather office will likely be interested in the measurement.

17
17/113

The Quantitative Precipitation Forecast (abbreviated QPF) is the expected amount of liquid
precipitation accumulated over a specified time period over a specified area. A QPF will be
specified when a measurable precipitation type reaching a minimum threshold is forecast
for any hour during a QPF valid period. Precipitation forecasts tend to be bound by
synoptic hours such as 0000, 0600, 1200 and 1800 GMT. Terrain is considered in QPFs by
use of topography or based upon climatological precipitation patterns from observations
with fine detail. Starting in the mid to late 1990's, QPFs were used within hydrologic
forecast models to simulate impact to rivers throughout the United States. Forecast models
show significant sensitivity to humidity levels within the planetary boundary layer, or in the
lowest levels of the atmosphere, which decreases with height. QPF can be generated on a
quantitative, forecasting amounts, or a qualitative, forecasting the probability of a specific
amount, basis. Radar imagery forecasting techniques show higher skill than model
forecasts within 6 to 7 hours of the time of the radar image. The forecasts can be verified
through use of rain gage measurements, weather radar estimates, or a combination of
both. Various skill scores can be determined to measure the value of the rainfall forecast.

2.3.2

Estimation of Areal Precipitation

A single point precipitation measurement is quite often not representative of the volume of
precipitation falling over a given catchment area. A dense network of point measurements
and/or radar estimates can provide a better representation of the true volume over a given
area. A network of precipitation measurements can be converted to areal estimates using
any of a number of techniques include the following:
1) Arithmetic Mean - This technique calculates areal precipitation using the arithmetic
mean of all the point or areal measurements considered in the analysis. Areal rainfall is
obtained from point data by an arithmetic or area-weighted average of the rainfall
amounts. The general formula to calculate area-weighted averages is:

(2.1)
where:
The cross-product of each "n" sub-area and its corresponding rainfall amount are summed
and then divided by the total area.
18
18/113

2) Isohyetal Analysis - This is a graphical technique which involves drawing estimated


lines of equal rainfall over an area based on point measurements. The magnitude and
extent of the resultant rainfall areas of coverage are then considered versus the area in
question in order to estimate the areal precipitation value. In Isohyetal Method, rainfall
amounts from a set of gages are plotted on a map of the region. Lines connecting all
points of equal precipitation are then connected to create a isohyetal map (analogous
to contour lines on a topographic map). Obviously, this works best when there are
numerous raingages. The area between each isohyetal line is then measured (using a
planimeter if done manually or a computer algorithm if automated with GIS software).
This is the more accurate of the two methods; however, it's laborious if done manually
because the isohyetal lines must be redrawn and remeasured for every storm event.
3) Thiessen Polygon - This is another graphical technique which calculates station
weights based on the relative areas of each measurement station in the Thiessen
polygon network. The individual weights are multiplied by the station observation and
the values are summed to obtain the areal average precipitation.
The Thiessen polygon network can be constructed from the Delauney triangulation of a
scatter point set. A Delauney triangulation is a TIN that has been constructed so that
the Delauney criterion has been satisfied.

Delauney Triangulation and Corresponding Thiessen Polygon Network for a Set of Scatter
Points.

19
19/113

Each Thiessen polygon is constructed using the circumcircles of the triangles resulting
from a Delauney triangulation of the scatter points. The vertices of the Thiessen polygons
correspond to the centroids of the circumcircles of the triangles. Thiessen Method. This
technique has the advantage of being quick to apply for multiple storms because it uses
fixed sub-areas. It is based on the hypothesis that, for every point in the area, the best
estimate of rainfall is the measurement physically closest to that point. This concept is
implemented by drawing perpendicular bisectors to straight lines connecting each two
raingages. This yields, when the watershed boundary is included, a set of closed areas
known as Thiessen polygons. An example with four gages, A - D, is shown below:

As a computational example, assume the areas enclosed by the four Thiessen polygons
are: A = 11 Ac.; B = 9 Ac.; C = 8 Ac.; and D = 15 Ac. These areas are, of course, constant
and need be measured only once. For a given storm, the gages recorded the following
precipitation amounts: A = 1.81 inches, B = 2.25 inches, C = 2.07 inches and D = 1.53
inches. Using the above weighted averaging formula, the Thiessen average precipitation
estimate for the entire watershed is:

In contrast, the simple arithmetic average of precipitations is 1.91 inches. Differences


between arithmetic and Thiessen averages increase for non-uniform storms when the
Thiessen areas differ widely.

20
20/113

4) Distance Weighting/Gridded - This is another station weighting technique. A grid of


point estimates is made based on a distance weighting scheme. Each observed point
value is given a unique weight for each grid point based on the distance from the grid
point in question. The grid point precipitation value is calculated based on the sum of
the individual station weight multiplied by observed station value. Once the grid points
have all been estimated they are summed and the sum is divided by the number of grid
points to obtain the areal average precipitation
5) MAPX - This is a NWS-specific gridded technique. Areal runoff zone precipitation
estimates are made using the 4 x 4 km 1-hourly gridded precipitation estimates. The
arithmetic mean calculation technique is used to average the grid point estimates.
6) Index Stations - In some areas of the country (primarily mountainous areas), predetermined station weights based on climatology are used to compute basin average
precipitation.
Areal Precipitation Terminology As Used In ABRFC Hydrologic Modelling
MAP - Mean Areal Precipitation - Areal runoff zone precipitation estimate normally based
on point precipitation observations. The distance weighting calculation technique is used.
MAP is used as input to the river forecast model on a routine basis.
MAPX - Radar Based Mean Areal Precipitation - Areal runoff zone precipitation estimate
based on the 4 x 4 km WSR-88D 1-hourly gridded precipitation estimates. The arithmetic
mean calculation technique is used to average the grid point estimates. MAPX is used as
input to the river forecast model on a routine basis.
FMAP - Future Mean Areal Precipitation - Future or forecast areal runoff zone precipitation
estimate. The Weather Forecast Offices (WFOs) develop precipitation forecasts based on
input from sources which may include meteorlogical model output, national guidance
products, local forecast procedures and individual forecaster experience. After weather
analysis is complete, the WFO forecaster uses a computer program to draw isohyets of
forecast precipitation and then the program performs an automated isohyetal analysis
calculation technique to convert to areal estimates. The forecast precipitation information is
generated for four 6-hour periods. The area of coverage is that of each WFOs area of
responsibility. The ABRFC Hydrometeorological Analysis and Support (HAS) function
mosaics the input from the WFOs so as to cover the entire ABRFC area of responsibility.

21
21/113

The HAS function also coordinates any required changes in the individual WFO QPF
information. FMAP is used as input to the river forecast model on a routine basis.

2.4

Evaporation and it its Estimation

Evaporation is the slow vaporization of a liquid and the reverse of condensation. A type of
phase transition, it is the process by which molecules in a liquid state (e.g. water)
spontaneously become gaseous (e.g. water vapor). Generally, evaporation can be seen by
the gradual disappearance of a liquid from a substance when exposed to a significant
volume of gas. Vaporization and evaporation however, are not entirely the same
processes. For example, substances like caesium, francium, gallium, bromine, rubidium
and mercury may vaporize, but they do not evaporate as such.
On average, the molecules in a glass of water do not have enough heat energy to escape
from the liquid, or else the liquid would turn into vapor quickly (boil). When the molecules
collide, they transfer energy to each other in varying degrees, based on how they collide.
Sometimes the transfer is so one-sided for a molecule near the surface that it ends up with
enough energy to escape.
Liquids that do not evaporate visibly at a given temperature in a given gas (e.g. cooking oil
at room temperature) have molecules that do not tend to transfer energy to each other in a
pattern sufficient to frequently give a molecule the heat energy necessary to turn into
vapor. However, these liquids are evaporating, it's just that the process is much slower and
thus significantly less visible.
Evaporation is an essential part of the water cycle. Solar energy drives evaporation of
water from oceans, lakes, moisture in the soil, and other sources of water.
In hydrology, evaporation and transpiration (which involves evaporation within plant
stomata) are collectively termed evapotranspiration. Evaporation is caused when water is
exposed to air and the liquid molecules turn into water vapor which rises up and forms
clouds.
Evaporation is the process by which a liquid becomes a gas. It usually occurs at the
surface of a liquid,

22
22/113

2.4.1

Factors Influencing the Rate of Evaporation

1)

Concentration of the substance evaporating in the air

If the air already has a high concentration of the substance evaporating, then the given
substance will evaporate more slowly.
2)

Concentration of other substances in the air

If the air is already saturated with other substances, it can have a lower capacity for the
substance evaporating.
3)

Concentration of other substances in the liquid (impurities)

If the liquid contains other substances, it will have a lower capacity for evaporation.

4)

Flow rate of air

This is in part related to the concentration points above. If fresh air is moving over the
substance all the time, then the concentration of the substance in the air is less likely to go
up with time, thus encouraging faster evaporation. This is the result of the boundary layer
at the evaporation surface decreasing with flow velocity, decreasing the diffusion distance
in the stagnant layer.

5)

Inter-molecular forces

The stronger the forces keeping the molecules together in the liquid state, the more energy
one must get to escape.

6)

Pressure

In an area of less pressure, evaporation happens faster because there is less exertion on
the surface keeping the molecules from launching themselves.

23
23/113

7)

Surface area

A substance which has a larger surface area will evaporate faster as there are more
surface molecules which are able to escape.

8)

Temperature of the substance

If the substance is hotter, then evaporation will be faster.


The actual rate of evaporation from a standardized "pan" open water surface outdoors, at
various locations nationwide. Others do likewise around the world. The US data is
collected and compiled into an annual evaporation map. The measurements range from
under 30 to over 120 inches (3,000 mm) per year

2.5
2.5.1

Evapo-Transpiration and its Estimation


Equations

A large number of more or less empirical methods have been developed over the last 50
years by numerous scientists and specialists worldwide to estimate evapo-transpiration
from different climatic variables. Relationships were often subject to rigorous local
calibrations and proved to have limited global validity. Testing the accuracy of the methods
under a new set of conditions is laborious, time-consuming and costly, and yet evapotranspiration data are frequently needed at short notice for project planning or irrigation
scheduling design. To meet this need, guidelines were developed and published in the
FAO Irrigation and Drainage Paper No. 24 'Crop water requirements'. To accommodate
users with different data availability, four methods were presented to calculate the
reference crop evapo-transpiration (ETo): the Blaney-Criddle, radiation, modified Penman
and pan evaporation methods. The modified Penman method was considered to offer the
best results with minimum possible error in relation to a living grass reference crop. It was
expected that the pan method would give acceptable estimates, depending on the location
of the pan. The radiation method was suggested for areas where available climatic data
include measured air temperature and sunshine, cloudiness or radiation, but not measured
wind speed and air humidity. Finally, the publication proposed the use of the BlaneyCriddle method for areas where available climatic data cover air temperature data only.

24
24/113

These climatic methods to calculate ETo were all calibrated for ten-day or monthly
calculations, not for daily or hourly calculations. The Blaney-Criddle method was
recommended for periods of one month or longer. For the pan method it was suggested
that calculations should be done for periods of ten days or longer. Users have not always
respected these conditions and calculations have often been done on daily time steps.
Advances in research and the more accurate assessment of crop water use have revealed
weaknesses in the methodologies. Numerous researchers analysed the performance of
the four methods for different locations. Although the results of such analyses could have
been influenced by site or measurement conditions or by bias in weather data collection, it
became evident that the proposed methods do not behave the same way in different
locations around the world. Deviations from computed to observed values were often found
to exceed ranges indicated by FAO. The modified Penman was frequently found to
overestimate ETo, even by up to 20% for low evaporative conditions. The other FAO
recommended equations showed variable adherence to the reference crop evapotranspiration standard of grass.
To evaluate the performance of these and other estimation procedures under different
climatological conditions, a major study was undertaken under the auspices of the
Committee on Irrigation Water Requirements of the American Society of Civil Engineers
(ASCE). The ASCE study analysed the performance of 20 different methods, using
detailed procedures to assess the validity of the methods compared to a set of carefully
screened lysimeter data from 11 locations with variable climatic conditions. The study
proved very revealing and showed the widely varying performance of the methods under
different climatic conditions. In a parallel study commissioned by the European Community,
a consortium of European research institutes evaluated the performance of various evapotranspiration methods using data from different lysimeter studies in Europe.
The studies confirm the overestimation of the modified Penman introduced in FAO
Irrigation and Drainage Paper No. 24, and the variable performance of the different
methods depending on their adaptation to local conditions. The comparative studies may
be summarized as follows:

25
25/113

The Penman methods may require local calibration of the wind function to achieve
satisfactory results.
The radiation methods show good results in humid climates where the aerodynamic term
is relatively small, but performance in arid conditions is erratic and tends to underestimate
evapo-transpiration.
Temperature methods remain empirical and require local calibration in order to achieve
satisfactory results. A possible exception is the 1985 Hargreaves' method which has
shown reasonable ETo results with a global validity.
Pan evapo-transpiration methods clearly reflect the shortcomings of predicting crop
evapo-transpiration from open water evaporation. The methods are susceptible to the
microclimatic conditions under which the pans are operating and the rigour of station
maintenance. Their performance proves erratic.
The relatively accurate and consistent performance of the Penman-Monteith approach in
both arid and humid climates has been indicated in both the ASCE and European studies.
The analysis of the performance of the various calculation methods reveals the need for
formulating a standard method for the computation of ETo. The FAO Penman-Monteith
method is recommended as the sole standard method. It is a method with strong likelihood
of correctly predicting ETo in a wide range of locations and climates and has provision for
application in data-short situations. The use of older FAO or other reference ET methods is
no longer encouraged.

2.5.2

Penman-Monteith Equation

In 1948, Penman 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.

26
26/113

The Penman-Monteith form of the combination equation is:

(2.2)
where Rn is the net radiation, G is the soil heat flux, (es - ea) represents the vapour
pressure deficit of the air, a is the mean air density at constant pressure, cp is the specific
heat of the air, represents the slope of the saturation vapour pressure temperature
relationship, is the psychrometric constant, and rs and ra are the (bulk) surface and
aerodynamic resistances. The Penman-Monteith approach as formulated above includes
all parameters that govern energy exchange and corresponding latent heat flux (evapotranspiration) from uniform expanses of vegetation. Most of the parameters are measured
or can be readily calculated from weather data. The equation can be utilized for the direct
calculation of any crop evapo-transpiration as the surface and aerodynamic resistances
are crop specific.
Aerodynamic resistance (ra)
The transfer of heat and water vapour from the evaporating surface into the air above the
canopy is determined by the aerodynamic resistance:

(2.3)
where
ra aerodynamic resistance [s m-1],
zm height of wind measurements [m],
zh height of humidity measurements [m],
d zero plane displacement height [m],
zom roughness length governing momentum transfer [m],
zoh roughness length governing transfer of heat and vapour [m],
k von Karman's constant, 0.41 [-],
uz wind speed at height z [m s-1].

27
27/113

The equation is restricted for neutral stability conditions, i.e., where temperature,
atmospheric pressure, and wind velocity distributions follow nearly adiabatic conditions (no
heat exchange). The application of the equation for short time periods (hourly or less) may
require the inclusion of corrections for stability. However, when predicting ETo in the wellwatered reference surface, heat exchanged is small, and therefore stability correction is
normally not required.
Many studies have explored the nature of the wind regime in plant canopies. Zero
displacement heights and roughness lengths have to be considered when the surface is
covered by vegetation. The factors depend upon the crop height and architecture. Several
empirical equations for the estimate of d, zom and zoh have been developed. The
derivation of the aerodynamic resistance for the grass reference surface is presented in
Box 1.
(Bulk) surface resistance (rs)
The 'bulk' surface resistance describes the resistance of vapour flow through the
transpiring crop and evaporating soil surface. Where the vegetation does not completely
cover the soil, the resistance factor should indeed include the effects of the evaporation
from the soil surface. If the crop is not transpiring at a potential rate, the resistance
depends also on the water status of the vegetation. An acceptable approximation to a
much more complex relation of the surface resistance of dense full cover vegetation is:
BOX 1. The aerodynamic resistance for a grass reference surface
For a wide range of crops the zero plane displacement height, d [m], and the roughness
length governing momentum transfer, zom [m], can be estimated from the crop height h [m]
by the following equations:
d = 2/3 h
zom = 0.123 h
The roughness length governing transfer of heat and vapour, zoh [m], can be
approximated by:
zoh = 0.1 zom
Assuming a constant crop height of 0.12 m and a standardized height for wind speed,
temperature and humidity at 2 m (zm = zh = 2 m), the aerodynamic resistance ra [s m-1]
for the grass reference surface becomes (Eq. 4):
28
28/113

where u2 is the wind speed [m s-1] at 2 m.

(2.4)
where
rs (bulk) surface resistance [s m-1],
rl bulk stomatal resistance of the well-illuminated leaf [s m-1],
LAIactive active (sunlit) leaf area index [m2 (leaf area) m-2 (soil surface)].
The Leaf Area Index (LAI), a dimensionless quantity, is the leaf area (upper side only) per
unit area of soil below it. It is expressed as m2 leaf area per m2 ground area. The active
LAI is the index of the leaf area that actively contributes to the surface heat and vapour
transfer. It is generally the upper, sunlit portion of a dense canopy. The LAI values for
various crops differ widely but values of 3-5 are common for many mature crops. For a
given crop, green LAI changes throughout the season and normally reaches its maximum
before or at flowering (Figure 5). LAI further depends on the plant density and the crop
variety.
The bulk stomatal resistance, rl, is the average resistance of an individual leaf. This
resistance is crop specific and differs among crop varieties and crop management. It
usually increases as the crop ages and begins to ripen. There is, however, a lack of
consolidated information on changes in rl over time for the different crops. The information
available in the literature on stomatal conductance or resistance is often oriented toward
physiological or eco-physiological studies.

29
29/113

Fig. 5. Typical presentation of the variation in the active (green) Leaf Area Index over
the growing season for a maize crop
The stomatal resistance, rl, is influenced by climate and by water availability. However,
influences vary from one crop to another and different varieties can be affected differently.
The resistance increases when the crop is water stressed and the soil water availability
limits crop evapo-transpiration. Some studies indicate that stomatal resistance is
influenced to some extent by radiation intensity, temperature, and vapour pressure deficit.
The derivation of the surface resistance for the grass reference surface is presented in Box
2.
BOX 2: The (bulk) surface resistance for a grass reference crop
A general equation for LAIactive is:
LAIactive = 0.5 LAI
which takes into consideration the fact that generally only the upper half of dense clipped
grass is actively contributing to the surface heat and vapour transfer. For clipped grass a
general equation for LAI is:
LAI = 24 h
where h is the crop height [m].

30
30/113

The stomatal resistance, rl, of a single leaf has a value of about 100 s m-1 under wellwatered conditions. By assuming a crop height of 0.12 m, the surface resistance, rs [s m1], for the grass reference surface becomes (Eq. 2.3):

Reference surface
To obviate the need to define unique evaporation parameters for each crop and stage of
growth, the concept of a reference surface was introduced. Evapo-transpiration rates of the
various crops are related to the evapo-transpiration rate from the reference surface (ETo)
by means of crop coefficients.
In the past, an open water surface has been proposed as a reference surface. However,
the differences in aerodynamic, vegetation control and radiation characteristics present a
strong challenge in relating ET to measurements of free water evaporation. Relating ETo to
a specific crop has the advantage of incorporating the biological and physical processes
involved in ET from cropped surfaces.
Grass, together with alfalfa, is a well-studied crop regarding its aerodynamic and surface
characteristics and is accepted worldwide as a reference surface. Because the resistance
to diffusion of vapour strongly depends on crop height, ground cover, LAI and soil moisture
conditions, the characteristics of the reference crop should be well defined and fixed.
Changes in crop height result in variations in the roughness and LAI. Consequently, the
associated canopy and aerodynamic resistances will vary appreciably with time. Moreover,
water stress and the degree of ground cover have an effect on the resistances and also on
the albedo.
The FAO Expert Consultation on Revision of FAO Methodologies for Crop Water
Requirements accepted the following unambiguous definition for the reference surface:
"A hypothetical reference crop with an assumed crop height of 0.12 m, a fixed surface
resistance of 70 s m-1 and an albedo of 0.23."

31
31/113

The reference surface closely resembles an extensive surface of green grass of uniform
height, actively growing, completely shading the ground and with adequate water. The
requirements that the grass surface should be extensive and uniform result from the
assumption that all fluxes are one-dimensional upwards.
The FAO Penman-Monteith method is selected as the method by which the evapotranspiration of this reference surface (ETo) can be unambiguously determined, and as the
method which provides consistent ETo values in all regions and climates.

2.5.3

FAO Penman-Monteith Equation

The panel of experts recommended the adoption of the Penman-Monteith combination


method as a new standard for reference evapo-transpiration and advised on procedures
for calculation of the various parameters. By defining the reference crop as a hypothetical
crop with an assumed height of 0.12 m having a surface resistance of 70 s m-1 and an
albedo of 0.23, closely resembling the evaporation of an extension surface of green grass
of uniform height, actively growing and adequately watered, the FAO Penman-Monteith
method was developed. The method overcomes shortcomings of the previous FAO
Penman method and provides values more consistent with actual crop water use data
worldwide.

Fig. 6. Characteristics of the hypothetical reference crop


32
32/113

From the original Penman-Monteith equation (Equation 2.3) and the equations of the
aerodynamic (Equation 4) and surface resistance (Equation 2.4), the FAO PenmanMonteith method to estimate ETo can be derived (Box 6):

(2.5)
where
ETo reference evapo-transpiration [mm day-1],
Rn net radiation at the crop surface [MJ m-2 day-1],
G soil heat flux density [MJ m-2 day-1],
T mean daily air temperature at 2 m height [C],
u2 wind speed at 2 m height [m s-1],
es saturation vapour pressure [kPa],
ea actual vapour pressure [kPa],
es - ea saturation vapour pressure deficit [kPa],
slope vapour pressure curve [kPa C-1],
psychrometric constant [kPa C-1].
The reference evapo-transpiration, ETo, provides a standard to which:
evapo-transpiration at different periods of the year or in other regions can be compared;
evapo-transpiration of other crops can be related.
The equation uses standard climatological records of solar radiation (sunshine), air
temperature, humidity and wind speed. To ensure the integrity of computations, the
weather measurements should be made at 2 m (or converted to that height) above an
extensive surface of green grass, shading the ground and not short of water.
Data
Apart from the site location, the FAO Penman-Monteith equation requires air temperature,
humidity, radiation and wind speed data for daily, weekly, ten-day or monthly calculations.
Location

33
33/113

Altitude above sea level (m) and latitude (degrees north or south) of the location should be
specified. These data are needed to adjust some weather parameters for the local average
value of atmospheric pressure (a function of the site elevation above mean sea level) and
to compute extraterrestrial radiation (Ra) and, in some cases, daylight hours (N). In the
calculation procedures for Ra and N, the latitude is expressed in radian (i.e., decimal
degrees times

/180).

BOX 3 Derivation of the FAO Penman-Monteith equation for the hypothetical grass
reference crop
With standardized height for wind speed, temperature and humidity measurements at 2 m
(zm = zh = 2 m) and the crop height h = 0.12 m, the aerodynamic and surface resistances
become (Boxes 4 & 5):
ra = 208/u2 s m-1, (with u2 wind speed at 2 m height)
rs = 70 s m-1
(1 + rs/ra) = (1 + 0.34 u2)
Rn and G is energy available per unit area and expressed in MJ m-2 day-1. To convert the
energy units for radiation to equivalent water depths (mm) the latent heat of vaporization,
is used as a conversion factor. The conversion from energy values to equivalent depths of
water or vice versa is given by (Eq. 20):

By substituting cp with a rearrangement of Eq. 8:

and considering the ideal gas law for

a:

where TKv the virtual temperature, may be substituted by:


TKv = 1.01(T+273)
results in:

[MJ m-2 C-1 day-1]

34
34/113

where
cp specific heat at constant pressure [MJ kg-1 C-1],
a mean air density at constant pressure [kg m-3],
ra aerodynamic resistance [s m-1],
psychrometric constant [kPa C-1],
ratio molecular weight of water vapour/dry air = 0.622,
latent heat of vaporization [MJ kg-1],
u2 wind speed at 2 m [m s-1],
R specific gas constant = 0.287 kJ kg-1 K-1,
T air temperature [C],
P atmospheric pressure [kPa],

[MJ m-2 C-1 day-1]


or, when divided by

= 2.45),

[mm C-1 day-1]


A positive value is used for the northern hemisphere and a negative value for the southern
hemisphere.
Temperature
The (average) daily maximum and minimum air temperatures in degrees Celsius (C) are
required. Where only (average) mean daily temperatures are available, the calculations
can still be executed but some underestimation of ETo will probably occur due to the nonlinearity of the saturation vapour pressure - temperature relationship (Figure 11). Using
mean air temperature instead of maximum and minimum air temperatures yields a lower
saturation vapour pressure es, and hence a lower vapour pressure difference (es - ea),
and a lower reference evapo-transpiration estimate.
Humidity
The (average) daily actual vapour pressure, ea, in kilopascals (kPa) is required. The actual
vapour pressure, where not available, can be derived from maximum and minimum relative
humidity (%), psychrometric data (dry and wet bulb temperatures in C) or dewpoint
temperature (C).

35
35/113

Radiation
The (average) daily net radiation expressed in megajoules per square metre per day (MJ
m-2 day-1) is required. These data are not commonly available but can be derived from the
(average) shortwave radiation measured with a pyranometer or from the (average) daily
actual duration of bright sunshine (hours per day) measured with a (Campbell-Stokes)
sunshine recorder.
Wind speed
The (average) daily wind speed in meters per second (m s-1) measured at 2 m above the
ground level is required. It is important to verify the height at which wind speed is
measured, as wind speeds measured at different heights above the soil surface differ.
Missing climatic data
Situations might occur where data for some weather variables are missing. The use of an
alternative ETo calculation procedure, requiring only limited meteorological parameters,
should generally be avoided. It is recommended that one calculate ETo using the standard
FAO Penman-Monteith method after resolving the specific problem of the missing data.
Differences between ETo values obtained with the FAO Penman-Monteith equation with,
on the one hand, a limited data set and, on the other hand, a full data set, are expected to
be smaller than or of similar magnitude to the differences resulting from the use of an
alternative ETo equation.

2.6

Possible Questions

Q1. How precipitation is formed?


Q2. Define the entire process of formation of precipitation.
Q3. How precipitation can me measured?
Q3. What are the methods of areal precipitation measurement?
Q5. Area of catchment like a circle of diameter 100 km. Given below the data of five gauge
points, Estimate the mean areal precipitation by thiessen polygon method.

36
36/113

Station

Coordinates

(30,80)

(70,100)

(100,140)

(130,100)

(100,70

132.2

96.2

145.5

101.5

Precipitation(cm) 80.5

Coordinate of centre is (100,100). Answer: 120 cm


Q6. Long-term observations at a streamflow-measuring station at the outlet of a catchment
in a mountainous area gives a mean annual discharge of 65 m3/s. An isohyetal map for the
annual rainfall over the catchment gives the following areas closed by isohyets and the
divide of the catchment:
Isohyete

Area (km2)

Isohet (cm)

Area (km2)

140-135

50

120-115

600

135-130

300

115-110

400

130-125

450

110-105

200

125-120

700

Calculate (a) the mean annual depth of rainfall over the catchment, (b) the mean annual
runoff.
Q7. Define evaporation and discuss its mechanism.
Q8. How evapo-transpiration can be estimated? Name two well known method of Evapotraspration.

37
37/113

STREAM FLOW MEASUREMENT

3.1

Introduction

One can ask question "How much water is flowing in this river?" Answer is, it is a process
involving two concepts:

Stream stage

Streamflow

Streamflow (Discharge): The rate of water flow (volume/unit time) passing a given cross
section of a stream. Some common units include:
cubic feet per second, cfs, ft3/s
cubic meters per second, m3/s
gallons per minute, gpm
Stream gaging: Two types of field measurements is the basis for all streamflow work: (1)
river stage (water surface elevation) and (2) cross-sectional area.
The location along the river where these measurements are taken is referred to as a
gaging site. A permanent facility at this site is referred to as a gaging station. Determination
of river discharge requires that the velocity and cross-section area be measured at the
station in some systematic manner. This process is referred to as stream gaging.
Therefore, flow through a stream can be measured by direct measurement. Also some
methods are available to measure stream flow in ungaged streams.

Velocity Meters

Acoustic Doppler Current Profiler (ADCP)

Weirs and Flumes

Dilution Gaging

38
38/113

For a gaged streams flow are usually measured by the area- velocity method. The velocity
is measured using current meter.

3.2

Measurement of Stage and Velocity

Stream stage (also called stage or gage height) is the height of the water surface, in feet,
above an established datum plane where the stage is zero. The zero level is arbitrary, but
is often close to the streambed. You can get an idea of what stream stage is by looking at
this picture of a common staff gage, which is used to make a visual reading of stream
stage. The gage is marked in 1/100th and 1/10th foot intervals.
Streamflow, or discharge, is the volume of
water flowing past a fixed point in a fixed
unit of time. For water flow in streams, the
U.S. Geological Survey expresses the
value in cubic feet per second (ft3/s). For
example, when rain has not fallen for a
while, Peachtree Creek often is at a
baseflow stage of about 3 feet. The rating curve (see chart below) shows that at a stage of
3 feet streamflow is 76 ft3/s. Since one cubic foot of water contains 7.48 gallons, it might be
easier to understand this streamflow value if you consider that 76 ft3/s is about 568 gallons
of water flowing each second.

3.3

Stage-Discharge Relation: Rating Curve

Draw a graph with an x-axis and y-axis. The x-axis, the horizontal line, will be the
streamflow measurement. The y-axis, the vertical line, will be the staff gage reading. Place
a dot on the graph where each streamflow and corresponding staff gage measurement
intersect. Draw a smooth, curved line between the points. Now you have a stage-discharge
relationship. From now on, you can simply take the gage reading and estimate the stream
flow from your prediction curve.

39
39/113

As convenient as a stage-discharge relationship is, it still needs to be supported by real


data. The more data points you use to develop your graph, the better. The graph is
accurate only for the stream flows that fall within the data range you used to create the
graph. For example, if all your measurements were taken during June through September
when stream flows were low, the graph could not be used to predict high flows in
December. Be sure to collect data during a wide range of flow conditions. In general, if you
have about four data sets from the low-flow period and four from the high-flow period, you
can comfortably prepare the graph. Make periodic checks of the discharge curve,
especially after periods of flooding. Recalibrate the curve if the periodic checks indicate the
relationship has changed. Eventually, natural changes in the stream bottom will result in a
change in the relationship between flow and gage height.

Fig. 7 Stage vs streamflow curve

This chart, Figure 7, known as a rating curve, shows that there is a relation between
stream stage and streamflow. The stage-streamflow relation is used to relate water level to
an associated streamflow. The rating curve for a specific stream location is developed by
making successive steamflow measurements at many different stream stages to define
and maintain a stage-streamflow relation. These steamflow measurements and their
corresponding stages are then plotted on a graph. Continuous streamflow throughout the
year can be determined from the rating curve and the record of river stage.
The rating curve is crucial because it allows the use of stream stage, which is usually
easily determined, to estimate the corresponding streamflow at virtually any stream stage.

40
40/113

3.3.1

Relation of Stream Stage to Streamflow is Always Changing

Rating curves are not static - they occasionally must be recalculated. Rating curves
frequently shift due to changes in the factors that determine the relation between stream
stage and streamflow. These factors are:

Slope of the stream (affects velocity)

Roughness of the channel

Area of the channel at each stream stage

Backwater effects (when a tributary enters a larger river)

Filling in, scouring out, channel changes of river banks

Consider what can happen to a stream channel during a large flood. The Figure 8 below
shows a streambed before and after a flood, thus changing the relation between the
stream stage, in feet, and the amount of water flowing at that stage. The colored area
represents how much water is flowing. Both diagrams show the same 5-foot stage, but
more water is flowing after the flood because the streambed profile has changed and now
there is more area for water to flow. Scouring occurs more often on the outside edge of a
curve in a stream, whereas sand buildup occurs on the inside edge.

Fig. 8 Gage reading before and after bed scour

41
41/113

3.3.2

Method of Estimation of Flow by Direct Measurement

In order to accurately determine streamflow, measurements must be made of its width,


depth, and speed (velocity) of the water at many horizontal and vertical points across the
stream. To develop a stream-stage/streamflow relation (rating curve), streamflow must be
measured at many different stages. The well-developed rating curve allows for estimation
of streamflows at virtually any stream stage. More simply, if a stream is measured at
stages of 3.5, 6, 7.1, 9, and 10.2 feet, then an estimate can be made for a streamflow at 8
feet -- that is the goal.

Fig. 9 Current meter for velocity measurement


The stream-measurement procedure is to go across the stream at selected intervals and
measure the total depth and the velocity of the water at selected depths at each interval
across the stream. The picture shows a current meter (attached above the torpedo-looking
weight), which is lowered into the stream and measures water velocity. The spinning cups
on the current meter measure velocity.
In the diagram, the hydrologist would take a measurement of how fast the water is moving
at every green 'X', and would then determine the areas between all of the measured
intervals, such as the one shown by the purple box.

42
42/113

Fig. 10
In the diagram above, water depth/velocity measurements are obtained horizontally across
the stream at 1, 3, 5, 7, and 9 feet (the vertical lines in the diagram). At each location,
measurements of velocity and total depth are obtained. Depending on the depth and flow
conditions, one or more velocity reading(s) are obtained in each vertical. For our example,
a water depth/velocity measurement is obtained at a point 5 feet from the edge of the
stream. The total depth is slightly more than 3 feet and velocity readings are obtained at
depths of 1, 2, and 3 feet (the 'X's on the 5-foot vertical line). The purple box represents an
area that is midway between this measurement point and the measurement points on
either side. The purple area is 2 feet across and one foot high, or 2 square feet. The
measured velocity at the big X in the purple box is is 2 feet per second. To compute the
amount of water flowing in that purple area each second, multiply the area of the purple
box times the velocity of the water:
(1) 2 feet wide x 1 foot high = 2 square feet
(2) 2 square feet x 2 feet per second = 4 cubic feet per second.
To compute the total stream streamflow the hydrologist has to create imaginary purple
boxes between all of the 'X's and, using the velocity of the water in every box, compute the
streamflow for each purple area. Summing the streamflows for all the purple areas will give
the total streamflow. Actually, the example above is a simplified explanation of how
streamflow is measured. When an actual measurement is made, the hydrologist takes
measurements at about 20 points across the stream. The goal is to have no one vertical
cross-section contain more than 5 percent of the total stream discharge.

43
43/113

3.4

Questions

Q1. Why do we measure streamflow?


Q2. How do we measure streamflow?
Q3. Relating streamflow and stage?
Q4. The following data were collected during a stream-gauging operation in a river.
Compute the discharge.
Distance from left water

Depth (m)

Velocity (m/s)

edge (m)

At 0.2 d

At 0.8 d

0.0

0.0

0.0

0.0

1.5

1.3

0.6

0.4

3.0

2.5

0.9

0.6

4.5

1.7

0.7

0.5

6.0

1.0

0.6

0.4

7.5

0.4

0.4

0.3

9.0

0.0

0.0

0.0

Q5. The following are the data obtained in a stream-gauging operation. A current meter
with a calibration equation V = (0.32 N+0.032) m/s where N + revolutions per second was
used to measure the velocity at 0.6 depth. Using the mid-section method, calculate the
discharge in the stream.
Distance

12

15

18

20

22

23

24

0.50

1.10

1.95

2.25

1.85

1.75

1.65

1.50

1.25

0.75

80

83

131

139

121

114

109

92

85

70

180

120

120

120

120

120

120

120

120

150

from right
bank (m)
Depth
(m)
Number
of revolution
Time (s)

44
44/113

Q6. The following are the coordinates of a smooth curve drawn to best represent the
stage-discharge data of a river., Draw the stage-Discharge Curve(Rating curve)
Stage (m)

20.80

Discharge 100

21.42

21.95

22.37

23.00

23.52

23.90

200

300

400

600

800

1000

(m3/s)

45
45/113

4.1

HYDROGRAPH ANALYSIS

Introduction

The generated runoff hydrograph is a continuous integration of all factors that affect runoff
flow. The complexity and interaction of site factors on runoff and infiltration processes
makes it difficult to identify a single component of the hydrograph that accurately
characterizes the entire runoff event. A technique was developed to separate the runoff
hydrograph into segments representative of different portions of the flow event. Each
segment grouping is analyzed for treatment and/or site factor differences or influences on
the runoff. Comparing the treatment or site impacts on each hydrograph component allows
a more detailed interpretation of the runoff and infiltration processes. This approach to
runoff hydrograph analysis makes it possible to quantitatively assess differences in rainfall
simulator runoff results and provide insight into why hydrographs may be similar or
different.

4.2

Hydrograph and its Components

There are two meanings for hydrographs both coming from hydro- meaning water, and graph meaning chart. A hydrograph plots the discharge of a river as a function of time.

Fig. 11 A typical Hydrograph

46
46/113

My rising limb has taken a long time to get there, slightly less than 2 years. I hope that my
peak discharge spans from the 2nd of November till the 16th and there is a high discharge.
And I can picture my receding limb being a steep drop. At the end of it all, I dont suppose
my base flow would increase that much.
The discharge is measured at a certain point in a river and is typically time variant.

Rising limb - The part of the hydrograph up to the point of peak discharge.

Falling limb - The part of the hydrograph after the peak discharge.

Peak discharge - The highest point on the hydrograph when there is the greatest
amount of water in the river.

Lag time - Period of time between peak rainfall and peak discharge.

Stream hydrograph. Increases in stream flow follow rainfall or snowmelt. The gradual
decay in flow after the peaks reflects diminishing supply from groundwater.
Types of hydrograph can include:

Storm hydrographs

Flood hydrographs

Annual hydrographs

Surface water hydrograph


In surface water hydrology, a hydrograph is a time record of the discharge of a stream,
river or watershed outlet. Rainfall is typically the main input to a watershed and the stream
flow is often considered the output of the watershed; a hydrograph is a representation of
how a watershed responds to rainfall. They are used in hydrology and water resources
planning.

47
47/113

4.2.1

Factors Affecting Shape of Hydrograph

A watershed's response to rainfall depends on a variety of factors which affect the shape of
a hydrograph:

Watershed topography and geology (i.e. bedrock permeability)

The area of a basin receiving rainfall

Land-use (e.g. agriculture, urban development, forestry operations)

Drainage density

Duration of rainfall and precipitation intensity and type

Evapo-transpiration rates

River geometrics

The season

Previous weather

Vegetation type and cover

River conditions (e.g. dams)

Initial conditions (e.g. the degree of saturation of the soil and aquifers)

Soil permeability and thickness

A hydrograph is often compared to a hyetograph of the watershed.

4.3

Flow Duration Curve

The flow duration curve is a plot that shows the percentage of time that flow in a stream is
likely to equal or exceed some specified value of interest. For example, it can be used to
show the percentage of time river flow can be expected to exceed a design flow of some
specified value (e.g., 20 cfs), or to show the discharge of the stream that occurs or is
exceeded some percent of the time (e.g., 80% of the time).
The basic time unit used in preparing a flow-duration curve will greatly affect its
appearance. For most studies, mean daily discharges are used. These will give a steep
curve. When the mean flow over a long period is used (such as mean monthly flow), the
resulting curve will be flatter due to averaging of short-term peaks with intervening smaller
flows during a month. Extreme values are averaged out more and more, as the time period
gets larger (e.g., for a flow duration curve based on annual flows at a long-record station).
48
48/113

Step 1: Sort (rank) average daily discharges for period of record from the largest value to
the smallest value, involving a total of n values.
Step 2: Assign each discharge value a rank (M), starting with 1 for the largest daily
discharge value.
Step 3: Calculate exceedence probability (P) as follows:
P = 100 * [ M / (n + 1) ]
P = the probability that a given flow will be equalled or exceeded (% of time)
M = the ranked position on the listing (dimensionless)
n = the number of events for period of record (dimensionless)
What kind of graphs and charts can be generated?
A flow duration curve is a plot of discharge vs. percent of time that a particular discharge
was equalled or exceeded. The area under the flow duration curve (with arithmetic scales)
gives the average daily flow, and the median daily flow is the 50% value. It is useful to
graph the data on probability paper. Graphing software programs, such as Sigma Plot,
have the capability to plot on a log vs. probability scale.
What does this particular information tell you about your stream?
A flow duration curve characterizes the ability of the basin to provide flows of various
magnitudes. Information concerning the relative amount of time that flows past a site are
likely to equal or exceed a specified value of interest is extremely useful for the design of
structures on a stream. For example, a structure can be designed to perform well within
some range of flows, such as flows that occur between 20 and 80% of the time (or some
other selected interval).
The shape of a flow-duration curve in its upper and lower regions is particularly significant
in evaluating the stream and basin characteristics. The shape of the curve in the high-flow
region indicates the type of flood regime the basin is likely to have, whereas, the shape of
the low-flow region characterizes the ability of the basin to sustain low flows during dry
seasons. A very steep curve (high flows for short periods) would be expected for raincaused floods on small watersheds. Snowmelt floods, which last for several days, or
regulation of floods with reservoir storage, will generally result in a much flatter curve near
49
49/113

the upper limit. In the low-flow region, an intermittent stream would exhibit periods of no
flow, whereas, a very flat curve indicates that moderate flows are sustained throughout the
year due to natural or artificial streamflow regulation, or due to a large groundwater
capacity which sustains the base flow to the stream.

4.4

Effective Rainfall

Apart from soil, air and sunlight, crops need water to grow. This water can be supplied to
the crops by rainfall (also called precipitation), by irrigation or by a combination of rainfall
and irrigation.
If the rainfall is sufficient to cover the water needs of the crops, irrigation is not required.
If there is no rainfall, all the water that the crops need has to be supplied by irrigation.
If there is some rainfall, but not enough to cover the water needs of the crops, irrigation
water has to supplement the rain water in such a way that the rain water and the irrigation
water together cover the water needs of the crop. This is often called supplemental
irrigation: the irrigation water supplements or adds to the rain water.
Part of the rain water percolates below the root zone of the plants and part of the rain water
flows away over the soil surface as run-off (Fig. 12). This deep percolation water and runoff water cannot be used by the plants. In other words, part of the rainfall is not effective.
The remaining part is stored in the root zone and can be used by the plants. This
remaining part is the so-called effective rainfall. The factors which influence which part is
effective and which part is not effective include the climate, the soil texture, the soil
structure and the depth of the root zone.

Fig. 12 Part of the rain water is lost through deep percolation and run-off

50
50/113

If the rainfall is high, a relatively large part of the water is lost through deep percolation and
run-off.
Deep percolation: If the soil is still wet when the next rain occurs, the soil will simply not be
able to store more water, and the rain water will thus percolate below the root zone and
eventually reach the groundwater. Heavy rainfall may cause the groundwater table to rise
temporarily.
Run-off: Especially in sloping areas, heavy rainfall will result in a large percentage of the
rainwater being lost by surface run-off.
Another factor which needs to be taken into account when estimating the effective rainfall
is the variation of the rainfall over the years. Especially in low rainfall climates, the little rain
that falls is often unreliable; one year may be relatively dry and another year may be
relatively wet.
In many countries, formulae have been developed locally to determine the effective
precipitation. Such formulae take into account factors like rainfall reliability, topography,
prevailing soil type etc. If such formulae or other local data are available, they should be
used.
If such data are not available. Table 1 could be used to obtain a rough estimate of the
effective rainfall.
Table 1 Rainfall or precipitation P) and effective rainfall or effective precipitation (Pe) in
mm/month
P
(mm/month)
0
10
20
30
40
50
60
70
80
90
100
110
120

Pe
(mm/month)
0
0
2
8
14
20
26
32
39
47
55
63
71

P
(mm/month)
130
140
150
160
170
180
190
200
210
220
230
240
250

Pe
(mm/month)
79
87
95
103
111
119
127
135
143
151
159
167
175
51
51/113

Example 1:
Estimate the effective rainfall in mm/month if the rainfall is 60 mm/month. From Table 6 it
can be seen that the effective rainfall is 26 mm/month. This means that out of 60
mm/month, some 26 mm can be used by the plants; and it is estimated that the remaining
(60 - 26 =) 34 mm is lost through deep percolation and run-off.
Question
Determine the effective rainfall for the following monthly rainfall figures: 65, 210, 175 and 5
mm.
ANSWER (see Table 6)

P (mm/month) Pe (mm/month)
65
29
210
143
175
115
5
0
Example 2:
The effective rainfall hyetograph is calculated as in the following Table.
Interval
Rainfall depth (cm)
Loss @ 0.25 cm/h
For 6 h
Effective rainfall (cm)

4.5

4.5.1

1st 6 hours
3.5
1.5

2nd 6 hours
7.5
1.5

3rd 6 hours
5.5
1.5

2.0

6.0

4.0

Unit Hydrograph

Principles and Definitions

The unit hydrograph (UH) of a catchment is defined as the hydrograph resulting from an
effective rainfall of 1mm evenly distributed over the basin during the time D. The following
main characteristics of the unit hydrograph are considered (Figure 13):
52
52/113

Peak discharge of the unit hydrograph, up;


Base time tb is the total duration of the unit hydrograph;
Increase time or time to peak tp is the time between the start point of the hydrograph and
the peak;
Concentration time tc is the time between the end of rainfall and the end of the hydrograph;
Lag time tlag is the time between the gravity centre of rainfall and the peak of the
hydrograph.

Fig. 13 Characteristics of the unit hydrograph


The following hypotheses are considered while the linear unit hydrograph is used as a
transfer function:
The unit hydrograph reflects the ensemble of the physical characteristics of the river basin.
The effective rainfall is uniformly distributed over the catchment.
The shape characteristics of the unit hydrograph are independent of time. Therefore
duration of the unit hydrograph is constant regardless the effective rainfall intensity.

53
53/113

The response of the catchment to effective rainfall is linear: On the one hand, for a
particular reference duration of effective rainfall, the ordinates of the hydrograph are
proportional to rainfall. In other words, for a certain reference duration of an effective
rainfall having the amount h, the ordinate of the catchment hydrograph at the time t is
h.u(t), where u(t) is the ordinate of the unit hydrograph at time t (Figure 14). This property
is called proportionality.

Fig. 14 Linearity hypothesis of the unit hydrograph


On the other hand, for an effective rainfall composed of n segments of equal reference
duration, each of them having different amounts h(t), the resulting hydrograph is obtained
by summation of partial hydrographs that are shifted with the reference duration as shown
in Figure 15.

Fig. 15 Principle of superposition


54
54/113

This principle, used in the case of composed rainfall, is named the superposition or the
additivity principle and the procedure for assessing the hydrograph generated by a
composite effective rainfall is known as the convolution method. Proportionality and
additivity give to the unit hydrograph the attribute of linearity. According to this principle of
computation the total hydrograph produced by a composite effective rainfall having for
example three segments of h1, h2, and h3, is obtained as shown in Table 2. Eight
ordinates u1, u2, u3,.. u8 of the unit hydrograph, each of them corresponding to the
abscissa 1, 2, 3,..., 8 have been considered. According to the proportionality principle,
in the first line of Table 2 the hydrograph produced by the first segment of effective rainfall
is found. Then the hydrograph produced by the second segment of rainfall that is shifted
with a time interval (because the second segment of effective rainfall starts from the
beginning of 2). The procedure of obtaining partial hydrographs produced by segments of
rainfall continues and the ordinates of these hydrographs are added to result in the total
hydrograph.
Table 2

Application of the convolution method

Time

H1.u(t)

h1.u1 h1.u2 h1.u3 h1.u4 h1.u5 h1.u6 h1.u7 h1.u8

H2.u(t)
H3.u(t)

10

h2.u1 h2.u2 h2.u3 h2.u4 h2.u5 h2.u6 h2.u7 h2.u8


h3.u1 h3.u2 h3.u3 h3.u4 h3.u5 h3.u6 h3.u7 h3.u8

h1.u3
Total
h1.u2 +
h2.u2
hydrograph h1.u1 +
ordinates
h2.u1 +
h3.u1

h1.u4
+
h2.u3
+
h3.u2

h1.u5
+
h2.u4
+
h3.u3

h1.u6
+
h2.u5
+
h3.u4

h1.u7
+
h2.u6
+
h3.u5

h1.u8
h2.u8
+
h3.u8
h2.u7 +
h3.u7
+
h3.u6

Invariance, proportionality and superposition (additivity) are fundamental principles of the


linear unit hydrograph method. These hypotheses have to be carefully considered since
they are not always respected. For example, small basins have different unit hydrographs
according to rainfall intensity. The higher the rainfall intensity, the smaller the base time
and the higher the peak discharge of the unit hydrographs. Also, linearity is not always
guaranteed, especially in urban areas.

55
55/113

In general, the ordinates Qi of the discharge hydrograph (at the instant i) that is produced
by an effective rainfall having segments hi of the same reference duration , are given by
the following relation:

(4.5.1)
Looking at Table 4.5.1 we have:

as

(4.5.2)

as

*
*(The total number of ordinates of the unit hydrograph is N-M+1, hence 10-3+1=8)

In the case that effective rainfall has a continuous variation in time, the ordinate of the
hydrograph at the moment t expressed as a function of the unit hydrograph and a certain
effective rainfall, is given by the Duhamel integral of convolution, namely:

(4.5.3)

56
56/113

4.5.2

Unit Hydrograph Derivation

The discrete convolution equation allows for determining the ordinates of the unit
hydrograph of a certain reference duration on the basis of the recorded hyetograph of
effective rainfall and the resulted discharge hydrograph. This procedure is called
"deconvolution" (Chow et al., 1988;). Suppose that there are M pulses of effective rainfall hi
, i = 1, M (number of rainfall segments of equal duration ) and N pulses of direct runoff Qj ,
j = 1,N (number of the ordinates of the flood wave hydrograph). It is observed that the
number of ordinates of the unit hydrograph is Nu = N - M + 1. N equations may be written
for Qj , (j = 1, 2,.........., N), in terms of N - M + 1 unknown values of the unit hydrograph, as
presented for example in Table 2.
Furthermore, the unit hydrograph ordinates can be derived by either starting from the first
equation of the system 4.5.2 up to the eighth equation, inclusively (Method of substitution
forwards) or commencing with the tenth equation of the same system back to the third
equation, inclusively (Method of substitution backwards) with Nu = 8.
Method of substitution forward:

Method of substitution backward:

(4.5.4)

57
57/113

If another set of effective rainfall and discharge hydrograph is used, the solution obtained
by use of these two methods does not result in an identical unit hydrograph. This variation
of the unit hydrograph with rainfall-runoff events is partially explainable by the nonlinearity
in the rainfall-runoff relation, approximations induced by applying methods and models for
deriving effective rainfall, and assumption of even spatial and temporal distribution of
rainfall.
Another unique approximate solution, the method of successive approximation (Collins
Method) can be used. This method consists of the following steps:
Assume a unit hydrograph and apply it to the elements of the effective rainfall hyetograph
except the largest one;
Subtract the resulting hydrograph obtained in the previous step from the actual one and
reduce the residual to unit hydrograph terms;
Compute a weighted average of the assumed unit hydrograph and the residual unit
hydrograph, and use it as a revised approximation for the next trial;
Repeat the previous steps until the residual unit hydrograph does not significantly differ
from the assumed one.
Deconvolution may be used for determining the unit hydrograph from multipeaked
hydrographs, making use of matrix calculation methods. Equation 4.5.5 can be written in
form of a matrix product, as follows (Chow, 1988):

(4.5.5)

Or

58
58/113

(4.5.6)

Assume that there exists a solution for the unit hydrograph ordinates that results in a
computed hydrograph of direct runoff, as:

(4.5.7)

But, the above written equation can be also expressed as

(4.5.8)

with all equations satisfied.


The solution consists in minimising the difference

between actual and computed

runoff hydrographs. Linear programming or least square methods may be used (Chow,
1988).

4.5.3

The "S" Hydrograph

The "S" hydrograph is defined as the integral of a hydrograph of a continuous effective


rainfall having a duration greater than the concentration time tc of the basin. Since the
concentration time of the basin represents the necessary travel time that a drop of effective
rain that falls at the endmost point of the catchment would take to reach the outlet, the
entire area of the basin contributes to the formation of the maximum peak discharge of the
hydrograph. Consequently, after the moment when the duration of the rainfall exceeds the
concentration time the "S" curve becomes horizontal, assuming uniform temporal and
spatial characteristics of effective rainfall.

59
59/113

Fig. 16 The construction of the "S" curve


The "S" curve allows for deriving the unit hydrograph of certain duration T of effective
rainfall, starting from a unit hydrograph having the reference duration , through the
following steps (Figure 16):
shift two "S" hydrographs with a time interval T,
abstract the ordinates of one hydrograph from the other and
divide by the ratio T/
the unit hydrograph of effective rainfall of duration T is derived.

Fig. 17 Computation of the unit hydrograph of reference duration


For cases where the duration T is a multiple n integer of (T = n), the unit hydrograph can
be found more easily by applying the principle of convolution. Thus the unit hydrograph of
duration is shifted n times with the duration , the ordinates are added and the resulted
hydrograph is divided by n.

60
60/113

4.5.4

The Instantaneous Unit Hydrograph (IUH)

4.5.4.1 Definition and Properties


The instantaneous unit hydrograph is defined as a unit hydrograph produced by an
effective rainfall of 1 mm and having an infinitesimal reference duration (in other words the
duration tends towards zero). In terms of the systematic conception this means that the
instantaneous unit hydrograph represents the response of the catchment to an
instantaneous impulse. Of course, this is only a theoretical concept and cannot be realised
in actual catchments, but it is useful as the IUH can be analytically computed and, on the
basis of the "S" curve, a certain unit hydrograph of any reference duration may be derived
from it. Moreover, the IUH can be related to the catchment geomorphology.
It is reminded that the product of convolution of two mathematical functions is defined as:

(4.5.16
where:
g() represents the intensity as a function of time t
f(t)

is the transfer function

Let S(t) be the dimensionless ordinates of the time t of the "S" curve of a river basin.
According to Figure 4.5.5 the ordinates of the unit hydrograph of a duration are given by
the relation:
(4.5.17)
where u(t) has the dimensions [h-1, min-1], S is dimensionless and is a duration
The ordinates of the instantaneous hydrograph u0 corresponds to a duration

hence:
(4.5.18)

and
(4.5.19)

61
61/113

Furthermore, the intensity function should be now defined. Let Ie() represent the effective
rainfall intensity function that is multiplied by the area of the river catchment A and by the
duration in order to determine effective rainfall volume.
corresponds to the volume

According to Figure 17 the product

generated by the intensity Ie(). Integrating these elementary discharges over the duration
of rainfall, the product of convolution is given by:

(4.5.20)

Fig. 17

Convolution of the intensity function Ie(t) with the instantaneous unit


hydrograph u0(t) giving the discharge hydrograph Q(t)

The boundaries of integration are:

should not exceed the total duration of effective rainfall, that is


t - should not exceed the base time of the unit hydrograph Tb , that is

;
;

hence the following condition will be satisfied:


The product of convolution allows for maintenance of the linearity and invariance
hypotheses of the theory of the unit hydrograph.
In the following paragraphs instantaneous unit hydrographs resulting from different
conceptual models are presented.
62
62/113

4.5.4.2 Instantaneous Unit Hydrograph Resulting From Conceptual Models


The conceptual unit hydrographs result from the representation of the integration process
of the effective rainfall as a cascade of reservoirs that fill and deplete according to certain
laws. Let us consider one reservoir as presented in the Figure 18 that is characterised by
three functions that are the rate of filling I(t), the storage S(t), and the rate of depletion
(drainage) Q(t).

Fig. 18 A linear reservoir


The behaviour of the function of this reservoir is ruled according to the following laws:
Continuity law:

(4.21)
Storage law:
(4.22)

where K is the storage coefficient and it is a constant that confers to the reservoir the
property of linearity.
In effect, replacing 4.22 in 4.21 one obtains:

(4.23)

63
63/113

The solution of this equation is of the type:

(4.24)

where M(d) and N(d) are the differential operators relative to the I and Q, respectively. The
ratio M(d)/N(d) is defined as the transfer function of the system. The methods for
establishing above relations as proposed by several authors is discussed below.

4.5.4.3 The Instantaneous Unit Hydrograph of Nash


The instantaneous unit hydrograph proposed by Nash (1957)(After Chow, 1988) is a
conceptual model that uses the effect of "lamination" of an effective rainfall having the
depth of 1mm and duration that tends towards zero. This lamination is performed by a
cascade of n linear reservoirs having identical storage coefficient K (Figure 19).

Fig. 19 The Nash concept for deriving the instantaneous unit hydrograph
Equation (4.23) may be written as:
(4.25)
Considering the equation of convolution the expression of Q(t) is:

64
64/113

(4.26)
The integration of equation (4.25) is obtained by multiplying both terms from the right and
left side with et/K and one obtains:
(4.27)
Thus the output from the first reservoir becomes the input in the second one and so on.
Therefore, replacing t - by yields:
(4.28)
Again the convolution equation allows for determining the output from the second reservoir:
(4.29)
or:
(4.30)
Reiterating the procedure for n reservoirs the general expression of the ordinates of the
Nash instantaneous unit hydrograph is obtained as:
(4.31)
where (n) is the gamma function defined as:
(4.32)
.......
with the following iterative relation:
(4.33)

Relation 4.31 is the probabilistic distribution function with two parameters. It is well
known that the integral of the probabilistic distribution function is always equal to 1, which
is the volume of the unit hydrograph.

65
65/113

The increase time tp and the peak discharge Qp are derived by solving the equation dQ/dt
= 0, as follows:
(4.34)
and:
(4.35)

Finally, the ordinates of the instantaneous unit hydrograph might be expressed as a


function of the increase time and the peak discharge, as given below:
(4.36)

The instantaneous unit hydrograph derived by Nash is determined only as a function of two
parameters, namely n and K . The influence of these parameters upon the shape and the
magnitude of the hydrograph are presented in Figure 20. One observes that while the
parameter n influences the shape of the hydrograph, the parameter K determines the
position of the peak discharge. The higher the parameter K, the further the peak discharge
is shifted with respect to abscissa. When the parameter n augments the peak discharge Q
augments too, since, under the circumstances that tp is kept constant, the increase of the
parameter n corresponds to the diminution of the storage coefficient; and therefore the
wave becomes sharper.

Fig. 20 Sensitivity of the Nash model to the parameters n and K(tp)

66
66/113

Another example of the sensitivity of the Nash instantaneous unit hydrograph to the
parameter n while the storage K parameter is kept constant is presented in Figure 21.

Fig. 21 Sensitivity of the Nash model to the parameters n and K


where the moments are computed as presented in Figure 22

.:
Fig. 22 The method of moments for deriving the instantaneous hydrograph

67
67/113

Variation in values of n and K determined by application of the method of moments to


many flood events over a given catchment.
Applying the "S" curve method, the unit hydrographs specific to the considered reference
duration are determined for all flood events. These unit hydrographs are multiplied by
convolution with the corresponding effective rainfalls, resulting in computed hydrographs.

4.6

4.6.1

The Geomorphologic Unit Hydrograph

Principles and Methods of Determination

The transfer function of the catchment is directly influenced by the geomorphologic


characteristics as presented in Figure 23. This influence is mainly due to the topology of
the river network as it has been described by the classification of Strahler and by the ratios
of Horton and Schumm. These ratios or geomorphologic parameters are:
The proportionality coefficient Rb of the number of tributaries Nu of u order against the
number of tributaries Nu+1 of u + 1 order:
(4.62)

This coefficient is also called the bifurcation coefficient.


The proportionality coefficient Rl of the length of tributaries Lu of u order against the length
of tributaries Lu+1 of u + 1 order:
(4.63)

The proportionality coefficient Ra of the area of tributaries Au of u order against the area of
tributaries Au+1 of u + 1 order:

(4.64)

68
68/113

Fig. 23 Influence of coefficients of bifurcation on the flood hydrograph


Rodrigues-Iturbe and Valdes (1979) developed an analogy of the catchment, presented in
Figure 24. In this Figure circles represent river courses of i order, which is inscribed in the
centre.
The starting point of a drop of rainfall is, in this example, a river course of the order 1, 2 or
3. Assuming that this starting point is randomly occurring, there are the probabilities p01,
p02 and p03 that the drop of water commences its travel from 1, 2 or 3. These probabilities
represent the initial state of the catchment. Following the flow of the water drop within the
river network, denote p12 the probability that the drop travels from a watercourse of first
order to a watercourse of second order. In general pij is defined as the probability of
transition from the state i to the state j with pij = 0 when j <1 (as one passes always from
upstream with an inferior state to downstream with a superior one).

Fig. 24 Schematisation of a catchment of 3rd order


The following hypotheses are made:
Effective rainfall falls only on the catchment area except the watercourses;
Effective rainfall over an area of a given order is drained exclusively by a river course of
the same order.
69
69/113

Transition from one state to another is achieved provided that the order of the new state is
superior to the antecedent one.
All water drops cross the outlet section of the river basin;
The passage from one state to another (from a watercourse to another one) depends only
on the site where the water drop is found. In other words, the water drop has no memory of
its trajectories up to the present watercourse.
This analogy allows for development of a unit hydrograph function of the above-mentioned
geomorphologic parameters. Nevertheless, there is no general expression of the
geomorphologic unit hydrograph because this expression changes with the order of the
catchment.
However a definition of the geomorphologic unit hydrograph can be given, namely: The
geomorphologic unit hydrograph is the statistical density function of a particle of water
remaining over a river basin.
The following general equations allow for calculation of the two main parameters of the unit
hydrograph, namely the peak discharge Qp and the increase time tp (time-to-peak).

(4.65)

(4.66)

where:
L

length of the main watercourse

mean velocity of the main watercourse

Rl,
Rb,Ra

main geomorphologic characteristics of the above-mentioned river basin

It is worth mentioning that the geomorphologic unit hydrograph is not linear because its
main characteristics Qp and tp vary with the velocity of the main river course.

70
70/113

4.7

Regionalisation of the Unit Hydrograph (Synthetic Unit Hydrograph)

The unit hydrograph derived from data recorded on rainfall and the corresponding flood
hydrograph applied only for the basin for which the data was available. The synthetic unit
hydrograph is computed on the basis of data recorded at gauging stations and might be
used for assessing the flood wave in ungauged catchments having determined effective
rainfall over the basin.
There are three types of synthetic unit hydrographs (Chow et al., 1988)
Those relating the unit hydrographs characteristics (peak discharge, time to peak, base
time, etc.) to the morphologic parameters of the catchment. The procedure of such
geographical interpolation is called regionalisation.
Those based on a procedure that makes dimensionless the unit hydrograph (see the USSoil Conservation Service unit hydrograph). This could be used for assessing the unit
hydrograph in the similar regions with those for which the dimensionless unit hydrograph is
available.

4.7.1

The Synthetic SCS Unit Hydrograph

The Soil Conservation Service (SCS) has performed a synthesis of a great number of unit
hydrographs for river basins of various sizes located in different regions of the USA. In
order to compare these unit hydrographs their ordinates and abscissas were normalised.
Were divided the ordinates by the peak discharges up while the abscissas were divided by
the time to peak tp. Furthermore, the non-dimensional ordinates and abscissas were
averaged obtaining the non-dimensional synthetic hydrograph (SUH), the abscissas t/tp
and ordinates of which u(t)/up are given in Table 4.7.1.

71
71/113

Table 4.7.1

The US-Soil Conservation Service synthetic unit hydrograph

Time
t/tp
0,0

SUH
u(t)/up
0

Time
t/tp
1,1

SUH
u(t)/up
0,99

Time
t/tp
2,4

SUH
u(t)/up
0,147

0,1

0,03

1,2

0,93

2,6

0,107

0,2

0,1

1,3

0,86

2,8

0,077

0,3

0,19

1,4

0,78

3,0

0,55

0,4

0,31

1,5

0,68

3,2

0,04

0,5

0,47

1,6

0,56

3,4

0,029

0,6

0,66

1,7

0,46

3,6

0,021

0,7

0,82

1,8

0,39

3,8

0,015

0,8

0,93

1,9

0,33

4,0

0,011

0,9

0,99

2,0

0,28

4,5

0,005

1,0

1,0

2,2

0,207

5,0

The synthetic unit hydrograph and its corresponding "S" curve are presented in Figure 25.
Also in the same Figure the triangular approximation of the SCS-synthetic unit hydrograph
and its "S" curve is shown.

Fig. 25 The SCS-synthetic unit hydrograph and its triangular approximation


The following variables presented in the graph are defined:
Duration of effective rainfall D ;
The concentration time tc between the end of effective rainfall and the inflexion point of the
hydrograph;

72
72/113

The lag time t1 between the mid-point of the effective rainfall and the time of the peak
discharge (approx. t1 = 0.66 tc);
The increase time of the hydrograph tp;
The decrease time of the hydrograph tr;
The base time of the hydrograph tb.
This triangular SCS-synthetic hydrograph is defined by the parameters up (maximum
discharge) and tp. In the triangular hydrograph 37.5% of the volume of the flood wave is
found just before the peak, hence tp/tb = 0.375 or
(4.8.1)

The volume of the hydrograph is:


(4.8.2)
From equation 4.7.1 one obtains the expression of the peak discharge as a function of
increase time:
(4.8.3)

(4.8.4)
with

Considering that:
(4.8.8)

and replacing the relation 4.7.5 in 4.7.4. the parameter K = 0,75


Linking the SCS-synthetic unit hydrograph with the reference duration of effective rainfall,
D, one obtains:
(4.8.9)

As such, SCS consider that the following relation is valid for the United States:

73
73/113

(4.8.10)

4.7.2

Snyder's Synthetic Unit Hydrograph

Snyder (1938) made a study of the unit hydrograph expressed as a function of several
characteristics of the basin. First of all, Snyder standardised the unit hydrograph as one
those rainfall duration tr is related to the catchment lag time tlag by

(4.8.11

For a standard unit hydrograph the following relations have been derived:

(4.8.12)

where:
tlag catchment lag time in hours
L

length of the main river course in km from the outlet to the upstream divide line

Lc

distance in km from the outlet to the point of the river nearest the centroid of the
catchment

C1 coefficient derived from the gauged catchments in the same region (1 for the
English system)

The specific peak discharge [m3/km2] of the standard hydrograph is given by:
(4.8.1
3
where:
C2 = 2,75 (640 for the English system)
Cp coefficient derived from the gauged catchments in the same region

74
74/113

For a given effective rainfall tR that is different of durations tr of the standard hydrograph
the "derived unit hydrograph obtained from the data on effective rainfall and corresponding
hydrograph" would be defined by several parameters as shown in Figure 26.
Peak discharge of the unit hydrograph ; upR
Lag time defined as the time difference between the centroids of effective rainfall and the
unit hydrograph peak ; tlagR
Base time as defined above ; tb
Widths W50 and W75 at 50% and 75% of the unit hydrograph peak discharge
If then tlaf R = 5.5 tR and tR = tr and upR = up and Ct and Cp are computed applying
equations 4.7.12 and 4.7.13
If tlaf R 5.5 tR then the lag time of the standard basin is:

(4.8.14)
and values of tlag and tr are obtained from equations 4.8.11 and 4.8.12, which are solved
simultaneously. Values of the coefficients Ct and Cp are computed from 4.8.12 and 4.8.13
with upR = up and tlaf R = tlag.

Fig. 26 Snyder's unit hydrograph; (a) Standard unit hydrograph (tlag= 5.5 tr )
(b) Required unit hydrograph (tlag 5.5 tr). (Chow et al., 1988)
75
75/113

Once the coefficients Ct and Cp are derived for both cases tlaf R = 5.5 tR or tlaf R 5.5 tR
the unit hydrograph indicators of an ungauged catchment situated in a region that is
similar, from the point of view of physiographical characteristics of the landscape, with the
basins from which the coefficients Ct and Cp were derived, are given by:
The peak discharge is
(4.8.15)

The base time is derived assuming the shape of the unit hydrograph is a triangle of an
unit area (1mm):
(4.8.16
)
where:
C3 = 5,56 (1290 for the English system)

The widths (hours) of unit hydrograph corresponding to a discharge of the percentage


75% and 50% is computed as:
(4.8.17)

where:
Cper = 1,22 (440 for the English system) for per = 75%
Cper = 2,14 (770 for the English system) for per = 50%

Mention is made that practically one-third of the width is distributed before the peak time
and two-thirds after the peak.

76
76/113

4.8

Workout Examples and Exercises

Example 4.8.1

Given below are the ordinates of a 6-h unit hydrograph for a


catchment. Calculate the ordinates of the DRH due to a rainfall
excess of 3.5 cm occurring in 6 hr.

Time
0
(h)
0
UH
ordinate
(m3/s)

12

15

18

24

30

36

42

48

54

60

69

25

50

85

125

160

185

160

110

60

36

25

16

The desired ordinates of the DRH are obtained by multiplying the ordinates of the unit
hydrograph by a factor of 3.5 as in Table 4.8.1. Note that the time base of DRH is not
changed and remains the same as that of the unit hydrograph. The intervals of coordinates
of the unit hydrograph (shown in column 1) are not in any way related to the duration of the
rainfall excess and can be any convenient value.
Table 4.8.1
Time
(h)
1
0
3
6
9
12
15
18
24
30
36
42
48
54
60
69

Calculation of DRH Due to 3.5 cm ER Example 4.8.1


Ordinate of 6-h
unit hydrograph
(m3/s)
2
0
25
50
85
125
160
185
160
110
60
36
25
16
8
0

Ordinate of 3.5 cm
DRA
(m3/s)
3
0
87.5
175.0
297.5
437.5
560.0
647.5
560.0
385.0
210.0
126.0
87.5
56.0
28.0
0

77
77/113

Example 4.8.2

The effective rainfall hyetograph is calculated as in the following


table.

Interval
Rainfall depth (cm)
Loss @ 0.25 cm/h
For 6 h
Effective rainfall (cm)

Question 4.8.1

1st 6 hours
3.5
1.5

2nd 6 hours
7.5
1.5

3rd 6 hours
5.5
1.5

2.0

6.0

4.0

Following are the ordinates of a storm hydrograph of a river draining


a catchment area of 423 km2 due to a 6-h isolated storm. Derive the
ordinates of a 6-h unit hydrograph for the catchment.

Time from
start
of
storm (h)
Discharge
(m3/s)
Time from
start
of
storm (h)
Discharge
(m3/s)

-6

12

18

10

10

30

87.5

54

60

66
26.0

39.0 31.5

Example 4.8.3

24

30

36

42

48

115.5 102.5 85.0

71.0

59.0

47.5

72

78

84

90

96

102

21.5

17.5

15.0

12.5

12.0

12.0

Given the ordinates of a 4-h unit hydrograph (columns 1 and 2 in)


derive the ordinates of a 12-h unit hydrograph.

The calculations are performed in a tabular form in Table 4.8.2. In this


Column 3 = ordinates of 4-h UH lagged by 4 h
Column 4 = ordinates of 4-h UH lagged by 8 h
Column 5 = ordinates of DRH representing 3 cm ER in 12 h
Column 6 = ordinates of 12-h UH = (column 5)/3

78
78/113

Table 4.8.2

Calculation of a 12-h Unit Hydrograph from a 4-h Unit Hydrograph

Example 4.8.3

Time
(h)
A
1

0
4
8
12
16
20
24
28
32
36
40
44
48
52

0
20
80
130
150
130
90
52
27
15
5
0

Example 4.8.4
Table 4.8.3

Ordinates of 4-h UH
(m3/s)
B
C
Lagged by
Lagged by
4h
8h
3
4
0
20
0
80
20
130
80
150
130
130
150
90
130
52
90
27
52
15
27
5
15
0
5
0

DRH
(Col.
(2+3+4)

Ordinate of
12-h UH
(m3/s)
(Col. 5)/3

0
20
100
230
360
410
370
272
169
94
47
20
5
0

0
6.7
33.3
76.7
120.0
136.7
123.3
90.7
56.3
31.3
15.7
6.7
1.7
0

Solve Example 4.8.3 by the S-curve method.(Table:4.8.3)

Determination of a 12-h Unit Hydrograph by S-Curve Method-

Time
(h)

Ordinate
Of 4-h
UH
(m3/s)

S-curve
Addition
(m3/s)

1
0
4
8
12
16
20
24
28
32
36
40
44
48
54

2
0
20
80
130
150
130
90
52
27
15
5
0

3
0
20
100
230
380
510
600
652
679
694
699
699

S-curve
Ordinate
(m3/s)
(Col. 2+
Col. 3)
4
0
20
100
230
380
510
600
652
679
694
699
699
699
699

S-curve
lagged by
12 h
(m3/s)

(Col. 4Col. 5)

5
0
20
100
230
380
510
600
652
679
694
699

6
0
20
100
230
360
410
370
272
169
94
47
20
5
0

Col. 6
(12/4)
= 12-h UH
ordinates
(m3/s)
7
0
6.7
33.3
76.7
120.0
136.7
123.3
90.7
56.3
31.3
15.7
6.7
1.7
0

79
79/113

Exercise 4.8.2

A 6-h unit hydrograph for a catchment of area 1000 km2 can be


approximated as a triangle with base of 69 h. Calculate the peak
discharge of this unit hydrograph.

Example 4.8.3

Two catchments A and B are considered meteorologically similar.


Their catchment characteristics are given below.

Catchment A
L = 30 km
Lca = 15 km
A = 250 km2

Catchment B
L = 45 km
Lca = 25 km
A = 400 km2

For catchment A, a 2-h unit hydrograph was developed and was found to have a peak
discharge of 50 m3/s. The time to peak from the beginning of the rainfall excess in this unit
hydrograph was 9.0 h. Using Snyders method, develop a unit hydrograph for catchment B.
For Catchment A:
tR = 2.0 h
Time to peak from beginning of ER =

tR
t p 9.0h
2

tp = 8.0 h
From Eq.

t 'p

21
1R
tp
22
4

21
t p 0.5 8.0
22

tt

7.5 22
7.857h
21

From Eq.
tp = Ct (L Lca)0.3
7.857 = Ct (3015)0.3
Ct = 1.257
From Eq.
Qp = 2.78 Cp A/tp
50 = 2.78Cp250/8.0
Cp = 0.576

80
80/113

For Catchment B: Using the values of Ct = 1.257 and Cp = 0.576 in catchment B, the
parameters of the synthetic-unit hydrograph for catchment B are determined.
From Eq.
Tp = 1.257 (4525)0.3 = 10.34 h
By Eq.

tr

10.34
1.88h
5.5

Using tR = 2.0 h, i.e. for a 2-h unit hydrograph, by Eq.

t 'p 10.34

21 2.0

10.37h
22 4

By Eq.

Qp

2.78 0.576 400


10.37

= 61.77 m3/s, say 62 m3/s


From Eq.

W50

5.87
44h
(62 / 400)1.08

W75

44
25h
1.75

Time base: tb = 72+(310.37) = 103 h


tb = 5 (10.37+10) 58h
Considering the values of W50 and W75 and noting that the area of catchment B is rather
small,
tb 58 h is more appropriate in this case.
Exercise 4.8.4

A 6-h unit hydrograph for a catchment of area 1000 km2 can be


approximated as a triangle with base of 69 h. Calculate the peak
discharge of this unit hydrograph.

81
81/113

HYDROLOGIC ROUTING

5.1

Introduction

Hydrologic routing is the process of converting a hydrograph that passes through some
part of a flow system to allow for the changes that occur during its passage. There are
three main types of hydrologic routing:

catchment routing, which converts a rainfall excess hyetograph into a hydrograph at


the catchment outlet, allowing for the distribution of rainfalls over the catchment
surface, and various lags or delays along flow paths;

channel routing, which allows for the changes in hydrographs as they flow along
river or channel reaches, caused by variations in the channel geometry which result
in storage effects,

reservoir routing, which allows for storage effects in a concentrated, level pool
reservoir.

Broadly defined, flow routing is an analytical procedure intended to trace the flow of water
through a hydrological system, pond, conveyance, or porous media, given some runoff
event hydrograph as input. The procedure determines the flow hydrograph at a point
downstream, from known or assumed flow hydrographs at one or more points upstream. If
the flow is a runoff event such as a flood, then the procedure is specifically known as flood
routing. Routing by lumped system methods is called hydrological routing. These methods
calculate the flow as a function of time alone. Routing by distributed system methods is
called hydraulic routing, and the flow is calculated as a function of both space and time
throughout the system.
For hydrologic routing the input I(t), output Q(t), and storage S(t) are related by the
continuity equation:

( 5.1 )

82
82/113

If an inflow I(t) is known Equation 5.1 cannot be solved directly to obtain the outflow Q(t),
because both Q and S are unknown. A second relationship, the storage function, is
required to relate I, S, and Q. Coupling the continuity equation with the storage function
provides a solvable combination of two equations and two unknowns.
The specific form of the storage functions to be employed in hydrologic routing depends on
the nature of the system being analysed. In reservoir routing by the level-pool method,
storage is a non-linear function of Q only, i.e. S = f(Q), and the function, f(Q), is determined
by relating reservoir storage and outflow to reservoir water level. In channel routing by the
Muskingum method, storage is linearly related to I and Q. Similarly, in porous media,
storage is a function of outflow which depends on storage in the media and the underlying
soils.

5.2

Reservoir Routing

Level-pool routing is a procedure for calculating the outflow hydrograph from a pond
reservoir, assuming a horizontal water surface, given its inflow hydrograph and storagedischarge characteristics. When a reservoir has a horizontal water surface, its storage is a
function of its water-surface elevation, or depth in the pool. Likewise, the discharge is a
function of the water surface elevation, or head on the outlet works. Combining these two
functions yields the invariable single-valued function.
Integration of the continuity equation (Equation 5.1) over the discrete time intervals
provides an expression for the change in storage over the j th time interval j t , Sj+1 - Sj,
which can be rewritten as:

( 5.2 )
The inflow values at the beginning and end of the j th time interval are Ij and Ij+1,
respectively, and corresponding outflow values are Qj and Qj+1. The values of Ij and Ij+1,
are known because they are pre-specified (i.e. the inflow hydrograph ordinates). The
values Qj and Sj are known at the j th time interval. Hence Equation 5.2 contains two
unknowns, Qj+1 and Sj+1, which are isolated by multiplying by 2/t and rearranging the
result to produce:

83
83/113

( 5.3 )
In order to calculate the outflow Qj+1 from Equation 5.3, a storage-discharge function
relating 2S/t + Q and Q is needed. The method of developing this function using stagestorage and stage-discharge relationship is shown in Figure 27.

Fig. 27

Development of the Storage-Discharge Function for Hydrologic Pond


Routing

84
84/113

The relationship between water-surface elevation and reservoir storage can be obtained
using topographic maps or from field surveys. The stage-discharge relationship is derived
from hydraulic equations relating head and discharge for various types of spillway and
outlet works. The value of t is the same as the time interval of the inflow hydrograph.
For a given water-surface elevation, the values of storage S and discharge Q are
determined. Then, the value of 2S/t+Q is calculated and plotted against Q. In routing the
flow through the j th time interval, all terms in the right-hand-side of Equation 5.4 are
known, and so the value of 2Sj+1/t+Q can be computed. The corresponding value of
Qj+1 can be determined from the storage-discharge function 2S/t+Q versus Q. To set up
the data for the next time interval, the value 2Sj+1/t-Qj+1 is calculated by:
( 5.4 )
The computation is repeated iteratively for subsequent routing periods. Input requirements
for this routing method are:

the storage-discharge relationship

the storage-indication relationship

the inflow hydrograph

initial values of the outflow rate (Q1) and storage (S1)

the routing interval (t)

5.3

Channel Routing

A widely used hydrologic method for routing flows in conveyance systems is the
Muskingum method. It models the storage volume of flow in a channel reach by a
combination of wedge and prism storage (Figure 28). When the flood wave is advancing,
inflow exceeds outflow and a positive wedge of storage is produced. When the flood is
receding, outflow exceeds inflow and a negative wedge results. In addition, a prism of
storage is formed by a volume of (approximately) constant cross-section along the length
of the channel reach.

85
85/113

Assuming that the cross-sectional area of the flow is directly proportional to the discharge
at the section, the volume of prism storage is KQ, where K is a coefficient of
proportionality. The volume of wedge storage is assumed to be equal to KX(I-Q), where X
is a weighting factor having the range 0 X 0.5. The total storage is defined as the sum
of the two storage components S=KQ+KX(I-Q) which can be rearranged to give the linear
storage function for the Muskingum method:

(5.5 )

Fig. 28 Prism and Wedge Storage in a Channel Reach


The value of X depends on the shape of the modeled wedge storage, X = 0 corresponds to
reservoir (level-pool) storage and Equation 5.5 reduces to S = KQ.X = 0.5 for a full wedge.
In most natural stream channels, X is between 0 and 0.3, with a mean value near 0.2.
Great accuracy in determining X is usually not necessary because the Muskingum method
is not sensitive to this parameter.
The parameter K is the time of travel of the flood wave through the reach. Although
techniques exist that allow for the values of K and X to vary according to flow rate and
channel characteristics (e.g., the Muskingum-Cunge method), for hydrologic routing the
values of K and X are assumed to be specified and constant throughout the range of flow.
The change in storage over the time interval t (from j to j +1) is:

(5.6)
The change in storage can also be expressed by Equation 5.5 and combining Equations
5.2, Eq. 5.6 yields the Muskingum routing equation:

(5.7)

86
86/113

where,

( 5.8 a )
( 5.8 b )
( 5.8 c )
Note that C1 +C2 +C3 = 1.

The values of K and X are determined using observed


inflow and outflow hydrographs in the channel reach.
By assuming various values of X and known values of
inflow, successive values of K can be computed using:
(5.9)
The value of X that produces a loop closest to a single
line is taken to be the correct value for the reach. The
parameter K, from Equation 28 is the slope of the line
for the value of X. Figure 5.3 shows the procedure of
destermining the value of X and K.

Fig. 28
Procedure for Determining X and
K Values

87
87/113

5.4

One-Dimensional Hydraulic Routing

Procedures for distributed-flow hydraulic routing are popular because they compute flow
rate and water level as functions of both space and time. The methodologies are based
upon the Saint-Venant equations of one-dimensional flow. In contrast, the lumped
hydrologic-routing procedures discussed in the previous sections compute flow rate as a
function of time alone.
Unsteady flow is described by the conservation form of the Saint-Venant equations. This
form provides the flexibility to simulate a wide range of flows from gradual flood waves in
rivers to pipe flows, and can simulate lateral inflows or outflows such as weirs and
pumping. The equations are given below (Chow et al, 1988).
Continuity:
( 5.10)
Momentum:

( 5.11)
where,
x

longitudinal distance along the conveyance

time

cross-sectional area of flow

A0

cross-sectional area of dead storage (off-channel)

lateral inflow per unit length along the conveyance

water-surface elevation

vx

velocity of lateral flow in the direction of flow

Sf

friction slope

Se

eddy loss slope

width of the conveyance at the water surface

Wf

wind shear force

momentum correction factor

acceleration due to gravity

The Saint-Venant equations operate under the following assumptions:

88
88/113

(i) The flow is one-dimensional with depth and velocity varying only in the longitudinal
direction of the conveyance. This implies that the velocity is constant and the water surface
is horizontal across any section perpendicular to the longitudinal axis.
ii) There is gradually varied flow along the channel so that hydrostatic pressure prevails
and vertical accelerations can be neglected.
iii) The longitudinal axis of the channel is approximated as a straight line.
iv) The bottom slope of the channel is small and the bed is fixed, resulting in negligible
effects of scour and deposition.
v) Resistance coefficients for steady uniform turbulent flow are applicable, allowing for a
use of Manning's equation to described resistance effects.
vi) The fluid is incompressible and of constant density throughout the flow.
The momentum equation consists of terms for the physical processes that govern flow
momentum. When the water level or flow rate is changed at a certain point in a channel
with a subcritical flow, the effects of these changes propagate back upstream. These
backwater effects can be incorporated into distributed routing methods through the local
acceleration, convective acceleration, and pressure terms.
Hydrologic (lumped) routing methods may not perform well in simulating the flow conditions
when backwater effects are significant and the drain or channel slope is mild, because
these methods have no hydraulic mechanisms to describe upstream propagation of
changes in the flow momentum.
Example 5.1: Route the following flood hydrograph through the reservoir:
Time
(h)
Inflow
(m3/s)

12

18

24

30

36

42

48

54

60

66

10

30

85

140

125

96

75

60

46

35

25

20

The initial conditions are: when t = 0, the reservoir elevation is 101.60 m,

89
89/113

A time increment t = 6 h = 0.0216 hrs is chosen. Using the known storage-elevationdischarge data, the following table is prepared.
Elevation
100.00
(m)
Outflow Q 0
(m3/s)
2.S
310.2

100.50

101.00

101.50

102.00

102.50

102.75

103.00

10

26

46

72

100

116

130

331.5

385.3

451.8

524.0

597.2

627.8

672.2

(m3/s)

2S

Q vs elevation is prepared from this data.


t

A graph depicting Q vs elevation and

At t = 0, Elevation = 101.60, from Fig, Q = 12 m3/s and

2S
Q 340m 3 / s

2S

Q 340 24 316m s / s

t
1
For the first time interval of 6 h,
I1 = 10, I2 = 30, Q1 = 12, and

2S

Q (10 30) 316 356m 3 / s

t
2
2S

Q is 100.74,
t
2

From Fig the reservoir elevation for this


For the next time increment

2S

Q 356 2 12 322m 3 / s

t
1
The procedure is repeated in a tabular form (Table 5.1) till the entire flood is routed.
Using the data in columns 1, 8 and 7, the outflow hydrograph and a graph showing the
variation of reservoir elevation with time are plotted.
In this method also, the accuracy depends upon the value of t chosen; smaller values of
t give greater accuracy.

90
90/113

Table 5.1
Time
(h)

Reservoir Routing Example 5.1


I
(m3/s)

(I1+I2)

1
0

2
10

3
40

6
12
18
24
30
36
42
48
54
60
66

60
85
140
125
96
75
60
46
35
25
20

115
225
265
221
171
135
106
81
60
45

2S

(m3/s)
4
316
322
357
392
403
400
391
380
372
361
344
335

2S

(m3/s)
5
(340)
336
437
582
657
624
571
526
486
453
421
389

Elevation
(m)

Discharge
Q (m3/s)

6
100.60

7
12

100.74
101.38
102.50
102.92
102.70
102.32
102.02
101.74
101.51
131.28
101.02

17
40
95
125
112
90
73
57
46
37
27

Example 5.2: Route the following hydrograph through a river reach for which K = 12.0 h
and x = 0.20. At the start of the inflow flood, the outflow discharge is 10 m3/s.
Time (h)
Inflow
(m3/s)

0
10

6
20

12
50

18
60

C0

12 0.20 0.5 6
0.6

0.048
12 12 0.2 0.5 6 12.6

C1

12 0.2 0.5 6
0.429
12.6

C2

12 12 0.2 0.5 6
0.523
12.6

24
55

30
45

36
35

42
27

48
20

54
15

Table 5.2 Muskingum Method of Routing-Example 5.2


t = 6 h
Time (h)
0
6
12
18
24
30
36
42
48
54

I (m3/s)
10
20
50
60
55
45
35
27
20
15

0.048 I2

0.429 I1

0.523 Q1

0.96
2.40
2.88
2.64
2.16
1.68
1.30
0.96
0.72

4.29
8.58
21.45
25.74
23.60
19.30
15.02
11.58
8.58

5.23
5.48
8.61
17.23
23.85
25.95
24.55
21.38
17.74

Q (m3/s)
10.00
10.48
16.46
32.94
45.61
49.61
46.93
40.87
33.92
27.04

91
91/113

The calculations are shown in Table 5.2. For the first time interval, 0 to 6 h,
I1 = 10.0
I2 = 20.0
Q1 = 10.0
From Eq. (5.8),

C1 I1 = 4.29
C0 I2 = 0.96
C2 Q1 = 5.23
Q2 = 10.48 m3/s

For the next time step, 6 to 12 h, Q1 = m3/s. The procedure is repeated for the duration of
the inflow hydrograph. By plotting the inflow and outflow hydrographs the attenuation and
peak lag are found to be 10 m3/s and 12 h respectively.

92
92/113

FREGUENCY ANALYSIS OF HYDROLOGIC EVENTS

Flood frequency analyses are used to predict design floods for sites along a river. The
technique involves using observed annual peak flow discharge data to calculate statistical
information such as mean values, standard deviations, skewness, and recurrence
intervals. These statistical data are then used to construct frequency distributions, which
are graphs and tables that tell the likelihood of various discharges as a function of
recurrence interval or exceedence probability.

6.1

Prediction Methods of Extreme Events and Confidence Limit

Flood frequency distributions can take on many forms according to the equations used to
carry out the statistical analyses. Four of the common forms are:

Normal Distribution

Log-Normal Distribution

Gumbel Distribution

Log-Pearson Type III Distribution

Each distribution can be used to predict design floods; however, there are advantages and
disadvantages of each technique. As widely used, in this note later two methods are
described.

6.1.1

Gumbels Equation for Practical Use

Equation giving the value of the variate X with a recurrence interval T is used as

xT x K n 1

(6.1)

where n 1 = standard deviation of the sample

( x x) 2
N 1

(6.2)

K = frequency factor expressed as


93
93/113

yT yn
Sn

(6.3)

in which yT = reduced variate, a function of T and is given by

yT ln . ln
T 1

(6.4)

YT 0.834 2.303 log log


T 1

or

(6.5)

yn = reduced mean, a function of sample size N and is given; for N 00, yn 0.577
Sn = reduced standard deviation, a function of sample size
N and is given in Text Books; for N 00, Sn 1.2825

These equations are used under the following procedure to estimate the flood magnitude
corresponding to a given return based on an annual flood series by Gumbel distribution.
1.

Assemble the discharge data and note the sample size N. Here the annual
flood value is the variate X. Find x and n 1 for the given data.

2.

Using Tables (Chow,1988) )determine yn and Sn appropriate to given N.

3.

Find yT for a given T by Eq. (6.5)

4.

Find K by Eq. (6.3).

5.

Determine the required xT by Eq. (6.1).

6.1.2

Log-Pearson Type III Distribution

According to the U.S. Water Advisory Committee on Water Data (1982), the Log-Pearson
Type III Distribution is the recommended technique for flood frequency analysis. Therefore,
this analysis is examined in detail here with a step-by-step tutorial.
Log-Pearson Type III Distribution
What is it?

94
94/113

The Log-Pearson Type III distribution is a statistical technique for fitting frequency
distribution data to predict the design flood for a river at some site. Once the statistical
information is calculated for the river site, a frequency distribution can be constructed. The
probabilities of floods of various sizes can be extracted from the curve. The advantage of
this particular technique is that extrapolation can be made of the values for events with
return periods well beyond the observed flood events. This technique is the standard
technique used by Federal Agencies in the United States.
How is it calculated?
The Log-Pearson Type III distribution is calculated using the general equation:
(6.6)
where x is the flood discharge value of some specified probability,
the log x discharge values, K is a frequency factor, and

is the average of

is the standard deviation of the

log x values. The frequency factor K is a function of the skewness coefficient and return
period and can be found using the frequency factor table. The flood magnitudes for the
various return periods are found by solving the general equation. The mean, variance, and
standard deviation of the data can be calculated using the two formulas below.

(6.7)

and
or
Next, the skewness coefficient Cs can be calculated as follows:

where n is the number of entries, x the flood of some specified probability and

is the

standard deviation. Excel functions can also be used to calculate the variance (=VAR( ) ),
standard deviation (=STDEV( ) ), and skewness coefficient (=SKEW( ) ).

95
95/113

The skewness estimate (Cs) computed using the equation above is called the station
estimate, meaning that the estimate incorporates data values only from the gaging station
of interest.
Error and bias in the skewness estimate increase as the number of observations (n)
decreases. The Interagency Advisory Committee on Water Data (IACWD) uses a
generalized estimate of the coefficient of skewness, Cw (for instantaneous peak flow data
only), based on the equation:
Cw = WCs + (1-W)Cm
where W is a weighting factor, Cs is the coefficient of skewness computed using the
sample data, and Cm is a regional skewness, which is determined from a map.
The weighting factor W is calculated to minimize the variance of Cw, where

Determination of W requires knowledge of variance of Cm [V(Cm)] and variance of


Cs[V(Cs)]. V(Cm) has been estimated from the map of skew coefficients for the United
States as 0.302 (IACWD, 1982). This simplifies the denominator of the above equation by
substitution of 0.302 for V(Cm).
The variance of the station skew Cs for log Pearson type 3 random variables can be
obtained from the results of Monte Carlo experiments by Wallis et al. (1974). They showed
that

where
A = -0.33 + 0.08 | Cs| if | Cs |

0.90 or

A = -0.52 + 0.30 | C s | if | C s | > 0.90,


B = 0.94 - 0.26 | C s | if | C s |

1.50 or

B = 0.55 if | C s | > 1.50


in which | C s | is the absolute value of the station skew (used as an estimate of population
skew) and n is the record length in years.
The coefficient K is then found using tabulated values according to Cw and the return
period for each discharge.

96
96/113

For a more detailed description of this method, please refer to the following text:
Bedient, Philip B. and Wayne C. Huber. Hydrology and Floodplain Analysis. Prentice-Hall,
Inc., Upper Saddle River, 2002.

What does this particular information tell you about your river?
The Log-Pearson Type III distribution tells you the likely values of discharges to expect in
the river at various recurrence intervals based on the available historical record. This is
helpful when designing structures in or near the river that may be affected by floods. It is
also helpful when designing structures to protect against the largest expected event. For
this reason, it is customary to perform the flood frequency analysis using the instantaneous
peak discharge data. However, the Log-Pearson Type III distribution can be constructed
using the maximum values for mean daily discharge data. A tutorial and example is
supplied for both instantaneous and mean daily data.

6.2

Hydrologic data series

Two types of data series are used in frequency analysis:

Annual maximum series: Yearly maximum values. Number of data should be grater
than 25.

Partial duration series: Selected a magnitude of data larger than an arbitrary base
value.

The recurrence interval of an event obtained by annual series (TA) and by the partial
duration series (Tp) are related by

Tp

1
ln TA ln(TA 1)

From this it can be seen that the difference between TA and Tp is significant for TA<10
years and that for TA>20, the difference is negligibly small.

97
97/113

6.3

Return Period and Exceedence Probability:

A key planning question for water control measures is, "How much protection will a given
soil and water conservation practice/structure give me?" Our interest is not in the
impossible task of predicting the date of a storm event which exceeds the design capacity
of a practice or structure. Rather, we need to know the likelihood of occurrence
(probability) of an event with a specified intensity and duration. This is called its return
period or frequency of occurrence.
An event has a return period (or recurrence interval) of T years if its magnitude is equaled
or exceeded once, on the average, every T years. The reciprocal of the return period is the
exceedance probability of the event, that is, the probability that the event is equaled or
exceeded in any one year. For example, the 50-year flood has a probability of 0.02 or 2%,
of being equaled or exceeded in any single year. It is important to note that the return
period implies nothing about the actual time sequence of an event. The 50-year flood does
not occur once every 50 years; it is expected, for example, that on the average , about
twenty 50-year floods can be expected to be experienced during a 1,000 year period.
The likelihood or probability of an event with a specified intensity and duration is called the
return period or frequency. The intensity of a storm can be predicted for any return period
and storm duration, from charts based on historic data for the location. The term 1 in 10
year storm describes a rainfall event which is rare and is only likely to occur once every 10
years, so it has a 10% likelihood any given year. The rain fall will be greater and the
flooding will be worse than the worst storm one would expect in any single year. The term
1 in 100 year storm describes a rainfall event which is extremely rare and which will occur
with a likelihood of only once in a century, so has a 1% likelihood in any given year. The
rain fall will be extreme and flooding to be worse than a 1 in 10 year event. As with all
probability events, it is possible to have multiple "1 in 100 Year Storms" in a single year.

98
98/113

INTRODUCTION TO GROUND WATER HYDROLOGY

7.1

Forms of Subsurface Water

Water in the soil mantle is called subsurface water and is considered in two

Saturated zone, and

aeration zone.

Saturated Zone
This zone, also known as groundwater zone is the space in which all the pores of the soil
are filled with water. The water table forms its upper limit and marks a free surface, i.e. a
surface having atmospheric pressure.
Zone of Aeration
In this zone the soil pores are only partially saturated with water. The space between the
land surface and the water table marks the extent of this zone. Further, the zone of
aeration has three subzones:
Soil Water Zone
This lies close to the ground surface in the major root band of the vegetation from which
the water is lost to the atmosphere by evapotranspiration.
Capillary Fringe
In this the water is held by capillary action. This zone extends from the water table upwards
to the limit of the capillary rise.
Intermediate Zone
This lies between the soil water zone and the capillary fringe.
The thickness of the zone of aeration and its constituent subzones depend upon the soil
texture and moisture content and vary from region to region. The soil moisture in the zone
of aeration is of importance in agricultural practice and irrigation engineering. The present
chapter is concerned only with the saturated zone.

99
99/113

A water table is the free water surface in an unconfined aquifer. The static level of a well
penetrating and unconfined indicates the level of the water table at that point. The water
table is constantly n motion adjusting its surface to achieve a balance between the
recharge and outflow from the subsurface storage. Fluctuations in the water level in a dug
well during various seasons of the year, lowering of the groundwater table in a region due
to heavy pumping of the wells and the rise in the water table of an irrigated area with poor
drainage, are some common examples of the fluctuation of the water table. In a general
sense, the water table follows the topographic features of the surface. If the water table
intersects the land surface the groundwater comes out to the surface in the form of springs
or seepage.

7.2

Aquifers and Confining Bed

Aquifer - "A water-bearing layer of rock, or of unconsolidated sediments, that will yield
water in a usable quantity to a well or spring."
Confining Bed - "A layer of rock, or of unconsolidated sediments, that retards the
movement of water in and out of an aquifer and possesses a very low hydraulic
conductivity."
From these definitions, all rocks can be thought of as either aquifers or confining beds.
Moreover, aquifers are often considered to be unconfined or confined.
An aquifer is considered unconfined if water only partially fills the aquifer materials and
water freely rises and declines along the unsaturated/saturated zone boundary. These
unconfined aquifers are often referred to as water-table aquifers and wells that are opened
to these unconfined aquifers indicates the position of the water-table.
A confined aquifer is generally defined when water completely fills the aquifer materials
and is overlain by a confining bed. A common term for a confined aquifer is an artesian
aquifer. The water level from a well that permits water solely from a confined aquifer to
enter the well will stand at some point above the top of the confined aquifer but not
necessarily above the land surface. The water level in a well open to a specific confined
aquifer stands at the level of the potentiometric surface. If the potentiometric surface is
above land, the well is often considered as a free-flowing artesian well.
100
100/113

7.3

Properties of a Aquifers

Porosity: Our understanding of porosity provides us the ability to quantify the volume of
water contained in our aquifers. Porosity is best understood as the ratio of the voids to the
total volume of an unconsolidated or consolidated material.

Note: n = porosity as a decimal fraction


Vt = the total volume of a material
Vs = the volume of the solids in the material
Vv = the volume of the voids
Often, porosity is expressed as a percentage by multiplying the ratio by 100. Porosity also
depends on the range of grain size (sorting) and shape of the subject material, but not on
the size. Fine-grained materials tend to be better sorted than coarse-grained materials,
thereby exhibiting greater porosities.

By looking at the above table one might conclude that fine-grained materials have the
greatest ability to supply water to a well. This is not the case because physical properties
such as adhesion (the molecular attraction of two different substances) and cohesion (the
molecular attraction of similar substances) limit an aquifer's ability to release water into a
well. Water stored in an aquifer that is retained as a film on the aquifer material or in very
small openings is called specific retention. Conversely, water stored in an aquifer that will
drain under the influence of gravity is called specific yield. The term effective porosity is
often used to describe specific yield. With concepts of specific yield and retention in mind,
let's look at our porosity chart once again.

101
101/113

As we can conclude, finer-grained materials such as clay have a high porosity but also
have a high specific retention and low specific yield when compared to their coarse-grained
counterparts, thereby limiting its ability to provide water to a well.
Hydraulic gradient: As previously discussed, water entering an unconfined or confined
well will stand at a particular level. This level is often termed as the hydraulic head and is
actually the sum of three components - the pressure head, elevation head and velocity
head. The velocity head is often disregarded because ground water movement in most
cases is relatively slow. In practical applications, a depth to ground water measurement is
obtained and subtracted from the top of the well casing elevation to measure total head.
Note that the datum plane illustrated below is often calibrated to sea level.

Fig. 30
When ground water level measurements are collected for a specific aquifer and contoured
in a map perspective, a ground water surface map can be generated. Remember, water
level measurements obtained from an unconfined aquifer represents the water table
surface, and represent the potentiometric surface for water levels obtained from a confined
aquifer.
The direction of ground water movement can be understood in the fact that ground water
always flows in the direction of decreasing head. The rate of movement on the other hand
is dependent on the hydraulic gradient, which is the change in head per unit distance. The
change in head measurement is ideally in the direction where the maximum difference of
head decrease occurs. In the example below, the hydraulic gradient is determined to be
.00641 ft./ft. (the change in head divided by the change in distance). Notice the units are
foot by foot but can be described in more inconsistent units such as foot per mile.
102
102/113

Fig. 31
The direction and hydraulic gradient can be determined if the following data are available
for three wells located in a triangular pattern.
(1) Relative geographic position of the wells.
(2) Distance between wells.
(3) Total head (ground water level) at each well.
The solution is outlined and illustrated below.
(a) Identify the well that has the intermediate water level.
(b) Calculate where the elevation of the well with the intermediate head would fall along a
straight line between the wells having the highest and lowest heads.
(c) Draw a straight line from the well with the intermediate head to its elevation equivalent
plotted along the straight line between the wells having the highest and lowest heads. This
line represents a segment of the water-level contour whereby the total head is the same as
the intermediate well
(d) Draw a line perpendicular to the water-level contour line just plotted with the well of
highest or lowest head. This line is the line that parallels ground water direction.
(e) Now, calculate the difference between the head of the well and that of the intermediate
head contour by the distance between the well and the contour to reveal the hydraulic
gradient.
Thus, ground water flow direction is towards the south east at a hydraulic gradient of
.00641 ft/ft.

103
103/113

Fig. 32

7.4

Basic Equation of Ground Water Flow

Darcys Law and Hydraulic Conductivity


In the mid-1800s the French engineer Henry Darcy successfully quantified several factors
controlling ground water movement. These factors are expressed in an equation that is

commonly known as Darcy's Law.

By rearranging Darcy's Law and solving for hydraulic conductivity (K) in common units we
can get a sense of what hydraulic conductivity really represents

Thus, in practical terms hydraulic conductivity is the volume of water flowing through a 1 ft.
x 1 ft. cross-sectional area of an aquifer under a hydraulic gradient of 1 ft./ 1 ft. in a given
amount of time (usually a day). If we cancel out our units, we see that hydraulic
conductivity is usually expressed in ft./day

104
104/113

Hydraulic conductivity ranges approximately 12 orders of magnitude depending upon


differing water transmitting characteristics of aquifer materials. As one may conclude from
the chart below, the volume of water that can flow from our sandy and cavernous
carbonate aquifer materials of the coastal plain are in most cases far greater than our
rocks within the Piedmont and Blue Ridge provinces

Fig. 33
There are also some other terms used when regarding the spatial distribution of hydraulic
conductivity within an aquifer. If the value of hydraulic conductivity is the same in all
directions, the the aquifer is said to be isotropic. If hydraulic conductivity is different in
different directions, the aquifer is said to be anisotropic. The terms homogeneous and
heterogeneous are used when comparing hydraulic conductivity at separate locations of
the aquifer, regardless whether the value is the same or different in all directions. The
aquifer is homogeneous if hydraulic conductivity is the same and heterogeneous if
different. The diagram below shows the four possible combinations when describing the
hydraulic conductivity of aquifers

Fig. 34
Another important factor controlling ground water movement is its velocity. The ground
water velocity equation can be derived from a combination of the velocity equation of
hydraulics and our familiar friend, Darcy's Law.

105
105/113

We can solve for ground water velocity by; 1. substituting discharge (Q) by the crosssectional area at a right angle to flow direction (A) times velocity (v) and, 2. cancelling our
units. When we do this we discover that velocity (v) is equal to hydraulic conductivity (K)
times hydraulic gradient (dh/dl).

In practical terms we know that ground water does not move in open space, rather it
moves through aquifer materials that impedes ground water velocity. To account for this
we need to include effective porosity (specific yield) to accurately quantify ground water
velocity. Remember, hydrogeologists use effective porosity because this value better
represents water flowing through an aquifer under the forces of gravity. Thus the velocity
equation is modified

Let's take a look at the difference of ground water velocity when comparing characteristics
of a sandy substrate to that of a clay substrate. As we can see, ground water tends to
move quicker in materials with higher effective porosities and hydraulic conductivities

106
106/113

7.5

Transmissivity (T)

Transmissivity (T) is the volume of water flowing through a cross-sectional area of an


aquifer that is 1 ft. x the aquifer thickness (b), under a hydraulic gradient of 1 ft./ 1 ft. in a
given amount of time (usually a day). If we think about our definition of hydraulic
conductivity, we can conclude that transmissivity (T) is actually equal to hydraulic
conductivity (K) times aquifer thickness (b). Or otherwise denoted as T = Kb. We can also
conclude that transmissivity is expressed as ft2/day because if T = Kb, then T =
(ft./day)(ft./1). It is difficult to understand the differences between "T" and "K" when first
introduced to these terms - the illustration below should hopefully bring it all together.

Fig. 35
The differences in transmissivities in our coastal plain vary greatly. Some of our
Cretaceous age aquifers have transmissivites that are low as 100 to 1,000 ft2/day.
Conversely, transmissivity in the Eocene age Castle Hayne Limestone can be as high as
50,000 ft2/day!!!

107
107/113

7.6

Storage Coefficient (S)

You'll often hear hydrogeologists speak of the "T" and "S" of aquifers. We now have an
understanding of "T" or transmissivity. The "S" is used to represent the storage coefficient
of an aquifer which is the volume of water released from an aquifer per 1 foot surface area
per 1 foot change in head. Notice that we are not speaking of water flowing through an
aquifer, rather we are referring an aquifer's ability to store water. Mathematically, the
storage coefficient is dimensionless as the equation below illustrates:

The size of the storage coefficient is dependent whether the aquifer is unconfined or
confined. In regards to a confined aquifer, water derived from storage is relative to; (1) the
expansion of water as the aquifer is depressurized (pumped) and, (2) compression of the
aquifer. In a confined aquifer setting, the load on top of an aquifer is supported by the solid
rock skeleton and the hydraulic pressure exerted by water (the hydraulic pressure acts as
a support mechanism). Because of these variables, the storage coefficient of most
confined aquifers range from 10-5 to 10-3 (0.00001 to 0.001). Conversely, in an unconfined
aquifer setting, the predominant source of water is from gravity drainage and the expansion
of water and compaction of the rock skeleton is negligible. Thus, the storage coefficient is
approximate to value of specific yield and ranges from 0.1 to about 0.3.

7.7

Cone of Depression

As water is withdrawn from a well, the water level in the well begins to decline as water is
removed from storage in the well. The head in the well will fall below the level of the
surrounding aquifer and water begins moving from the aquifer into the well. The water level
will continue to decline and the flow rate of water into the well will increase until the inflow
rate is equal to withdrawal rate. Water from the aquifer must converge on the well from all
directions and the hydraulic gradient must get steeper near the well. For this reason the
resultant 3-D shape of water withdrawal is a called a cone of depression.

108
108/113

Fig. 36
The aquifer parameters of transmissivity (T) and storage coefficient (S) are variables that
can dictate the shape of the cone of depression. Relative to a confined aquifer, the
expansion of water in response to depressurizing (pumping) is very small, yet compression
of the rock skeleton is great. This permits the cone of depression to expand and deepen
rapidly when pumped. Thus, lower "S" values create deeper and wider cones than higher
"S" value aquifers that do not deepen and expand as readily. In regards to transmissivity,
aquifers of low "T" develop deep and narrow cones of depression and aquifers of high "T"
area characterized by shallow and wide cones. The reasons for this are fairly complex and
require more background information to be explained fully. However, think about the "T"
relationship this way - a greater decline in hydraulic head would be required to move water
through a less transmissive aquifer than a more transmissive aquifer.

7.8

Workout Examples

Example 7.1: At a certain point in an unconfined aquifer of 3 km2 area, the water table was
at an elevation of 102.00. Due to natural recharge in a wet season, its level rose to 103.20.
A volume of 1.5 Mm3 of water was then pumped out of the aquifer causing the water table
to reach a level of 101.20. Assuming the water table in the entire aquifer to respond in a
similar way, estimate (a) the specific yield of the aquifer and (b) the volume of recharge
during the wet season.
(a)
(b)

Volume pumped out = area drop in water table Sy


1.5 106 = 3 106 2.0 Sy
Sy = 0.25
Recharge volume
= 0.25 (103.20-102.00) 3 106
= 0.9 Mm3
109
109/113

Example 7.2: A field test for permeability consists in observing the time required for a
tracer to travel between two observation wells. A tracer was found to take 10 h to travel
between two wells 50 m apart when the difference in the water-surface elevation in them
was 0.50 m. The mean particle size of the aquifer was 2 mm and the porosity of the
medium 0.3. If v = 0.01 cm2/s, estimate (a) the coefficients of permeability and intrinsic
permeability of the aquifer (b) the Reynolds number of the flow.
The tracer records the actual velocity of water

Va

50 100
0.139cm / s
10 60 60

Discharge velocity V = n Va = 0.3 0.139 = 0.0417 cm/s

0.50
1 10 2
50
4.17 10 2
4.17cm / s
Coefficient of permeability =
1 10 2
K
Intrinsic permeability, K 0
g
4.17 0.01

4.25 10 5 cm 2
981

Hydraulic gradient i

Since 9.87 10-9 cm2 = 1 darcy


K0 = 4307 darcys

(b) Reynolds number Re =

Vd a
v

Taking da = mean particle size, Re =

0.0417 2
1

0.834.
10
0.01

Example 7.3: A 30-cm diameter well completely penetrates a confined aquifer of


permeability 45 m/day. The length of the strainer is 20 m. Under steady state of pumping
the drawdown at the well was found to be 3.0 m and the radius of influence was 300 m.
calculate the discharge.
In this problem, referring to
rw = 0.15 m
R = 300 m
sw = 3.0 m
B = 20 m
K = 45/(60 60 24) = 5.208 10-4 m/s
T = K B = 10.416 10-3 m2/s
Using following Eq.

110
110/113

2Ts w
2 10.416 10 3 3

300
ln R / rw
ln
0.15

= 0.02583 m3/s = 25.83 lps


= 1550 lpm
Example 7.4 A 30-cm well completely penetrates an unconfined aquifer of depth 40 cm.
After a long period of pumping at a steady rate of 1500 lpm, the drawdown in two
observation wells 25 and 75 m from the pumping well were found to be 3.5 and 2.0 m
respectively. Determine the transmissivity of the aquifer. What is the drawdown at the
pumping well?
(a)

1500 10 3
0.025m 3 / s
60

h2 = 40.0 2.0 = 38.0

h1 = 40.0 3.5 = 36.5 m

r2 = 75 m

r1 = 25 m

From the following Eq,

K h22 h12
ln r2 / r1

0.025

K 382 36.52

ln 75 / 25

K = 7.823 10-5 m/s


T = K H = 7.823 10-5 40 = 3.13 10-3 m2/s
(b)

At the pumping well, rw = 0.15 m

K h12 hw2
ln r1 / rw

0.025

7.823 10 5 36.52 hw2

ln 25 / 0.15

hw2 = 811.84 and

h= 28.49 m

Drawdown at the well, sw = 11.51 m

111
111/113

7.9

Exercises

Exercise 7.9.1 In a field test a time of 6 h was required for a tracer to travel between two
observation wells 42 m apart. If the difference in water-table elevations in these wells were
0.85 m and the porosity of the aquifer is 20%, calculate the coefficient of permeability of
the aquifer.
Exercise 7.9.2 A confined aquifer has a thickness of 30 m and a porosity of 32%. If the
bulk modulus of elasticity of water and the formation material are 2.2 105 and 7800 N/cm2
respectively, calculate (a) the storage coefficient and (b) the baro-metric efficiency of the
aquifer.
Exercise 7.9.3 A 45-cm well penetrates a 30 m unconfined aquifer completely. Under a
steady pumping rate for a long time the drawdowns at two observation wells 15 and 30 m
from the well are 5.0 and 4.2 m respectively. If the permeability of the aquifer is 20 m/day,
determine the discharge and the drawdown at the pumping well.

112
112/113

References
Bras, R. L., (1990). Hydrology, an Introduction to Hydrologic Science. Addison Wesley.
Chow, V.T., Maidment, D., and Mays, L. W., (1988). Applied Hydrology. McGraw Hill.
Ramrez, J. A., 2000: Prediction and Modeling of Flood Hydrology and Hydraulics. Eds.
Ellen Wohl; Cambridge University Press.
Bedient, Philip B. and Wayne C. Huber. Hydrology and Floodplain Analysis. Prentice-Hall,
Inc., Upper Saddle River, 2002.
Subramanya, K, 1984: Engineering Hydrology, Tata McGraw-Hill Publishing Company
Limited, New Delhi.

113
113/113

Anda mungkin juga menyukai