George M. Hornberger, Patricia L. Wiberg, Jeffrey P. Raffensperger, Paolo D'Odorico - Elements of Physical Hydrology-Johns Hopkins University Press (2014)
George M. Hornberger, Patricia L. Wiberg, Jeffrey P. Raffensperger, Paolo D'Odorico - Elements of Physical Hydrology-Johns Hopkins University Press (2014)
2
Elements of Physical Hydrology
SECOI'ID EDlTtO'll
George M. Hornberger
Patricia L. Wiberg
Jeffrey P. Raffensperger
Paolo D'Odorico
3
© 2014 Johns Hopkins University Press
All rights reserved. Published 2014
Printed in the United States of America on acid-free paper
98765432 1
A catalog record for this book is available from the British Library.
Special discounts are available for bulk purchases of this book. For more
il!formaliull. please colllact Special Sales at 4/0-5/6-6936 or
[email protected].
4
Contents
Preface
Appendixes
1 Units, Dimensions, and Conversions
2 Properties of Water
3 Basic Stalistics in Hydrology
Answers to Example Problems
Glossary
References
Index
5
Preface
6
notwithstanding, the book is still designed to accompany an undergraduate
course in physical hydrological science. It is aimed at upper-level
undergraduates majoring in environmental or Earth scicnces. The coverage
presupposes a modest background in calculus and physics.
We use several conventions in an attempt to make the book "user
friendly." Each chapter has introductory and concluding remarks that seek
to place the material presented in the chapter in the context of some
contemporary environmental issue. Tenns that appear in boldface at their
first occurrence are found in the Glossary. Terms that appear in italics
deserve emphasis. Supporting material is contained in three appendixes: a
review of units and dimensions (Appendix I ); a tabulation of certain
properties of water (Appendix 2); and a review of some elementary
statistical concepts (Appendix 3) .
Over the ycars we have benefited from ideas and data shared by many
colleagues and students. We are grateful to all who have contributed to our
education, but will refrain from attempting to produce a comprehensive
list. We do want to acknowledge specifically colleagues who either have
taught the course at UVA that stimulated the first edition of the book in the
past or who are currently in the teaching rotation for the course: John
Albertson, Keith Beven, Keith Eshleman, John Fisher, Peter German,
Aaron Mills, Matt Reidenbach, and Todd Scanlon. In addition, we
especially thank Margot Bjoring for the preparation of the figures for this
edition of the book.
7
Elements of Physical Hydrology
8
1 The Science of Hydrology
1.1 Introduction
1.1 Introduction
9
of intense scmtiny and speculation. In arid regions, the fair allocation and
wise use of freshwater resources are signifieant challenges facing
governments and people, affecting relations between nations, states, cities,
and individual users. As a resource, water appears unlimited. However, the
twentieth century saw a tremendous growth in the use of water, as well as
an increase in the threat of its contamination, and the trends have
continued in the new millennium.
Hydrological science has aspects related to "curiosity-driven" questions
and to "problcm-driven" questions. The first aspect relates to questions
about how the Earth works, and specifically about the role of water in
natural processes. The second relates to using scientific knowledge to
provide a sound basis for the proper use and protection of water resources.
Whence we may conclude that the water goes from the rivers to the sea and
from the sea 10 the rivers, thus constantly circulating and returning, and that
all the sea and rivers have passed through the mouth of the Nile an infinite
number of times. The conclusion is that the saltiness of the sea must
proceed from the many springs of water which, as they penetrate the earth,
find mines of salt, and these they dissolve in part and carry with them 10 the
ocean and other seas, whence the clouds, the begetters of rivers, never carry
it up.
10
flourished and, until about 1950, pragmatic considerations dominated
hydrology. Primarily due to the development and availability of digital
computing, thcoretical approaches in hydrology have incrcasingly allowed
hydrological theories to be subjected to rigorous mathematical analysis.
Freshwater resources are needed to meet the needs of humans,
livestock, commercial enterprises, agriculture, mining, industry, and
thermoelectric and hydroelectric power. Ln today's world, the necessity of
solving water-supply problems has become obvious, with many regions
exhibiting signs of looming water shortages (Figure 1 . 1 ).
Most of the human consumption of freshwater resources is associated
with food production. (We use cOl/slimp/ion to refer to water that is used in
ways that return it to the atmosphere rather than to a stream, river, or
groundwater. For example, thermoelectric power plants may withdraw
large quantities of water from a river for cooling, but then return the bulk
of that water to the stream, albeit at a higher temperature.) Relative to the
consumption of water for food production, drinking, household, and
industrial uses of water are overall smaller (Figure 1.2). We need much
more water to produce the food we eat than the amount of water we use for
drinking or other activities. The global water crisis is more likely causing
hunger than thirst. A report from the office of the U.S. Director of National
Intelligence, Global Water Security
(http://www.dni.gov/index.php/aboutlorganization/national-intelligence
council-nie-publications) concludes: "Between now and 2040, fresh water
availability will not keep up with demand absent more effective
management of water resources. Water problems will hinder the ability of
key countries to produce food and generate energy, posing a risk to global
food markets and hobbling economic growth." The need for improved
water management, based on scientific knowledge, is of critical
importance in the coming decades (Jury and Vaux, 2007).
II
r
·f ....
·
• •
o o
tt! ; ,/
asins:'"
;
Water stress indicator
�o""'!�'o'b
.. ., ......
Sou"*,, Smakht"" Ailvenga
.nd DOlI, 20001.
Figure 1.1 Water resources are highly stressed in many parts of the world.
The water stress indicator is the fraction of available water appropriated
for use by humans. Fractions greater than one indicate the use of fossil
groundwater.
http://www.grida.no/graphicslib/deta iI/water-scarcity-index_14f3
12
o Agriculture
o Household
9.6%
o Industrial
85.8%
Although the rate of increase in water usc in some arcas has slackened
over the past several decades, the total consumption of water globally has
continued to increase (Figure 1.3 ). Population growth, depiction or
deterioration of freshwater resources, and changing demands (mosl
notably the tendency to adopt more water-demanding diets in emergent
countrics,-i.c., an increase in meat consumption) will tend to further stress
water resources in the future for many countries. Falkenmark et al. (2007)
indicate their view of the severity of the issue. "We are on the verge of a
new and more serious era of water scarcity, and it is clear that we will face
increasingly complex challenges. Water supply to different sectors will
become more challenging as supplies of blue water (e.g. water in rivers
and aquifers) become overstretched, while a scarcity of green water (e.g.
water in the soil) will limit food and biomass production."
World population is expected to increase in the next 50 years, stressing
water resources. Issues related 10 the qualifY of water supplies have
occupicd an increasingly important niche in hydrology. Ii is estimated that
13
80% of all diseases and over one-third of all deaths in developing countries
result from the consumption of contaminated water. Provision of basic
sanitation and water treatment in much of the world is still lacking. [n
1980, the United Nations launched the International Drinking Water
Supply and Sanitation Decade, with the goal of clean water and sanitation
services to those without them. Despite enonnous effort, expense, and
progress, at the close of the decade 1.8 billion people still had no access to
sanitation services, and nearly 1 .3 billion people still laeked access to
clean water. Population growth wiped out the progress achieved through
this effort (Gleick, 1993).
2000
Europe
�
North America
�
-
>-
,
Africa
� 1500
E Asia
..
-
South America
<:
0 Australia and Oceania
a.
'-
1000
-
E
"
fI)
<:
0
u
Q)
�
500
C\I
-
;:
o
1900 1920 1940 1960 1980 2000 2020
Figure 1.3 Water consumption grew during the twentieth eemury and is
projected to continue increasing in the future.
Data from Shiklomanov 1999.
14
processes (their assimilative capacity) depends on the quantity of water
flowing in them.
Understanding surface-water now is a requisite for water quality
studies. Similarly, knowlcdge ofsubsurfacc flow is nccessary for
understanding the movement of pollutants underground. To predict or
evaluate the effects of liquid-waste disposal in deep injection wells or the
use of solid waste as fill for reclamation of strip mines, one must
appreciate the mcchanics of water flow in rock and soil. The massive
cffort now undcr way in the Unitcd States to clean up sitcs wherc
groundwater has been contaminated represents another challenge for
hydrological science.
Major floods are the most dramatic and visible of hydrological
phenomena. The number of people afef cted by floods and the number of
lives lost havc increased in recent decades (Figure 1.4). With pressure for
increased usc of floodplains, the prediction and control of floods remain
among the most important applications of hydrology.
Addressing hydrological challenges, such as those related to floods or
groundwater contamination, requires a finn understanding of the basic
principles of the physics and chemistry of water. Hydrological science
uses the fundamentals of the basic sciences and mathematics to develop
explanations (modcls) of obselVed phenomena. One of the basic problems
in hydrology is a description of water motion. One goal of physical
hydrology is to identifY the paths of water movement on and beneath the
surface of the Earth. Using physical theory and associated mathematical
models, hydrologists seek to describe quantitatively the motion of water in
the nafUral environment. As Jury and Vaux (2005) put it, "There is little
question that science must play a critical role in fonning a succ·essful
solution to the world's emerging water problems." They go on to conclude
that "a new science of sustainability will be needed if the prospects for
managing and solving the world's emcrging watcr problcms are to bc
bright." Knowledge of physical hydrology will be a critical component of
this new science.
15
1010
8 Number aHected
8 Number killed
"
10'
-
a.
0
"
a.
-
0 10'
"
�
.0
E
"
Z
10'
10'
1975 1980 1985 1990 1995 2000
Year
Figure 1.4 Losses during river flooding have trended upward over the past
several decades.
Redrawn from Jonkman 2005.
16
atmosphere occurs in vapor phase (evaporation from the ocean surface),
liquid phase (rain onto the ocean surface), and solid phase (snowfall onto
the ocean surfacc).
Solar energy drives the hydrological cycle; gravity and other forces
also play important roles. The dynamic processes of water vapor fonnation
and transport of vapor and liquid in the atmosphere are driven largely by
solar energy. Precipitation and the now of water on and beneath the
Earth's surface are driven primarily by gravity. Within partially dry soil,
gravitational and other forces are responsible for the movement of water.
The hydrological cyclc can be considered to "start" anywhere, but let us
consider atmospheric water first (Figure 1.5). As hydrology is concemed
mainly with water at or near the Earth's surface, from our point of view
the dominant process involving atmospheric water is the precipitation of
water on the land surface. The portion of the precipitation that reaches the
land surface as solid precipitation (mostly snowfall) can be retained
temporarily on vegetation or ground surfaces, or accumulate in seasonal
snowpacks or in pennanent snowpacks known as glaciers. Considering
liquid precipitation (rain), a portion also can be retained temporarily on
vegctation surfaces or in surface depressions, a portion enters into the soil
(infiltration) and a portion nows over the land surface first into small
rivulets and ultimately into larger streams and rivers. This last process is
called sllIface rtlnoff; which can be augmented by runoff during periods of
snowmelt (snowmelt runofl). The portion of rainfall that infiltrates into the
soil can also follow one of scvcral paths. Somc of the watcr el'aporates
from the soil and somc is returned to thc atmosphere by plants
(frallspira/iol/). We often refer to the total evaporation and transpiration
from vegetated land surfaces as evapotranspiration. The remaining water
continues to move downward through the soil and recharges thc saturatcd
portion of the subsurface, becoming groundwater. At lower elevations,
groundwater discharges into streams and rivers or directly to the ocean
(groundwater rllnoj}). Water evaporates from the surface of the oceans and
thereby replcnishes the water in the atmosphere. Thus we have returned to
thc particular compartmcnt that we considered first, atmosphcric water.
Looking at the relative magnitude of thesc watcr nuxes (Figure 1.6), wc
observe that over the continents evapotranspiration is on averagc smallcr
than precipitation. The fraction of precipitation exceeding
evapotranspiration contributes to surface and groundwater runoff.
Conversely, over the oceans evaporation exceeds precipitation. Therefore,
precipitation alonc is not sufficient to rcmovc from the atmosphere above
the oceans all the water vapor coming from ocean evaporation. In fact, the
atmospheric circulation transports part of this water vapor on land masses,
I7
thereby allowing continental precipitation to exceed the evapotranspiration
ratc.
Snow Rain
• • • • • • I I f II
• • • • • •
• • • • • • I I II
• • • • • • I .. II
I' I I
11111 11111
Transpiration
Evaporation
I I I I I
I I I I I
• • •
Evaporalion
· . . . .
I I I I I
-
• • • • • • • • •
Groundwater runoff
• • • • • • • • • • •
. .
• • • • • • • • • • • •
. . . . . . . ,
. . .
·
· . . . . . . . . , . , . . . .
. .
• • • • • • • • • • • • • • • • • •
• • • • • • • • • • • • • • • • • • • • • • • • • •
. , . .
.
, . . . . . . . . . . . - - -
. . . . . . . . . " . . . . . . . . , . " .
. .
. . . . . . . .
· . . . . . . . . . . . . . . . - . .
• • • • • • • • • • • • • • • • • • • • • • • • • • • • •
• • • • • • • • • • • • • • • • • • • • • • • • •
• • •
18
1 .7c). Thus at the regional scale the water cycle may have an important
internal component, whereby part of regional evapotranspiration
contributes 10 a substantial part of the regional prccipitation. Known as
precipitation recycling, this phenomenon is usually quantitatively
expressed in tenns of the recycling ratio, which is the fraction of the total
regional precipitation that is contributed by water evapotranspiring from
the same region. The recycling ratio is overall greater during the growing
season, when evapotranspiration is more intense, and it increases with the
region size. Studies on precipitation recycling provide important insights
into the impact of terrestrial vegetation on the water cycle. This research
has shown how the removal of forest vegetation (e.g., deforestation) may
affect regional precipitation by changing (decreasing) transpiration and
precipitation recycling.
li
'''' Ocean
So 39 moisture
-0
li
0.
�o
- 0 •
0
et
-
i�
-
• �
· . .
0
et°
• • •
Surface
38
• • • •
discharge
• • • •
• • • • •
• • • • • • •
• • • • • • • •
Groundwater 1
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
•
discharge
•
· . . . . .
• • • • • • • • • •
· . . . . . . . . . . . . . ,
• •
.
•
.
•
.
•
.
•
. .
• •
. .
• .�
.
- ��
• •
. .
• •
.
• • --c---:--,-
-
• �
.
. . . . . . . " . , . , " , . , " " , . , '
· . . . . , . . . . . . . . . . . . . . . . . , . .
. . . . . . . . . .
. . . .
, . . . , . . . . .,. . , . . . .
• •
.
• • • • • • • • • • • • • • • • • • • • • • • • •
. . . . . . . . , . , . . . . . . . . , . . . ,
Figure 1.6 Flows within the hydrological cycle. Units are relative to the
annual precipitation on the land surface ( 1 00 I 19,000 km3 yr-!). Black =
arrows depict flows to the atmosphere, gray arrows depict flows to land or
oceans, and blue arrows indicate lateral flows.
Data from Maidmcllt 1993.
19
principle of conservation of mass, which often is referred to as a water
balance or water budget when used in this way. A simple statement of
conservation of mass for any particular compartment (usually referred 10
as a COil/rot volume) is that the time rate of change of mass stored within
the compartment is equal to the difference between the inflow rate and the
outflow rate. For example, if we are adding two grams of water to a bucket
every minute and one gram of water is leaking out each minute, then the
mass stored within the bucket is increasing at the rate of one gram per
minute. Symbolically, we can write this as:
20
Atmospheric tr ;';:::
�.:---...,
sport
Precipitation
--1 ----- - -- ----
Evapotranspiration
,...._
.. _.=.:e:
o c�a:.:
n;;& land transport
(a)
Atmospheric transport
---- .. II __.
.... --
...... ./'
Evapotranspiration Precipitation
Runoff
(b)
Atmospheric transport
- --- .... ....
,... ... --
Recycled precipitatio
Precipitation
Runoff
Evapotranspiration
(e)
21
Figure 1.7 The water cycle at the global (a), local (b), and regional (c)
scales.
Modified after Eltahir and Bras 1996.
dAl
=1'-0'' (1.1)
d,
where M= mass within the control volume [M]; t = time [T] ; I' = mass
inflow rate [M il]; and 0' = mass outflow rate [M il]. The expressions
in square brackets are the mass-length-time dimensions associated with the
defined quantity; for example, the dimensions of I' are mass per time or M
,I. (See Appendix I for a discussion of units, dimensions, significant
figures, and unit conversions.)
In many instances, the density of water can be taken as approximately
constant and the conservation law expressed in tenns of volume. The lenns
involving mass in Equation 1 . 1 can be expressed in tenns of density times
volume and density can then be canceled from both sides of the equation.
Thus, Equation 1.1 can be rewritten:
dV =/-0
- ' (1.2)
d,
where V = volume of water within the control volume [L3]; I = volume
inflow rate [L3,1]; and 0= volume outflow rate [L3,1].
22
not change significantly. Ovcr much longer time periods such as centuries
or millennia this may not be true if there is a dramatic shift in climatic
conditions. Ir thcre is no change in storage over time, we say that the
system is at steady slare. For any given control volume at steady state, a
completely general water budget equation can be written (using bars over
the terms to indicate that they are annual average quantities):
dV _
- = 0= p+r" +rj{l-r..... -rl:" -et. (1.3)
d,
dV
p r
' O (104)
d,
�
- "" - - et""
0'
(1.;)
23
the flows of water into and out of them is the residence time. The
residence time, Tr [T], is a measure of how long, 011 average, a molecule
of water spends in that reservoir before moving on to another reservoir of
the hydrological cycle. The residence lime is easily calculated for systems
at steady state, when the inflow and outflow rates are idemical:
y
T, =-. (1.6)
I
= 3 1 0 mm, and eJ = 490 mm (Figure 1 .6). (Note that we now are referring
to thc volumes divided by the areas being considercd. It is sometimes more
convenient to use depth rather than total volume, because the volumes can
be quite large; also, we are probably more familiar with the statement, "20
mm of precipitation was recorded at Smith Airport," than "Smith Airport
received 20,000 m3 of water".) On average, 39% of precipitation to the
continents runs ofT and 6 1 % is rcturncd to the atmosphere through
evapotranspiration. In other words, the runoff ratio V)m is equal to 0.39.
The balance is, of course, affected by many topographic and climatic
factors and the budgets for individual continents can be quite different
from the average (Table 1 .2). The budget for North America is p = 670
mm I� = 290 mm and el = 380 mm Thus, in North America 43% of
, , .
Table I.t. Sizes and residence times for major reservoirs in the
hydrological cycle
24
Pe rcenuge Percentage of
Volume (kml) Residence time (yr)
of total freshwater
lakes
25
Table 1.2. Average annual water budget for the continents (excluding
Antarctica)
Runoff fatio,
Continent p(mm] r,(mmj rt (mm) r�/p
Note: Values for average annual preeipilation, runoff, and evapotranspiration arc
reported as depths of water over each land area. The total volume may be
calculated by converting the depths to km and multiplying by the land areas.
Several estimates of these quantities exist, all of which are uncertain to some
degree; Glciek (1993) presents a summary.
26
Let us consider application of Equation 1.2 to a catchment. In this case,
V is the volume of water stored on or beneath the surface of the catchment.
One inflow is precipitation falling on the catchment. By definition, thcre is
no surface water inflow into the catchment (this is the primary reason for
using the catchment concept in hydrology, as discussed). There may be
both inflow and outflow of groundwater across catchment boundaries.
However, if the assumption is made that the groundwater divide coincides
with the surface water divide, then groundwater inflows can be neglected
in the formulation of the water budget equation. This assumption is often
valid but should be evaluated for each application. Outflows consist of loss
to the atmosphere (via evapotranspiration) and discharge at the river
station chosen in defining the basin. The conservation equation for the
catchment may be written:
27
where V = volume of waler stored in Ihe catchment [L3]; p = precipitation
rate [L3 .1]; rs = rate of surface runoff [L3 . 1 ]; rg = net rate of
groundwater runoff [LJ .1]; el = rate of evapotranspiration [LJ .1].
We can illustrate the use of Equation 1.7 by considering long-term
average conditions. For questions involving long-range regional planning,
average quantities are often adequate. In this case, changes in volume of
water stored in the eatehmcnt can bc neglected. If, in addition, we assumc
net groundwater runoff to be negligibly small, the budget equation
becomes:
I� .11' yr
Perfon11ing the indicated calculation we find that ,:� is 385 nun yr- I over
the basin area. Equation 1.8 indicates that el = p - ' = 695 mm yrl or that
:.
about 64% of the rain that falls in this humid region of the United States is
returned to the atmosphere through evapotranspiration. II is small wonder
that evaporation control is of major concern in arid regions of the world.
The use of a water budget for annual average conditions is
straightforward and provides useful infonnation about climatological mean
values. One reason that these applications of the conservation laws arc
simple and straightforward is that in dealing with annual averages we have
been able to neglect temporal changes in storage. We have not had to
consider how soils store and release water or how channel storage changes
with the passage of a flood wave. An understanding of hydrological
28
phenomena at time scales of hours, days, or even months requires that
changes in storage be described and that the relationship between the
forces applied to water "panicles" and the motion of these "particles" be
studied. Thus, the dYllamics of water flow must be studied.
1 . 5 Concluding Remarks
29
1.6 Key Points
30
1.7 Example Problems
B. In gallons?
Problem 4. The polar ice caps (area = 1 .6 107 km2) are estimated to
x
contain a total equivalent volume of2.4 x 107 km) of liquid water. The
average annual precipitation over the ice caps is estimated to be 5 inches
per year. Estimate the residence time of water in the polar ice caps,
assuming their volume remains constant i n time.
A. What is the volume ofwaler (in m3) evapotranspired for the year
(assume no change in water stored in the catchment)?
B. What is the depth of water (in mm) evapotranspired for the year (again,
assuming no change in watcr storcd in the catchmcnt)?
31
1.8 Suggested Readings
Oki, T., and S. Kanae. 2006. Global Hydrological Cycles and World Water
Resources. Science 3 1 3: I 068-1 072.
Univ. of Oregon. Global Water Balance Animations.
http://geography.uoregon.edulenvchange/clim animationsl
V6rosmarty. C. J., P. Green, J. Salisbury, and R. B. Lammers. 2000.
Global Water Resources: Vulnerability from Climate Change and
Population Growth. Science 289:284-288.
32
2 Precipitation and Evapotranspiration
2.1 Introduction
2.2 Precipitation
2.1 Introduction
Chapler 1 introduced the concept of the hydrological cycle and bri efly
described some of the important processes involved in the motion of water
at or near the Earth's surface. This chapter explores two processes of great
importance to the fields of hydrology and meteorology: precipitation and
evapotranspiration. Precipitation is the primary input of watcr to a
catchment. Evapotranspiration is often the primary output of water from a
catchment. A detailed treatment of the precipitation process is most often a
33
subject of meteorology and is beyond the scope of this book. Our
discussion of precipitation will focus on methods for directly or indirectly
quantifying precipitation inputs to catchments and on describing thcse
inputs using statistical techniques. Evapotranspiration, on the other hand,
depends both on properties of the land surface and the state of the near
surface air and is, therefore, well within the domain of hydrology. Our
concern with thesc processes is in knowing the rates, timing, and spatial
distribution ofthese water fluxes between the land and the atmosphere.
A significant portion of precipitation that falls on vcgetated catchmcnts
is intercepted by and temporarily stored on the surfaces of vegetation (for
example, within dense forest canopies). The remainder falls through to the
land surface, where it may infiltrate into the soil or run off. Water stored
on the surfaces of plants and at the surface of bare soil, as well as at the
surface of open water bodies, can be returned directly to the atmosphere by
evaporation. Water that infiltrates into the soil may be taken up by roots,
converted from liquid 10 vapor in micropores on leaf surfaces (stomata) of
plants, and released to the atmosphere in a process referred to as
transpiration. As it is very difficult to keep track of the fractions of water
lost by the various pathways, we adopt the convention of referring to the
combined processes of evaporation and transpiration as evapotranspiration.
Evapotranspiration represents a dominant outflow of water from most
catchments and accounts for approximately two-thirds of precipitation
over most contincntal land masses. Understanding the physical processes
and factors that govern the rale, timing, and distribution of this flux of
water from the Earth's surface is, therefore, an important goal of
hydrological science. In practice, direct measurements of evaporation from
surfaces (such as leaves, soils, or ponds) at the scale of the catchment are
unavailable, and we must rely on indirect measures of evapotranspiration
or inferences from empirical techniques and the use of proxy variables.
Later in this chapter we explore the process of evapotranspiration and
discuss some techniques to calculate or infer rates of this flux from the
catchment.
There arc a number of reasons for studying the processes of
precipitation and evapotranspiration. The most apparent reason derives
from the catchment annual water budget. Equation 1 .8 tells us that for a
known or specified value ofp, there are infinite combinations of el and '"s
that will satisry Equation 1 .8. What processes ultimately govern the way in
which these two annual outflows are proportioned in a particular
catchment? What controls the relative proportioning on a seasonal or
individual stonn basis? How might changes in land use and climate
influence the water budget for a particular catchment (or for an entire
34
continent)? Will clear-cutting a forest in a catchment cause an increase or
decrease in et or in rs?
The historical development of hydrological concepts indicates the
importance of understanding both precipitation (the ultimate supply of
freshwater) and evapotranspiration (which can reduce the supply of
freshwater and affect its quality by concentrating impurities). The earliest
measurements of precipitation, attributed to Kautilya, an Indian chancellor
of the exchequer during the fourth century BC (Biswas, 1972), were used
as a basis for taxation because agricultural production was presumed to be
proportional to rainfall amounts. Knowledge of precipitation and
evapotranspiration is crucial to water resources and agricultural questions,
especially in arid regions.
One historical example illustrates just how important it is to know the
magnitude of hydrological fluxes. In the early part oflast century, rapid
growth in the western and southwestern United States led to efforts to
"reclaim" the desert, mostly through management of the Colorado River.
To apportion the flow of the Colorado among the states that would use the
water (the Colorado River Compact of 1 922), it was necessary to
determine the amount of water available each year (i.e., p - el). This was
done by averaging the annual discharge measured at a single point on the
Colorado River over the available period of record ( 1 896-1921), which
turned out to be about 1,233 m3 eaeh year. Unfortunately, this period of
time turned out to be a particularly weI era (or, the following years were
particularly dly). From 1 922 to 1 976, the average annual discharge of the
Colorado River at the gaging station was 1 ,020 m3. When the budget was
calculated for the Colorado River Compact of 1922 and the water
apponioned among the states, there was not enough waler to go around! [t
should be clear that the temporal and spatial patterns of precipitation and
evapotranspiration within the Colorado River basin strongly influence
water availability and, hence, its use and management.
2.2 Precipitation
35
and solid). Precipitation is the dominant dcposition mcchanism by which
atmospheric moisture is cycled from the atmosphere to both oceans and
continents. Other mcchanisms (i.e., dircct dcposition of dcw and fog) may
be important in some instances, especially in coastal mountains.
There are three primary steps in the generation of precipitable water in
the atmosphere: creation of saturated conditions in the atmosphere,
condensation of water vapor into liquid water, and growth of small
droplets by collision and coalescence until they become large enough to
precipitate. Saturated conditions occur whcn thc air (which is a mixturc of
gases, including water vapor) has the maximum water vapor content it can
hold without the emergence of condensation (e.g., dew or fog). Water
vapor content can be expressed in tenns of water vapor density (or
absolute humidity), which is the volume of water vapor per volume of air.
However, in mierometeorology and hydrology, the partial pressure of
water vapor, or vapor pressure, e [M L-1 J2], which is proportional to
the vapor density, is typically used as a measure of how much water vapor
is in the air. The saturation vapor pressure, e.m [M L-I J2], is the value
of e for saturated conditions. This value strongly depends on the air
temperature; that is, wann air can hold more water vapor than can cool air,
or, conversely, cool air cannot hold as much water vapor as warm air
(Figure 2. 1). Thus, cooling an air mass tends to produce saturated
conditions, because e.<a/ is reduced. Cooling typically occurs when an air
mass is lifted vertically. Uplift can happen in a number of ways, such as
when (a) air masses rise over mountains or other topographic features
(referred to as orographic uplift), (b) wann air masses rise above cooler air
masses at fronts (frontal uplift), or (c) when heating of the Earth's surface
(especially during the summer) makes air near the surface less dense so
that it rises (convcction) and cools, often producing thunderstonns.
Condensation is simply the phase change whereby water vapor becomes
liquid or solid water. It requires not only the creation of saturated
conditions within the air, but also the presence of condensation nuclei,
small particles (size of 0.00 1-10 jlm), such as dust, pollutants, smoke from
burning biomass, sea salts or prcviously fornled water droplets or iee
particles. Cloud condensation nuclei provide surfaces on which the
condensation of water molecules can start. Condensation may produce
such small particles that they remain stable in the atmosphere. The while
clouds observed on a fair day, for example, arc composed of water droplets
that arc too small to precipitate. It is the coalescence of small droplets into
larger drops, through collision of small droplets with each other or with
larger drops that gives rise to precipitable raindrops or ice crystals. To
36
generate substantial amounts of precipitation, clouds also need to draw
atmospheric moisture from the surrounding air.
�
-
"
II) Supersaturated
II)
Q)
�
Co
Undersaturated
Temperature, T •
37
land areas is derived from local evapotranspiration, a process known as
precipitation recycling (sec Chapter 1 ). Although this indicates that local
evapotranspiration docs influence local precipitation, much of the
precipitated water must be transported significant distances across the
continents from the oceans. It is the large-scale motions of the atmosphere
that are responsible for the broad patterns that are observed in annual
precipitation.
Average annual precipitation onto the continents is geographically
extremcly variable, reflecting the influencc of a number of important
physiographic factors. For example, in the United States, average annual
precipitation ranges from a minimum about 40 mm yr- 1 at Death Valley,
California (in the Mojave Desert), to a maximum of nearly 12,000 mm yr-1
at the summit of Mt. Waialeale on the island of Kauai in Hawaii. The
lowest average annual precipitarion that has been measured anywhere on
the Earth is less than I mm yr- 1 at Ariea, Chile (the 59-year record
includes a period of 1 4 consecutive years totally devoid of precipitation).
In general, average annual precipitation onto the continents is a function
of: (a) latitude (precipitation highest in latitudes of rising air-O° and 60°
north and soulh�and lowest in latitudes of descending air-30° and 90°
north and soulh); (b) elevation (precipitarion usually increases with
elevation, a phenomenon known as the orographic effect); (c) distance
from moisture sources (precipitation is usually lower at greater distances
from the ocean, a phenomenon known as continentality); (d) position
within the continental land mass; (e) prevailing wind dircction; (t) relation
10 mountain ranges (windward sides typically cloudy and rainy, with
leeward sides typically dry and sunny, a phenomenon known as rain
shadow); and (g) relative temperatures of land and bordering oceans
(Eagleson, 1970). Average annual precipitation onto the oceans is thought
to bc similarly variablc.
Ovcr much of the world precipitation is also extremely variable in lime,
with infrequent large stonns (e.g., hurricanes and monsoons) delivering
large portions of the total annual precipitation. As with the spatial
variability of precipitation, the temporal variability has both random and
persistent components. For example, although the arrival of individual
hurricanes is not predictable far into the future, the season in which they
arrive is consistent from year to year. Accurate forecasts of how many will
appear, the amount of rainfall they will produce, and the path they will
follow cannot be made for any given year.
38
Two problems arise in quantifying precipitation input to a given land area:
how to measure precipitation at one or more points in space and how to
extrapolate these point measurements to determine the total amount of
water delivered to a particular land area. A variety of instruments or gages
are used to measure point precipitation (amount of precipitation deposited
at a particular station representing a point in space). Point precipitation
usually is expressed in depth units (volume divided by collector cross
sectional area). Point measurement devices arc generally of two types:
non-recording (storage) gages and recording gages. The first category
includes simple wedge- or funnel-shaped containers to collect precipitation
over a period of time between observations. There are two common types
of recording precipitation gages: weighing and tipping bucket gages.
Weighing gages collect precipitation, typically through a funnel and rccord
the weight of precipitation as a function of time. Tipping bucket gages
consist of a container with a funnel at the top leading to a pair of small
"buckets" attached to a fulcrum. When one bucket fills with water, it tips,
emptying the water and moving the other bucket bcncath the outlct of the
funnel. The device records the lime at which the bucket tips, and so the
time over which a certain amount of precipitation fell (e.g., 0.25 mm for a
bucket ofa certain size) is recorded. In regions that receive snowfall, point
gage data usually are expressed in liquid water equivalent (LWE), which is
the cquivalcnt amount of total precipitation if it had all fallen as rainfall.
Most precipitation gages are equipped with windbreak devices or shields
to minimize the measurement error caused by disruption of the airflow
pattern around the gage.
Rainfall rates measured at a point are tremendously variable in time
(Figure 2.2). The record of hourly precipitation over time (known as a
hyetograph) for a station in Virginia illustrates that precipitation
commonly is organized into discrete stonn events of varying intensity and
duration separated by interstonn periods of variable duration. Because we
may be interested in longer periods of time (perhaps an annual water
budget for a catchment), we can add hourly measurements together to
derivc daily or monthly hyctographs. Variability decreases as the period of
reporting increases, as illustrated by the daily and monthly hyetographs for
the Coweeta experimental watershed for the 1992 water year (Figure 2.3).
(In the United States, the U.S. Geological Survey water year begins
October I and runs through the end of the following September.) Both
annual prccipitation dcpths and temporal precipitation variability depend
on local climatology. For example, in monsoon-prone regions most of the
precipitation arrives in a certain season when large-scale atmospheric
flows drive moist air from oceanic areas onto terrestrial areas. In rain
39
forests and some coastal areas the precipitation is spread more evenly over
the year.
6 ,------,
5
E
-
E
�
c 4
0
-
CO
.
-
-
.
a. 3
.-
u
..
�
a.
,., 2
"
-
0
J:
1
O+rTT'"n
M m �
N M
• � � . �
01 0) 0) Ol
� �
� � � �
< < < <
Figure 2.2 A hyctograph for a site near Charlottesville, Virginia, showing
temporal variability in precipitation.
Data courtesy of Greg V. Jones.
40
�
�-
E E
-
" E
'"
E 0 -
E
"
-
..
-
-
a.
-
.-
u
.,
.-
a.
�
,..
..
-
-
Cl
� N � � � � m � �
• • • • • • • • • • •
§ � �
� <
c .c � ""
8 � � � � � � • '
Figure 2.3 Daily and monthly (inset) hyetographs for Ihe Coweeta
experimental watershed, North Carolina, water year 1992.
Data courtesy of Wayne Swank.
41
Thus far, we have discussed temporal variations in precipitation at a point
or station in space. In hydrology we typically want to estimate the total
volumc of water delivered to a given area (e.g., a catchment) within some
period of time. That is, we want to know the average precipitation depth
over the area, i.e.,
42
.s, .s,
s,
s ,·
'----.I
S3
'-- __
_�
. S3 .87 • S,
•
l '
p = - "" a/p," (2.2)
•
A L-
, ,
.
where PI (pj_ + 1'1+)12, the average p for each subregion; Pi- = value of
=
43
radar data become available on a routine basis, they may supplant the
weighting schemes outlined above, or at least provide a rational basis for
constructing isohyctal lincs.
WCalher radar has become an increasingly important tool for cstimating
the spatial distribution of rainfall, p(x, y). Radar signals reflect from
raindrops in the atmosphere and the characteristics of the reflected signal
can be related to rainfall rates. Radar is far from an absolutely accurate
measurement method, but it provides detailed infonnation on the time and
spacc distribution of rain and can bc particularly valuable for heavy
rainfall. The spatial distribution of total precipitation in a storm may
depend in part on the path of the storm (Figure 2.5). We also see from
these data the potential for significant spatial variability in precipitation
over a catchment. Imagine the difTerent estimates oftola] precipitation
from a single gage located in the southcrn end of the catchment as
compared with one in the northern end.
Measurements from satellites also may be used to infer precipitation
rates and snow-accumulation patterns. Snow-cover maps are produced
from data collected with NOAA's (National Oceanic and Atmospheric
Administration) Advanced Very High Resolution Radiometer (AVHRR).
These satellite data are used in conjunction with ground-based point
measurements to detennine spatial distributions of water in the snowpack
(e.g., see DeWalle and Rango, 2008). Whereas the satellite observations
estimate amounts of snow present on the ground, techniques are being
developcd to cxtract snowfall ratcs from the NEXRAD data. NEXRAD
sites are being installed presently around the United States, providing
coverage of most regions. With the data from these instruments being
made available over the Internet, it is reasonable to anticipate that they will
soon be applied widely in hydrology to estimate rainfall and snowfall rates
distributed in space and time over catchments.
44
a
./
10:17 a.m.
1 1 : 38 a.m.
45
(storm) events of varying amounts (storm depths). Average precipitation
intensity is the rate ofpreeipital'ion over a specified lime period, the
preeipilation depth divided by the time over which that depth is recorded.
For example, the data in Figure 2.2 are reported for each hour (i.e., the
hourly precipitation intensity). The hourly precipitation intensity for the
period shown varies from zero to about 22 mm hel. Average precipitation
intensities depend on the time period over which the computation is done.
That is, the variation in hourly precipitation intensity typically will be
much greater than the variation in 6-hour intensity (average over a longer
time), but less than that of 1 5-minute intensity (average over a shorter
time). Average precipitation intensity for a storm is the total depth of
precipitation for the stoml dividcd by the storm length. In general, the
longer the storm-event duration, the less the (average) storm intensity.
However, the greater the stoml duration, the greater the storm depth.
Hydrologists commonly employ a statistical technique known as
frequency analysis to describe systematically the temporal characteristics
of precipitation at a particular station. To illustrate the frequency analysis
technique, we will consider annual precipitation amounts for several
United States cities over a 64-year period (Table 2. 1). We assume that the
quantities in Table 2. 1 (annual precipitation) are samples of a random
variable. Ifwe look at how often (the frequency with which) values within
a certain range arc encountered, we find that a plot of magnitude versus
frequency displays a characteristic shape, a "bell-shaped" curve. In the
case of the data from Seattle (Figure 2.6), we find that the most frequent
values of annual precipitation are between 950 and 1000 mm. Higher and
lower values occur with lower frequency. We also could normalize these
data by dividing by the total number of observations (64) to determine the
relative frequency. For example, annual precipitation values between 950
and 1 000 mm occur with a relative frequency of 10/64, or 0.16.
Alternatively, we might say that values within that range occurred 16% of
the time. Note that, because annual precipitation is a continuous random
variable, we need to be concerned with the probability of a range of
values, rather than with one specific value. Also, note that these data
represent yearly precipitation amounts. As we will discuss shortly,
decreasing the averaging period (say, monthly or daily precipitation
amounts) often causes the shape of the curve to become more skewed and
look less like a bell.
Because annual precipitation amounts appear to follow a normal (or
"Gaussian") distribution, the mean and the standard deviation (Figure 2.6,
Table 2. 1 ) are the only parameters that are required to describe the
magnitude-frequency relationship. Once we have detennined the
46
parameters of the nom1al distribution (the mean and the standard
deviation), we can detcrmine the probability associated with a particular
range of annual precipitation values (Appcndix 3). As an example, we will
calculate the probability that annual precipitation in any given year will
exceed 1.0 m (this is referred to as the exceedance probability). The first
step is to calculate the z-value, which is a way of normalizing the data to a
distribution with zero mean and unit standard deviation. This is
accomplished by finding the difference between the sample mean and the
target value ( 1 .0 m in this example) and dividing this value by the sample
standard deviation:
x-x
== • (2.3)
'..
; = 392 x = 973
J.. _ 96 05.. _ 1 66
Denver
(1948-201 1 )
Seattle
(1 948-201 1 )
where x = the value of the variable (P) Ihat we are interested in (e.g., 1 . 1
m); x = the mean ofp; s = the standard deviation ofp. For the Seattle
..
47
data, the calculated z-value corresponding tox = 1 . 1 m is 0.77. For this z
value, the cumulative probability (Table A3.2 in Appendix 3) is 0.7794.
This valuc is thc probability that thc annual prccipitation will bc less than
1 . 1 m, so we need to subtract 0.7794 from 1 . 1 (the total probability) to
arrive at a value of 0.2206. Therefore, there is a 22% chance that annual
precipitation will exceed 1 . 1 m in Seattle. Another way of looking at the
same result is to say that precipitation in excess of 1 . 1 m occurs roughly
evcry 4-5 years or with a return period (the inversc of the cxcecdance
probability) of approximately 4-5 years.
Magnitude-frequency relationships for many hydrological variables are
useful to hydrologists and water resources planners. For example, in the
design of a reservoir for irrigation or hydropower it may be important to
know how ortcn part of the reservoir's storage capacity would remain
unused. To that end, hydrologists usc the methods described in this
subsection to calculate the probability that the annual rainfall in a region is
smaller than the minimum value required to reach full storage conditions
in the course of the water year. Although annual precipitation often is
describcd wcll by a normal distribution (Figure 2.6), not all hydrological
parameters are (see Haan, 2002). In fact, exccpt for annual totals
precipitation intensities (or amounts) typically do not follow a nonnal
distribution. We have restricted our application of the nomlal distribution
to annual precipitation amounts. If the precipitation averaging time is
decreased to a much shorter period, the distribution begins to diverge from
nonna!. This is shown in Figure 2.7 for a 50-year record of precipitation
on the southern coast of Ireland. In this figure, probability density is
analogous to a relative frequency in that it is a relative measure of how
likely certain magnitudes of precipitation are. Notice the extreme skewness
in the daily precipitation distribution, the moderate skewncss in the
monthly amounts, and the rclative absence of skewness in the annual
amounts. Hence, it would be incorrect to use the nonnal distribution to
describe a variable such as daily precipitation at this location. It should be
stressed that the probabilistic approach requires a significant amount of
data (a long time series) to produce meaningful analyses.
Table 2.1. Total annual precipitation (mm), 1948-20 I I , for selected cities
in the United States
48
Year Seattle Santa Barbara Oenver Chariotte§vllie
49
1983 1040 1035 514 1409
50
'.2..-- ----- --,--, 0.01 -,---------
. ( b)
-;,-
,.,
-
(� •
� 0.15 •
<
� �
� �
:0�
;:
-
0.1 ;: 0.005
0.05
'\
� , �::::p-.,.�
o 10
.
ro � 40 SO
,-f-'---
o
100 150SO
-c--r-r-�.....j
250 300 350 200
Dalty precipitation (mm) Monthly precipitation (mm)
�'o:.----7---,=,-----(/;c'il)
:; 0.0015
•
o.OOO:-J-.c.:;:::::.-��_�_"'�>-�
1i 0.001
£
Jl
51
as the return period, Trc1llnp of X* and is typically measured in years. The
return period of an event, X*, tells us that X* is exceeded on average every
Trc tum years. Based on this definition, the return period increases with
increasing values of X*. In other words events exceeding X* become more
rare as X* increases. It should be stressed that the definition of return
period is given as the mean recurrence interval, t, which means that the
rainfall event with return period of, say, 100 years (sometimes called "the
1OO-year rainfall") is exceeded on average (but not exactly) once every
100 years. We can have some I OO-year intervals with more than one event
greater than the hundred-year rainfall and others with none. It can be
shown that if events with X> X* are statistically independent, l/T,.cmr M is
equal to the exceedance probability of X* (i.e., Prob[X> X*] = I/Trctum;
Appendix 3).
The study of extreme rainfall events faces two challenges: First,
extreme rainfall events are by definition rare and the rainfall records that
are typically available to detennine their return period are often very shon.
For example, we might need to study events with the return period of 100
or 200 years, while having access to records much shoner than these return
periods. Second, as noted in the previous section, event-scale precipitation
does not exhibit the "bell-shaped" Gaussian distribution that we have
considered in the ease of annual precipitation. The propenies of the
distribution ofX arc in gencral unknown and depend on thc timc scalc !:J.t
at which Xis measured (see differences between daily and monthly
precipitation in Figure 2.7). Statistical theories, however, have shown that
the extreme values of any random variable tend to follow only three
possible classes of distributions. In particular, extreme precipitation is
typically distributed according to the Gumbel (or "Type I") distribution.
52
140
E 120
-
E
-
100
.c
T
a.
-
:
----- ,------ --- ! ! -- -"----
� BO
"C
-
-
�
- 60
c
�
.-
�
40
,.,
-
�
20
.-
0
1958 1962 1967 1972 1977
Year
Figure 2.8 Record of daily precipitation depth, X, for the city of Turin.
The return period of the extreme value X· = 88 mm d-' (dashedhorizollfai
line) is the mean of the recurrence intervals 'I' T2, and so on.
ProliX';?:X.]= I
T...,,,,,,
= l-CXP [ ( X'-f3)]
-cxp -
a
(2.4)
Equation 2.4 shows that the Gumbel distribution h�nly two parameters,
a and /3, that need to be estimated using the mean (X,) and the standard
deviation (ox) of X' as
We can use the precipitation record from the city of Turin to examine
this method. For every year we consider the maximum rainfall depth with
a duration of 1 , 3, 6, 12, and 24 hours (Table 2.2). To detennine the return
period of a rainfall depth, e.g., X· = 70 mm with the duration of 3 hours,
we first need to calculate the mean, X', and the standard deviation, 0x� of
53
the annual maxima of that duration (Table 2.2, third column). Using
equations (2.5) we detennine the corresponding values of a and P for the
3h duration (sec Table 2.2) and use Equation 2.4 to calculate the return
period of X*. We find that 70 mm of rainfall in 3 hours is exceeded on
average once every 40 years (Trelum = 40 yr). Alternatively, we might need
10calculate the rainfall depth corresponding to a given return period.
Solving Equation 2.4 for X* we obtain
(2.6)
where InO denotes the natural logarithm. Using Equation 2.6 we find that
the rainfall depth with a 3 hour duration and a return period of 100 years is
X* = 79.5 mm. As expected, the rainfall depth associated with the same
return period (Trelllrn = \00 yr) increases with increasing rainfall duration
(see Table 2.2).
54
Year lIt=1h lIt=3h lIt=6h at=12h lIt=24h
55
1970 18.0 26.4 27.6 28.2 46.8
56
2006 31.0 34.8 46.7 64.4 106.9
,.=100yrs)
X· (T",.. 69.5 79.5 91.3 115.1 151.0
Note: These values are used to calculate the mean, standard deviation, and the
parameters, a and P, of the Gumbel distribution, along with the rainfall values, X*,
with lOO-year return period. All values are expressed in mm.
2.3 Interception
Not all precipitation reaching the land surface is available for streamflow
or replenishing groundwater. Rather, a portion is temporarily stored by
vegetation (interception), where it is subject to evaporation. If we
consider the total amount of precipitation (gross precipitation) delivered to
a point within a land area (p), such as might be measured by a tipping
bucket gage placed in a clearing or above a forest canopy, some of that
precipitation in a vegctated area will be intercepted by the plants, some
will fall between plants to land on bare ground or ground covered by lower
vegetation or leaflitter, and some will run down the stems and trunks of
plants to the ground surface (Figure 2.9). Precipitation stored by the
canopy or leaf litter is subject to evaporation. We can define the total
interception (IT) as the sum of the canopy interception (lJ and litter
interception (I,). It is important to stress that, according to this definition,
interception is associated with the evaporation of rainwater stored in
canopy and litter. Interception should not be confused with the temporary
retention of water in canopy or litter before it eventually drips to the
ground.
The capacity of the canopy and litter to store water is limited, and the
rain interception and snow interception capacities are not the same value.
After a period of time during a stonn, these "reservoirs" will begin to
approach their limits, such that no additional water can be stored.
Precipitation that is not intercepted by the canopy or which leaves the
temporary storage of interception may be classified as either stemnow or
throughfall (see Figure 2.9). The fonner is exactly what the name implies
57
and thc latter refcrs to watcr that rcachcs the ground dircctly or by dripping
ofT leaves. This partitioning of prccipitation between vegetated surfaces
and thc soil is critical in dcfining thc reservoirs of watcr availablc for
direct evaporation from surfaces and for transpiration by plants of water
taken up from the soil through roots.
Not all precipitation reaches the ground surface in vegetated areas, such
that the combination of throughfall and stcmnow is less than the total
precipitation. Water storcd in vcgctation (Ie) or lcaf litter (I,) may return to
the atmosphere through evaporation.
2.4 Evapotranspiration
58
Evaporation is a physical process that occurs whenever a wet surface is
exposed to unsaturated air. Transpiration is the vaporization and transport
of plant water from leaf chloroplasts to the atmosphere through the
stomata, small cavities existing on leaf surfaces. It is difficult in practice
to separate evaporation (from wet surfaces) and transpiration (water
evaporating inside plants) from each other. Hence, we focus on the
combined quantity, evapotranspiration (et). As we saw in Chapter I,
looking at a global average, about two-thirds of the precipitation that fal1s
on the continents is evapotranspired. Of this amount, 97% is el from land
surfaces and 3% is open-water evaporation. It is through the process of
evapotranspiration that the sun's energy is introduced to drive the
hydrological cycle.
Both evaporation and transpiration require a diffusion mechanism (sce
2.4.2) to sustain a water vapor flux that removes vaporized water
molecules from the evapotranspiring surface. The two ingredients for
vaporizing water are energy and water. Hence, evapotranspiration is where
the surface-water balance and surface energy balance meet. The energy is
derived from the solar radiation and the water is typical1y provided by
local precipitation. Because both solar energy and available water are
necessary to cause evaporation (and transpiration), energy will limit the
rate of evapotranspiration at some times, and water availability will limit
the rate at other times. Considering water availability, we see that
evapotranspiration is a two-way street, where the amount of water present
affects the rate of evapotranspiration, which in tum affects the amount of
water present for subsequent allocation. When the near-surface soil is
saturated or nearly saturated with water, evapotranspiration may proceed at
a rate-known as the potential evapotranspiration rate-limited only by
the availability of energy. When the soil becomes drier the actual
evapotranspiration rate becomes reduced below that found for a wet
surface.
Most plants have openings (stomata) on their leaves to allow them to
take up carbon dioxide from the atmosphere. When the stomata are open,
plants transpire water. Unlike evaporation, transpiration is not control1ed
solely by physical conditions because plants regulate the rate at which
water is released in transpiration in a manner that varies by plant type. Of
the water taken up by plant roots about 95% is transpired through the
stomata. The remaining 5% or so is converted to biomass through
photosYnlhesis. Hence, to first order, the water taken up by the roots is
converted to vapor and lost to the atmosphere. When the availability of
soil water is limited, plants conserve it by restricting flow to the
atmosphere by contracting the stomata. However, the degree of restriction
59
varies considerably across plant species and even throughout the year for a
given species. This renders quantitative treatment possible only in an
average sense, say with an entire stand of trees trealed as a single "big
leaf' for which the resistance 10 water flow is handled as a simple function
of soil moisture content and time of year. Here hydrologic and ecologic
conditions must be explored simultaneously, to incorporate a quantitative
description of plant water use characteristics into the framework of
hydrology.
The direct measurement of evapotranspiration is a difficult task and in
most hydrological applications the rate of evapotranspiration is calculated
as a function of quantities that are easier to measure. To that end, we can
in principle use methods based on the water balance, the energy balance,
or the diffusion of water vapor (mass-transfer methods). As it will be
shown in the following subsection, water balance methods have poor
accuracy. Thus, energy balance and mass-transfer approaches are most
commonly used. To eliminate the dependence on some variables that are
difficult to measure, these two methods are often combined (combination
methods).
dV
P(/)+ r.,(t)+ 'gi(/) -,..,(1) - '.>"'0) - el(/). (2.7)
dt
0::
In this equation, the time rate of change in volume of water stored within
the control volume (left side) is balanced by the difference between the
inputs and the outputs. Each of the tenns on the right side of the equation
has units of volume per time [LJi1], and all are functions of time.
Equation 2.7 may be simplified for some problems, for example, by
neglecting rgi and rgo when considered unimportant. Writing Equation 2.7
for average annual quantities reduces it to Equation 1 .3. Once all the tenns
except et have been measured, then et may be computed as the residual to
force a balance that conserves mass in the system. An example of the use
of this equation for an entire calchment is presented in Chapter I. The
result is a very reasonable estimate of average annual evapotranspiration,
el, for the catchment. In many cases, however, the water balance approach
may suffer from the accumulation of errors in the measured variables. As
60
an example, consider a lake with a very rapid throughflow rate, which over
the course of a year does not see a net change in storage. Suppose that the
measurcd avcrage annual fluxes and associated errors for the lake arc:
Water vapor fluxes from the soil surface or through the stomata typically
occur by diffusion, the same process that is responsible for the spread of
contaminants in streams and lakes or the transfer ofheat through rock
surrounding an igneous intrusion. In a diffusive process the flux takes
place in the direction of decreasing concentration of the transported
quantity and the rate of transport is proportional to the concentration
gradient (i.e., the difference in concentration between two points in the
flow direction divided by the distance between the two points). In the case
of evapotranspiration the transported quantity is water vapor and its
concentration (i.e., mass of water vapor per unit volume, or water vapor
density) is proportional to the vapor pressure (see section 2.2). Therefore,
evapotranspiration is a function of the vapor pressure difference between
the evapotranspiring surface(s) and the overlying air (a) at a height Za
above the surface (Figure 2.1 0)
61
T,/ e
"
H,o
(2.8)
where eo and II" arc the vapor pressure and the wind speed at the height zO'
respectively, and es is the vapor pressure of the air in cont(lct with the
evapotranspiring surface. Because this air is saturated with water vapor
(i.e., eJ = esat), eSal can be calculated as a function of the surface
temperature, TJ (see Figure 2. 1 and Table 2.3).
The dependence of et on Jill reflects the fact that diffusion in a turbulent
atmosphere is morc effective with stronger wind speeds. (Turbulence is
discussed in Chapter 3.) In general the diffusion of water vapor from the
evaporating surface depends on the aerodynamic properties of the surface
and is enhanced by surface roughness. The dependence on roughness is
expressed through the K£ factor in Equation 2.8, which, in the case of
evaporation from a wet surface (i.e., without accounting for the effeet of
stomatal regulation) and under neutral conditions (i.e., in the absence of
convection or subsidence), can be expressed as
62
(2.9)
Table 2.3. Values of air density, saturation vapor pressure, tatent heat of
vaporization, psychrometric constant, and slope of saturation vapor
pressure curve as a function of air temperature
63
T ['e) p� [kg m41 e"" [kPal l..[kJ krl y(kPa'C-<1 So (kPa 'C-<I
64
33 1. 153 5.032 2.423x 101 0.0676 .
0 282
where Po is air density (sec Table 2.3), Pw is liquid water density, 1nO
indicates the natural logarithm, k = 0.4 is the VOIl Karman's universal
constant (dimensionless), Po is the atmospheric pressure, Zu is the height at
which "a and e" arc measured (Figure 2. 1 0), Zo is the roughness height, a
parameter that accounts for the roughness of the surface (Table 2.4); and z"
is a parameter known as displacement height, which accounts for the fact
that in the presence of a forest canopy the surface over which the wind
blows is not at the ground level but displaced by a height that is a function
of the tree height, II (i.e., zd:::: 0.65h) (Campbell and Nonnan, 1998). In the
case of transpiration the diffusion of water vapor from the stomata is
controlled by the plant, which can reduce the opening of the stomata to
limit water vapor losses in conditions of limited soil water availability.
Stomatal regulation can, therefore, reduce the value of K£ with respect to
the values given by Equation 2.9.
As expected, looking at Equations 2.8 and 2.9, evapotranspiration rates
are greater in the presence of dry air (i.e., eo < eS(l/) and stronger winds.
Evapotranspiration is also enhanced by the roughness of the surface. This
representation of evapotranspiration as a diffusive flux is due to the
English scientist John Dalton ( 1802).
65
fluxes across the boundary and the net change in energy held within the
volume. The energy may change among its possible forms (radiant,
thermal, kinetic, and potential), but it must be conserved.
All matter has internal energy, known as heat Q [M L21"2], which is
due to the kinetic and potential energy associated with individual
molecules. Heat is an extensive property (it depends on the amount of
material) and is expressed in units of calories or joules. There are two
Iypes ofheal: sensible and latent heat.
Table 2.4. Some example of surface roughness for a few types of ground
surfaces and vegetation covers
•
Surface Zo (cm)
Ice 0.001
Dry lake bed 0.003
Calm, open sea O.QI
Closely mowed grass 0.1
Snow-covered farmland 0.2
Tilled bare soil 0.2-0.6
Thick grass (50 cm high) 9
Forest (on level ground) 70-120
•
66
temperature is represented in degrees kelvin, which is indicated as "K").
We can use the specific heat capacity to determine how the temperature of
a given mass of water will change if we add energy (heat the water). The
heat capacity of water at 20°C is approximately 4.2 103 J kg-I K-l or 1.0
x
I!.Q = 12000)
I!.T= =2.9K
mc" 1.0kgx4.2xIOJJkg' K-'
I
--------
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
----- ------- ----
___ I
67
Figure 2.11 A control volume for energy conservation. Solar energy (R'l)
entering the control volume must be balanced by fluxes of energy out of
the volume and by the time rate of change in energy stored (dQldt).
(2.10)
This tells us that we need to add about 2.5 million joules of energy to
evaporate 1 kilogram of water. Other types of latent heat energy include
Note that evaporation of snow or ice involves a change in phase from solid
to vapor. This may occur in either two separate steps, melting and then
evaporation, or in one step as sublimation, which is the direct phase
change from ice to water vapor. In either case latent heat is added to
support the sum of the two unique phase changes, as evidenced by the fact
that A.s = A.,. + A",.
The net solar radiation (Rn, i.e., the fraction of solar radiation that is
received at the surface and is not reflected back to the atmosphere) warms
the exposed surfaces inside the control volume. When water is present
some of this heat energy is absorbed by the water to support the phase
change from liquid to vapor (evapotranspiration). Evapotranspiration will
not typically absorb all the energy, and so the surface continues to warm.
As the surface becomes warmer than the air and the underlying soil, heat
will be conducted from the hot surface to the air (H) and from the surface
down into the soil (G). During this process, both heat and water vapor
have been added to the air inside the control volume. The moist, warm air
becomes less dense than the surrounding air and tends to rise up and out of
the control volume. Therefore, we see that the main energy outlets for the
sun's energy that hits the Earth's surface are conduction into the soil,
conversion of liquid water to water vapor, and heating of the overlying air.
The energy used to evaporate the water is stored in the water vapor and is
removed from the system as the water vapor is removed. Finally, the latent
heat (evaporation energy) is converted back to thermal energy when and
where the water recondenses. From this we see that the flux of water vapor
induced by evapotranspiration is associated with a flux of latent heat, £,.
The latent heat flux is related to the rate of evapotranspiration through the
latent heat of vaporz
i ation:
68
EI
el "" . (2.11)
p�A.\·
69
� 200Wm-�
et: O'J I ",8.0xlO-i IUS-I :O.7anday-.1
PW./"'J "'(000
1 kgm-32-
X .)X 1 kg- )
�
with
(2.16)
In the previous sections we have considered two methods that can be used
to calculate the evapotranspiration rates. The mass-transfer method
outlined in Section 2.4.2 uses Equations 2.8 and 2.9 to detennine et as a
function of wind speed (uo)' air vapor pressure (eo), the saturation vapor
pressure at the surface (esol' which is a function of surface temperature Ts),
and a number of parameters that depend on air temperature and surface
characteristics (Tables 2.3 and 2.4). The energy balance method outlined
in Section 2.4.3 uses Equations 2.14 to 2.16 to calculate et as a function of
the net radiation (R), heat conduction to the ground (G), wind speed (110),
surface temperature (Ts), the air temperature (To) at the height (zo), and the
set of parameters reported in Tables 2.3 and 2.4.
It can be convenient to combine the two methods to eliminate one of
these variables, preferably a quantity that is hard to measure. The Bowen
ratio method eliminates the dependence on wind speed. It is based on the
70
definition of Bowen ratio (B)-the ratio between sensible and latent heat
fluxes-and can be expressed (using Equations 2.8, 2.9, 2.15, and 2.16) as
B=!i=r T,-T"
E,
( e_-e"
l. (2.17)
et we obtain
R,,-G
1.'1 = . (2.18)
p�A.,.(l + B)
The Bowen ratio method uses Equations 2.17 and 2.18 to calculate the
evapotranspiration rate. Because Equation 2.17 is obtained using
Equalions 2.8 and 2.9, it docs not account for the effect of stomatal
regulation and can only be used in the case of evaporation from a lake, or
other wet surfaces. We see from Equation 2.17 that as a surface becomes
wanner (Ts increases) and drier (eSat decreases), the Bowen ratio tends to
increase. Consequently, sensible heat flux increases relative to latent heat
flux. The Bowen ratio ranges from approximately 0.1 in very humid
regions (8 0.1 over the oceans) to values greater than I in arid regions.
-
e/=
Sa [R" G J + KIIIIAc"" (T,,)
- -
e" J. (2.19)
p A,.(S" + r)
•. (S" +r)
where S" is the slope of the saturation water pressure curve (Figure 2.1); its
values are reported in Table 2.3. As in the case of the mass transfer method
presented in Section 2.4.2, Equations 2.18 and 2.19 do not account for the
effect of stomatal regulation. Thus, the Bowen ratio method and the
Penman method can be used to estimate evaporation but not transpiration.
The Penman Equation 2.19 can be modified to account for stomatal
regulation (known as the Penman-Montieth method), though this
introduces a number of additional parameters. Brutsaert (2005) provides a
71
good explanation of the Penman-Montieth method.
72
600
1\ R"
500
G f\.
---
H
400
£,
-
300
"
,
E
=:
� .
-
><
200
,�
"
u..
-
(/ \
/ ,
-
100
' ,
',t
, ,
,
, ,
,
,
,
\.\\.
,
--'
,
1'"
c
�
, ,
o
�'"
,�-----
-
-
.-
.
-
-100 ..
, . , . .
, • .
, •
, .
, ..
,
. . .
E E E E E E E E E
• • • . . •
• • � � • • � � •
. . . .
0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 .. 0 0
N '" N '" N '" N iri N
� � � � �
The rate of el that occurs under the prevailing solar inputs and
atmospheric properties, if the surface is fully wet, is commonly referred to
as potential evapotranspiration (PET). By definition PET does not depend
on surface properties and is not affected by stomatal regulation. For a
catchment water balance, we arc interested in the actual et, because that is
the rate at which water is removed. The concept of PET is useful as a tool
if we have some means to estimate actual et from knowledge of the
potential rate and the nature of the surface conditions. By definition, when
the surface is wet the ralio of ei to PET will be unity. Conversely, when
the surface is completely dry, the ratio will go to zero. The transition
73
between these two extremes is shown in Figure 2.13 for the simple case of
evaporation from a bare soil field. The actual et was detemlined from
detailed measurements of wind velocity and vapor pressure in the
atmosphere; PET was calculated from measured atmospheric properties.
1 ,------,
o
I:
2 0.75 -
1ii
.
.
�
a.
-
on
I:
.. o
0.5 -
�
'0 o
a.
� o
.. o
-
o
i
o
.-
0.25-
(;
a.
-
O� __
-- -
--._
, -- -- -- - - --
----
- ,
._ -- -- --�
o 0.25 0.5 0.75 1
DRY SATURA TED
Degree of soil saturation
et to PET decreases as the available soil
Figure 2.13 The ratio of actual
moisture decreases.
Data courtesy ofJohn D. Albertson.
74
2.5 Concluding Remarks
75
Subsequent chapters will deal with the terrestrial portion of the
hydrological cycle. To proceed to a physical treatment of water flows in
streams and beneath the Earth's surface, we will need to gain a basic
understanding of fluid mechanics. The physics of fluids wi II be the subject
of Chapters 3 and 4.
76
along gradients of decreasing water vapor pressure {Section 2.4.2}
Problem 1. Two tipping bucket rain gages are used to collect the
following rainfall data:
77
Cumulative precipitation (mm) Cumulative precipitation (mm)
Station #Z
Time
Station'1
12:00 noon 13 11
4:00 p.m. 19 16
6:00 p.m. 19 17
8:00 p.m. 19 17
10:00 p.m. 19 17
2:00a.m. 19 17
4:00a.m. 19 17
A. Calculate the mean daily rainfall intensity for each station (mm hr-1 ).
B. Calculate the maximum 2·hour rainfall intensity for each station (mm
hr-1 ).
C. Calculate the maximum 6·hour rainfall intensity for each station (mm
hr-1).
D. Using the arithmetic average method and knowing that the drainage
basin area is 176 mi2, calculate the total volume of rainfall (m3)
delivered to the basin during the event.
78
A. Calculate the cross-sectional area (m2) of a United States Class A
evaporation pan through which inflows and outflows of water can pass.
Also, calculate the total storage volume of the pan (m3).
B. Initially, the pan contains 10.0 U.S. gallons of water. Calculate the
depth of water in the pan (mm).
E. The pan is emptied and refilled with 10.0 gallons of water and left in an
open field for another 24 hours. During this period, rain fell for a 3-hour
period at a constant intensity of 2.5 mm hr-1; after 24 hours, the volume
of watcr in the pan was 11.50 gallons. Calculate thc average
evaporation rate (mm he l ) from the pan during this period
Problem 3. For Problem 2 part 0, calculate the flux of latent heat from the
water in the pan to the atmosphere (W m-2). Use a water density, Pw =
1000.0 kg m-3.
A. Write a complete energy balance equation (i.e., including all tenns) for
the catchment for the month of June.
79
C. Using a mean Bowen ratio ofD.2D for the catchment, calculate the mean
daily flux of latent heat to the atmosphere (W m-2) and the mean
evapotranspiration rate (mm day-I) from the catchment. Usc a water
density,pw= 1000.0 kg m-3.
80
3 The Basis for Analysis in Physical
Hydrology: Principles of Fluid Dynamics
3 . 1 Introduction
3 .3 Forces on Fluids
3 . 5 . 1 Fluid acceleration
3 .8 Concluding Remarks
3 . 1 0 Example Problems
3 . 1 1 Suggested Readings
3.1 Introduction
81
processes governing fluid motion is called fluid dynamics. Flow of water
in streams and subsurface aquifers, infiltration of precipitation into soils,
cvaporation, design of flood control measures, and transport of
groundwater contaminants are just some of the problems hydrologists
study that depend on a knowledge of fluid dynamics. In hydrology we are
obviously most concerned with understanding the behavior of water, but
the principles discussed in this chapter apply to a broad range of fluids.
The notions of force balance and Newton's second law (force equals
mass times acceleration), familiar from basic solid mechanics, can be
extended to fluids, including water. The forces that cause fluids to move
are due to gravity, pressure differences, and surface stresses. In this
chapter we analyze relatively simple cases such as flow through a pipe or
hose, but the same principles apply to flow in streams and through soil and
rocks. These applications are explored in subsequent chapters. The basic
principles developed in this chapter also apply to the flow of air, lava,
glaciers, and the Earth's mantle. While all of these can be described as
fluids, they differ dramatically in their viscosity. The viscosity of lava is
almost a billion times that of air. We will find that fluid viscosity plays an
important role in the nature of fluid flow. Flows in which viscous forccs
dictate the nature of flow are called laminar flows; flows in which viscous
forces are relatively unimportant are often turbulent. Most groundwater
flow is laminar, while most surface-water flow is turbulenl.
82
exerts a tangential stress on the water surface.
The above definition is valid for any fluid. The rate at which
defonnation occurs, however, is not independent of the fluid itself. The
property of a fluid that describes the resistance to motion under an applied
shear stress is tenned viscosity. The viscosity of honey is greater than the
viscosity of water, which in tum is greater than the viscosity of air. This
common sense notion of viscosity corresponds with the scientific notion.
To fonnulate a precise definition of viscosity we will consider a simple
conceptual experimcnt. Imagine a very large, thin sheet of matcrial (area =
A) floating on a very shallow layer of water (depth = d) lying above a
horizontal boltom surface. Suppose further that the sheet is pulled (with
force F) across the surface of the water. This experiment can be
conceptualized in more concrete lenns as attaching a rope to the edge of a
square sheet of Styrofoam 2 meters on a side, floating this on a layer of
water that is 5 mm deep at the bottom of a large, flat-bottomed pool, and
pulling slowly on the rope to cause the Styrofoam to move tangentially
(parallel to the water surface) across the surface of the water. We would
find thai the fluid adjacent to the bottom of the pool would "stick" to the
bottom and not move and that the fluid adjacent to the underside of the
floating sheet of Styrofoam would stick to that surface and, therefore,
move with the same velocity as the sheet itself, say up/ale (Figure 3. 1).
After a short time we would find that the velocity of the water between the
sheet and the pool bottom is proportional to the distance above the bottom,
as long as the product lip/ale x d is less than about 0.002 m2 S-l; the reason
for this limit is explored in Example Problem 3 ( Section 3. 10). For
example, at a distance of d/2 units above the boltom-halfway between the
bottom surface and the floating plate-the water velocity would equal
IIp/O/e12. We would also find that the applied force divided by the area of
the plate is proportional to the surface (plate) velocity divided by the water
depth or
F 11[>1<",
-=JI (3.1)
A d
The constant of proportionality, 11, is the viscosity of the fluid.
Equation 3 . 1 is a specific case of an important relationship for
understanding the flow of water, air, and many other fluids. More
generally, FIA is the tangential force per unit area (shear stress) acting on
the fluid surface and lip/auld is the slope of the curve (a straight line in this
case) relating fluid velocity to distance above the bottom boundary. A
curve relating velocity to distance above the bottom boundary is called a
83
velocity profile; the slope of such a curve is called the velocity gradient.
We can rewrite Equation 3.1 to relate the shear stress to the velocity
gradient:
d"
=
r J1 d:: ,
(3.2)
d,
z water depth
84
molecules are pictured as being "smeared" or "averaged" to eliminate
spaces between atomic particles. Thus, we will use the terminology "at a
point" so that we havc the mathematical tools that are necessary, but it
should be borne in mind that the "point" actually represents average
conditions over a small volume. For our purposes explanation in physical
hydrology-this assumption is valid. (There are situations involving
rarefied gasses important in other branches of science for which the
continuum assumption may not bc acceptablc.)
Oncc the continuum assumption has bcen made, we can define the
density of fluid at a point. Density is simply mass per unit volume. If
density varies spatially, the ratio of mass to volume will vary with the
sample size. Hence, the definition of density at a point is the limit of the
ratio of mass to volume, as the volume under consideration shrinks to an
infinitesimal:
,. aM
p= lin . (3.3)
<1.1'-0 AV
mass 11M [M]; the "11" before V and Mjust reminds us that we are
considcring small volumes and masses. Of course, the continuum
assumption means that density " at a point" is really the average density of
a volume that is small compared with the macroscopic scale of the fluid
motions being studied but large compared to intennolecular distances. If
the density of a fluid does not vary spatially, the fluid is said to be
homogeneous. A fluid property related to density is unit weight. The unit
weight (y) of fluid is its weight per unit volume and is equal to the product
of density and the acceleration of gravity, that is y pg.
=
85
area that tends to compress a fluid. Values of density given in Table 3.1
are for a pressure of I atmosphere. The deviations in density caused by
changes in pressure are often ncgligible. In these cases, the fluid is said 10
be incompressible. However, there are cases of importance in
environmental sciences where the compressibility of water must be
considered and, thus, the variation of density with pressure taken into
account. These cases include the variation of water density with depth in
the oceans and with the variation of confining pressure in deep aquifers.
86
we consider in fluid dynamics-normal stresses and tangential stresses.
The latter, as we have already seen, are termed shear stresses. The inward
directed (compressive) normal stress, when applied to a fluid medium, is
referred to as pressure.
A special case of fluid motion is that ofa fluid at rest, i.e., the case of no
motion. If a fluid is not moving, there are no shear stresses present (by
definition) and only pressure need be considered when analyzing surface
forces. The special case of no motion has important applications and the
study involving such problems is referred to as fluid statics.
The variation of pressure with depth in a fluid at rest is one of the
fundamental relationships in fluid dynamics. The equation describing this
variation in pressure is known as the hyd rostatic equation. We will derive
this equation by considering a thin slice of fluid at rest and considering the
forces acting on it (Figure 3.2). The slice has a surface area A and a
thickness tlz. The horizontal pressure forces acting on the sides on the slice
must be equal and opposite so that they sum to zero. Ifnot, there would be
a net pressure force causing the slice of fluid to move sideways, at odds
with our assumption that the fluid is not moving. The upward force on this
slice, pA, is due to the pressurep on the bottom face of the slice. There are
two downward forces, the wcight of the slice itself and the force due to
pressure on the top face of the slice. The weight of the slice is given by the
product of density p times gravitational acceleration g times the volume of
the slice V (= Atlz). The pressure on the top face of the slice will differ
from the pressure on the bottom by some small amount I1p, so that the
downward-directed pressure force on the top face is (p + I1p)A. Because
the fluid is not moving, the upward and downward forces acting on the
slice must balance each other, giving:
which we can solve for the change in pressure !!.p from the top to the
bottom of the slice:
or
87
/;. P = _
pg
/;.=
(p+/;.p)A
pA pg6zA
Figure 3.2 A thin slice within a static fluid. Because the fluid is at rest, the
upward and downward forces acting on the slice must balance.
dp
-� -pg. (3.4)
do
Equation 3.4 is the hydrostatic equation. It indicates that the rate of
decrease of pressure with distance as we proceed upward (relative to
gravity) is pg, the unit weight of the fluid. The pressure is caused by the
weight of the overlying fluid. In many instances involving water we can
consider density to be constant and we can then integrate Equation 3.4 to
give pressure as a function of depth. Suppose the slice in Figure 3.2 is at a
level d units below the surface oflhe fluid and thaI z = 0 at the surface. We
will integrate Equation 3.4 from z = -d to z = 0, i.e., from the slice to the
surface. Rewriting Equation 3.4 in differential fonn,
dp = -pgd= (3.5)
88
p, -p � - pg[O - (-d)]
0'
p - p. = pgd. (3.6)
p= pgd. (3.7)
Thus, the pressure in a constant density, stalic fluid is the product of the
unit weight and the depth. Pressure increases linearly with depth.
As an example, let us calculate the pressure on the bottom of the 1-
meter high pond wall shown in Figure 3.3a. The unit weight of water is 9.8
kN m-3 and the depth is 1 meter. Therefore, the pressure at the bottom of
the pond, and hence at the base of the wall, is 9.8 kN m-2. The pressure on
the wall varies linearly with depth such that the pressure at 0.5 m is 4.9 kN
m-2 and the (gage) pressure at the surface is 0 kN m-2 (Figure 3.3b) . The
lotal force on the wall is the integral of pressure over area. Ifwe consider a
I -meter width ofthc wall (perpcndicular to the plane of the page in Figurc
3.3), then the force on the wall is detennined by:
(a) (b)
1 m
Pressure at bottom
= 9.8 kN m-2
89
Figure 3.3 Water pressure on a vertical wall (a) increases linearly with
depth at a rate of9.8 kN m-2 per meter (b).
kN kN
Force per meter of width =
J'I -9.8
m
3 :d: = 4.9
In
.
where d is the height of the mud column above the formation. For a
drilling mud with unit weight (Pm,,,' g) = 10.5 kN m- ,
3
90
Dril li ng mud fills the
pipe stem
Figure 3.4 Drilling mud, with a density typically 10 to 20% greater than
that of water, is used to stabilize wells by providing a hydrostatic pressure
at depth to counterbalance the formation fluid pressure.
So the required height of the drilling mud in the well is about 1 700 m. The
depth of the formation is 2000 m so there the balancing pressure of the
drilling mud is more than adequate to prevent a blowout.
91
Bernoulli equation are also considered.
92
derivatives. We write the local rate of change in temperature as the partial
derivativc of tcmperature with time, oTlot. This is just a notational
convenience for expressing the change of temperature with time at a fixcd
location (e.g., Washington). Likewise, aT/Oy is the partial derivative of
temperature with distance in the y direction or the rate of change of
temperature with distance at a fixed time (e.g., 8 a.m.). The total rate of
change oftcmperature with time is written as dTldt and is equal to aTlot +
v (oT/Oy) where v is the velocity in the y direction.
Acceleration is treated in an entirely analogous manner to temperature
in the above example. Total acceleration is the sum of local acceleration
and convective acceleration, and acceleration in the x-direction can be
written:
rill au au
= + 11 .
(3.8)
dl al ax
- - -
93
solvable, but one that nevertheless is general enough to describe practical
flow situations. An aim of hydrology is to determine what simplifying
assumptions are acceptable for giving useful predictions of the behavior of
real hydrological systems.
We will eonsider the problem of the flow of water through a garden
hose as the starting point for developing an equation for fluid motion. The
results, as we will see, apply to a far wider range of problems than just a
garden hose. The hose is connected to a house spigot with sufficient
pressure to cause an ample stream to leave the open end, perhaps 1 0 m
away. Clearly the pressure at the spigot will affect the flow in the hose. We
must also consider what other factors might be influencing the water in the
hose. As in the hydrostatic case (i.e., no motion), we approach the problem
by making an assessment of the body and surface forces acting on the
fluid.
We will consider an arbitrary short section of the hose of length ds
oriented in the s (an arbitrary) direction to begin constructing the force
balance for the water movement through this small volume (Figure 3.5).
Water is flowing into this small piece from the house, or upstream, end
and out of it at the downstream end.
Pressure causes a surface force to act on the small volume. There is
both an upstream pressure, PI' and a downstream pressure, p2' Thus there
is a force i n the downstream direction, PIA, and a force in the upstream
direction, py4. . Another force that acts on the volume of water in the short
section of hose is the weight of the water itself, that is, its mass multiplied
by the acceleration of gravity:
F;r (pdV)g.
: (3.9)
94
z
x
u.. 3.5 A section of a garden hose used to analyze the pressure and
gravitational force acting on a small volume of water.
The forces from pressure and the weight of the water are the only two
forces we will include in our simple model. It is easy to imagine a more
complex situation where other forces would be important. If, for example,
the fluid were beryllium, a relatively magnetic fluid, instead of water, then
it might be appropriate to consider magnetic forces. A morc obvious force
we have nOI listed is the frictional force between the water and the hose or
within the flowing water. Frictional forces are tangemial surface forces
due to the viscosity of the water and the fact that fluids in contact with a
surface "stick" to that surface. Although friction is certainly a real surface
force, we wish to keep our model simple, and so will neglect frictional
forces for the moment. We must keep i n mind, however, that this
assumption of negligible friction is not correct. We recognize from the
start that the final result will be useful only in cases when this assumption
does not cause too much error. In many cases in hydrology, the equation
95
resulting from this simple, frictionless case has to be modified, as we will
see.
To develop our flow model, knowledge of physical laws and
observations of the flow itself arc used. According to Newton ' s second
law, the product of the mass and acceleration of the fluid within the hose
must be equal to the sum of the forces acting on the fluid volume:
(3.10)
where Fg is the body force due to gravity and Fp is the surface force due to
pressure, the net force resulting from the sum of the force due to upstream
pressure, PIA, acting in the +s-direction (positive force) and the force due
to downstream pressure, P;0, acting in the -s-direction (negative force).
(3. 1 1 )
Over the small section of hose that we are considering, the pressure at the
downstream end,P2' differs from the upstream pressure, P ' by a small
I
amount dp,
dp
P2 = PI +-ds.
ds
The tenn dplds is the pressure gradient in the s direction. Substituting this
into Equation 3. 1 1 ,
or
dp
Fp = -- dsA . (3.12)
d,
The body force Fg acts in the vertical direction. We are interested only in
the component of this force acting along the axis of the hose, the s
direction. Fgs can be computed from a simple trigonometric relationship
(see Figure 3.5):
�, = F, sine. (3.13)
where e is the angle the hose is inclined from the horizontal. We can
further simplify this expression if we note that
96
do
sin O = --. (3.14)
d,
do
F.
" =-g-pAds.
d,
(3.15)
We now have all of the elements of the force balance in useful terms.
The mass of the fluid in the control volume is the density times the
volume. The forces are as indicated in Equations 3. 12 and 3. 1 5.
Assembling these in the fonn of Newton's second law (F= ma) gives:
de dp
(pAds)a = -g-=-pAds - - Ads,
d... d...
d"
a = u-, (3. 16)
d,
97
slope and velocity? To answer these questions, it is first necessary to solve
Equation 3.1 7. Our solution will be simplified if we recall from the rules
for differentiation that
(3.18)
�(�+:+...E....l =o.
ds 2g pg
(3.19)
1/2 P
- +: + - '" const:lnt. (3.20)
2g pg
112 L2 T2
Velocity head: -dlln -- = L
•
2g T2 L
98
fluid (pgV) gives the equivalent equation in units of energy:
111112
p V + mgz + 2 = eonstant.
We see that pV is the "flow work" or the work due to pressure, mgz is the
potential energy, and nlu2/2 is the kinetic encrgy. Thus, the Bcrnoulli
equation can also be thought of as a statement of the conservation of
energy.
At this point it is desirable to recall all of the assumptions needed to
derive the Bernoulli equation, namely:
a. no friction,
b. incompressible fluid,
c. homogeneous fluid,
d. flow steady with time.
In addition, the solution is valid only along the path of integration, the s
axis, a line everywhere parallel to the flow field that is referred to as a
streamline.
u' P
-+=+ constant H.
2g
= =
pg
99
Because the total head is constant, the sum of the velocity, elevation, and
pressure heads will be the same at every point on a streamline. Using
subscripts to identify points I and 2 on the streamline, the Bernoulli
equation can therefore be written in the useful fonn:
,
IIi PI IIi Pl
-+- =-+- 1 +-
,
1+- _ _
(3.21)
2g pg 2g pg
If a spot of dye were placed carefully on the surface of the tank, its
speed of motion would be seen to be very small with respect to the speed
at which the water exits the tank because the surface area of the tank is so
much larger than the cross-sectional area of the spigot at the outlet.
Because II I is so small relative to 112 (1/1 « 112), we can assume that 111 is
equal 10 zero in our analysis. The pressure on the surface would be the
local atmospheric pressure, as would be the pressure at the outlet of the
spigot, point 2. The small difference in atmospheric pressure due to the
elevation difference between I and 2 can be ignored. With these
considerations, we can now write,
_I
_ + Put",.,, = -
2
"2 + Z, + !P�m�"��"
pg 2g - pg
or
100
===
9 1"1
=� Inflow keeps water
II level constant
PI = ['2
-
'"
-
-
111 ", 0 1) :::r =
=
PaUfIOS
0 (gage)
-,
z �"
172
,
"
-
Datum (z 0)
=
-
Figure 3.6 Tank with steady flow of water. The Bernoulli equation
provides a relationship between the depth of water in the tank and the
velocity of water exiting the tank at point 2.
If we let d = zl-Z2 (the depth of water betwcen points I and 2), then,
(3.22)
Thus, the velocity of the water leaving the tank depends on the depth of
the waler in the tank.
If a plug were placed in the spigot, shutting off the flow, we could use
this same approach to compute the pressure at point 2. In this case, both II I
and //2 are zero but P2 is no longer equal to atmospheric pressure:
- -
-1 - + p,
- -2
pg
0'
P2 = pgd. (3.23)
This equation, for thes tatic situation, is identical to Equation 3.7, the
hydrostatic equation. Thus, we have shown that the hydrostatic equation is
a special case of the Bernoulli equation.
Let us return to our general fonn of the Bernoulli equation (Equation
3 .20),
101
u2
+=+ L = constant,
2g pg
-
and considcr the relationship betwcen pressure and velocity for thc garden
hose. If the velocity along a horizontal segment of a streamline (i.e., one
for which z did not change) were to increase, say because of a constriction
in the hose, we would conclude that the pressure would decrease. That is,
pressure and velocity are inversely related, assuming no change in
elevation, z.
Often we are more interested in the average or mean velocity of flow
through a hose or pipe than we are in the velocity along any particular
streamline. The mean velocity U at any cross section is the discharge Q
divided by the cross-sectional area A, that is, U = QIA. The mean velocity
also can be thought of as the avcrage value of the velocities at each point
in the cross-section. For the frictionless flow described by the Bernoulli
equation, the velocity at any point in a cross section is equal to the mean
velocity, so u = U. This is not true for flows in which friction is important.
One of the fundamental constraints on flow through pipes or channels is
conservation of mass: rate of inflow minus rate of outflow equals rate of
change of storagc (Eq. 1.1). In the case of steady flow in a full hose, there
are no changes in the amount of water in any segment of the hose at any
time (no change in storage) so inflow rate must equal outflow rate.
Because we are taking density to be constant, the volumetric inflow rate
must equal the volumetric outflow rate. The conservation of mass (or
volume, in this case) equation is oftcn referred to as the continuity
equation, simply written,
Q = UA = constant. (3.24)
The continuity equation and the Bernoulli equation are two of the
fundamental relationships of fluid mechanics.
As an example of the way conservation of mass can be used to provide
infonnation about flow, consider a section of hose where the diameter
changes from D1 to D2 , where D2 = 2D1 (Figurc 3.7). Becausc we have
assumed the flow to be steady, the same quantity of water flowing through
the narrow scction must be flowing through the wide section:
(3.25)
and thus,
(3.26)
102
where the cross-sectional area of the hose is A = rrD 2/4 and D is the hose
diameter. Substituting the expression for cross-sectional area into Equation
3 .26 gives,
lfDI2 = U, lfDi
UI 4 - 4
"\
•
•
•
VI
� �
•
•
•
• °2
•
•
"l-
•
/
I
2
A 1 - ( rrO I )/4
-
2
A2=(rr02)/4
Figure 3.7 An expansion joint in a hose. Discharge is the same at points I
and 2 (upstream and downstream ends of expansion joint) for steady flow
through the hose. As a result, if cross-sectional area increases, velocity
must decrease.
01
Thus, the velocity in the narrow section of the hose is four times the
velocity in the wider section.
If the hose is level (zl = z2), the Bernoulli equation is given by:
103
Ur + PI _ ULl + P2
::c
2g pg 2g pg
where the subscript i refers to position along the hose ( I , 2, and 3 in Figure
3.8).
104
•
•
I
hi
•
Figure 3.8 The height ofwaler spouting from holes punched in a hose
decreases along the length of the hose because offrictional losses of
energy (head loss) as the flow travels downstream in the hose.
u'
P +=+ - = H = constant.
PI!. 21!.
In fact, we observe a loss in total head along the hose. hL• which if added
to the other tenns in our equation would give a total head equal to the
105
initial sum H. That is, the Bernoulli equation must be modified to account
for the energy loss to friction:
U'-
J!.... + = + - + "t constant, (3.29)
pg 2g
=
where hL is called the head loss. Head loss is simply an empirical way of
dealing with the fluid friction that is dissipating energy (converting
mechanical energy to thennal energy) inside the hose over a specified
length of hose. For a fluid with a non·zero viscosity, like water, friction
causes fluid adjacent to a solid surface to "stick" to that surface. This
means the velocity of the water in contact with (stuck to) the wall of the
hose mllst be zero because the hose is not moving. As a result, there will
always be a velocity gradient away from the wall within a frictional flow
(Figure 3.9). Energy dissipation is a result of this velocity gradienl.
/ /
'.
•
•
•
•
•
,
,
,
,
,
,
,
,
,
I----�/
/'
,
,
,
\
Figure 3.9 Water "sticks" to the walls of the hose and vIscosity causes a
frictional energy (or head) loss. Velocity is zero at the walls of the hose
and a maximum at the centerline.
For flow through a circular pipe or hose, head loss between the two ends
of the hose depends on the viscosity of the fluid, the velocity oflhe flow,
hose diameter and length. Because head loss also is a function of the
roughness of the wall material itself, we must include a friction factor in
its definition:
L u�
"/. - f --, (3.30)
D 2g
106
wheref= friction factor [dimensionless], L= length [L], and D = diameter
[L).
lf we consider the special case of a horizontal hose of constant diameter
(and therefore constant cross-sectional area), velocity and elevation don't
change along the hose. We can write the Bernoulli equation including head
loss as
where ilL is the head loss between cross-sections I and 2 of the hose. If
velocity and elevation don't change along the hose, then VI V2, ZI Z2 '
= =
and thcrefore hI. (Pl-P2)/pg. If we cxpress head loss as head loss per unit
=
length, then
!!L = , PI - Pl , tJ.p , dp
(3.32)
_
_ _
_
_
=
__
_
I. pg I. pg L pg d'C
where Lis the length of the hose between cross-sections I and 2. Ifwe
combine Equations 3.30 and 3.32, we can obtain the following expressions
for velocity:
(3.33)
Note that dpldx is negative for flow in the positive x-direction, In other
words, there is a pressure drop in the direction of flow. Equation 3.33
suggests that if we can measure the pressure drop between two points
along a hose or pipe a distance Lapan and we know the friction factor, we
can calculate the velocity or discharge of flow through the pipe.
The problem of calculating or predicting the velocity at which water or
other fluids would flow through a given pipe, channel, or other conduit is a
problem that has great practical importance in the design and management
of water supply and sewer systems, waterways, and industrial pipelines.
Calculating velocity using Equation 3.33 requires knowledge of the correct
friction factor. Friction factor is a quantity that must be determincd
experimentally. Because of its importance, many measurements of friction
factorfhave been made under controlled laboratory conditions. A number
of these measurements of friction factor are ploned in Figure 3.10 against
the term pUD/�. This dimensionless term, the Reynolds number R, is
discussed in the next section. When plotted in this manner, the data for all
pipe diameters, discharges, and fluids describe a well-defined relationship
'07
for pipes that are smooth inside. We refer to this as a friction factor
diagram.
2X l0 ',-
---,C- __________________________________ --,
k,lD
----
----- --- ------- 4.0><10"
2.0><10" ilr
-- ----------- !!
------------·4.0Ml0..
�.
,::- .:::- -- -
,;: --- -
g�
Turbulent �
'0
. � '0'
- --
- - - - - -
,
- _
1.0><'0� �
10' 1.0><10"
5xl 0-3 +-
_____ ....,
______-r ,-
_______ _____ ...,
'-'- �
__
108
conducted by Sir Osborne Reynolds (1843-1912). Reynolds was an
English civil engineer noted for both his theoretical and applied studies in
fluid mechanics. Rcynolds's expcriment used a simple apparatus to
visualize a flow field (Figure 3.11). In this apparatus, dye is injected along
the centerline of a pipe in which fluid is flowing. The dye serves to show
the nature of the flow as the fluid moves along the tube. At low velocities
the dye rcmains approximately on the centerline, traveling downstream
with, and at the velocity of, the flow. Because of molecular diffusion there
will be a small flux of dye away from the centerline, but this will be so
small that no departure from the line will be observed with the naked eye.
Reynolds characterized this flow situation by describing the fluid as
moving in dislinellamilJae (Latin for layers), and today we refer to such
flow as laminar flow.
Now consider what happcns as the velocity of fluid flow in thc tube is
increased. At some value of increased velocity, the dye is observed to
become mixed within small, randomly located spots within the flow; with
higher velocities the dye rapidly becomes completely mixed with the
surrounding fluid in a short distance of flow. Rcynolds referred to this
flow as sinuous or disturbed flow. Today we refer to flows exhibiting such
violent mixing as turbulent flows. In this case the mixing takes place
much more rapidly than realized by molecular diffusion, and in fact we
call this mixing tllrblllellf diffllsion, We can conceptualize the turbulent
mixing as small parcels of the fluid "jumping" away from a streamline to a
different portion of the flow field. Most flows we observe in nature, for
example, clouds, rivers, ocean waves, are turbulent flows. In this regard,
we note that "laminar" and "turbulent" refer to properties of the flow, and
not properties of the fluid, For example, the water in the garden hose could
be in either laminar or turbulent flow without changing its propenies, that
is, density or viscosity.
109
'.
•
•
•
�--.:;-
•
1
•
---------
•
•
,
•
·
(a) .'
(b) .'
Figure 3.11 Reynolds 's experiment. In laminar flow (a), the dye remains a
thin line, moving with the fluid. The dye trace becomes convoluted in
turbulent flow (b).
To explain the last statement, we return to Reynolds's experiment. By
carefully changing the properties orthe fluids in his apparatus, as well as
the tube diameter and flow velocity, Reynolds empirically defined a
dimensionless number R which describes the flow properties:
=ULp
R (3.34)
I'
110
diameter hose, the Reynolds number would be calculated as follows:
1
u= 2 m 5- (mean velocity);
(2XO.03XiO')
53 000
=
R �
• •
(1.139xI0-')
which indicates that the flow is turbulent. The flow could be made laminar
by somehow reducing R to some value less than 2000. We could
accomplish this by:
III
have a tendency to deviate from the centerline, and the flow could become
turbulent. We can therefore take the ratio of the inertial to viscous force as
an index of whcn flow would be expcctcd to be laminar or turbulcnt. If the
viscous forces» inertial forces, the tlow is laminar, and if the inertial
forces» viscous forces, the flow is turbulent. We know that the inertial
force is proportional to mass x acceleration, which could be expressed as
(p)(L2)(velocity2). Similarly, the viscous force is proportional to the shear
stress x area, which could be expressed as (1l)(L)(velocity). Thus,
R (3.35)
viscous force JlLV Jl
The Rcynolds number is equal to the ratio of the inertial to viscous forces
within the fluid.
Let us return to our example of water flowing through the garden hose . If
the hose is narrow enough, the flow slow enough, or the viscosity high
enough, the flow could be laminar. In this case, the experimental
relationship between friction factor and Reynolds number indicated in
Figure 3.10 is linear and is given by:
1 =64,. 64JI.
(3.36)
R pUD
de 32p
D2
U=_dp . (3.37)
dr 32J1
The velocity given by Equation 3.37 is the mean velocity for laminar
flow through the pipe; this velocity times the pipe cross-sectional area
gives the discharge, Q UA. Steady laminar flow through a uniform pipe
=
is referred to as Poiseuille flow and Equation 3.37 for the mean velocity is
sometimes referred to as Poiscuillc's law. As depicted in Figure 3.9, the
actual velocity at any point in the pipe varies with distance between the
wall and the center of the pipe. !fwe integrated this velocity profile over
112
the pipe cross section, wc would obtain thc discharge Q; the mean velocity
Uis this discharge divided by pipe cross-sectional area A.
In deriving Poiseuilte's law from the Bernoulli equation we assumed
that the elevation along the hose or pipe is constant, that is, ZI Z2' We can
=:
generalize the equation by relaxing this assumption and allowing the hose
10 be inclined at an angle as in Figure 3.5. In this casc a term relatcd to thc
clevation difference is added to Poiseuille's law giving:
( dP dZ D' )--
U=- -+pg-
ds ds 32p
or
(.3.38)
If the flow velocity, pipe (or hose) diameter, and fluid viscosity combine to
produce a Reynolds number in excess of 4000, as is often the case, then
the flow will be turbulent and Poiseuilte's law does not apply. Instead, we
must return to Equation 3.33 and use Figure 3.10 to detennine the friction
factor. For values of Reynolds number bctween 4000 and 100,000, thc
friction factor for smooth pipes can be approximated by the expression:
f=0.316R-'I� (3.39)
113
walls bounding thc flow in addition to thc othcr factors. Wc will scc in thc
next chapter on channel flow that characterization of channel roughness is
an imponant but difficult pan of cstimating flow velocity in streams and
other turbulent surface flows.
Velocity 1------<--1
gradient
implies shear
stress or skin-
friction drag
Flow separation
at small scales
implies form drag
Figure 3.12 Flow through a rough-walled hose. The roughness length, kf '
is a measure of the size of the irregularities or unevenness of the surface.
The material in this chapter was meant to introduce you to the basics of
fluid mechanics. We should emphasize again that our objective is to apply
these fundamental principles of fluid flow to hydrological problems. It
may not yet be clear to you how a description of pressure distribution, flow
and fluid drag in a garden hose is relevant to our objective. The
applications to the "real world" are discussed in the succeeding chapters
and we will see that the garden hose example will assist us in problems
ranging from the relationship between river velocity and bed roughness to
the interpretation of water levels in wells.
"Real world" applications of the principles of fluid mechanics are not
restriclCd to the flow of water. The winds that blow over the surface of the
Eanh are flows of air in response to differences in atmospheric pressure or
other forces. rlows of air near the Earth's surface are characterized by
high Reynolds numbers because air has a relatively low viscosity and the
flows arc always "deep." In contrast, lava has a viscosity more than a
thousand times that of water (the value depends strongly on temperature)
114
while its density is only 2 to 3 times larger. As a result, flows of lava are
low Reynolds number, laminar flows that can be described by a fonn of
Poiseuille's equation. One of the many things that makes water interesting
is that there are commonly occurring examples of both laminar and
turbulent flow.
• The density of a fluid is the mass per unit volume at each point in the
fluid. A homogeneous fluid is one in which density is constant
throughout the fluid. The density ofwater is approximately 1000 kg m- 3.
The unit weight of a fluid is its weight per unit volume, pg [N m-3].
{Section 3.2}
• Two classes of forces act on fluids: body forces and surface forces. Body
forces, such as gravity, act unifonnly on each fluid clement. Surface
forces, such as pressure and friction, act on the surfaces of fluid
elements. In fluid mechanics we typically represent surface forces as
forces per unit area, or stresses. There are two types of stresses: nonnal
stresses and tangential stresses tenned shear stresses. The inward
directed (compressive) nonnal stress, when applied to a fluid medium, is
referred to as pressure. {Section 3.3}
• In a fluid at rest, pressure increases with depth at a rate of pg, the unit
weight of the fluid. This is expressed by the hydrostatic equation, dpldz =
115
below the surface, p = pgd. The pressure given by this relationship is
gage pressure, the pressure relative to atmospheric pressure. To obtain
absolute pressure, the atmospheric pressure acting on the fluid surface
must be added to p. {Section 3.4}
• The Bernoulli equation states that, for a frictionless flow, the sum of the
velocity head 112 /(2g) [L], elevation head z [L], and pressure head p/(pg)
[L] along a streamline is a constant tenned total head, H [L]: [1l2j(2g)] + z
+ [P/(pg)] = H. Thc assumptions made in deriving the Bernoulli equation
are I) no friction; 2) incompressible fluid; 3) homogeneous fluid; and 4)
steady flow. For frictionless flow, the velocity II on any streamline in a
cross section is equal to the mean velocity U, so that U can be substituted
for II in the Bernoulli equation to obtain an equation in temlS of mean
velocity. {Section 3.5.2}
flow through a horizontal pipe, head loss is related to the pressure drop
along the pipe: ilL = (PI - P2)/(pg). {Section 3.6}
• Head loss for flow through a pipe or tube is related to mean velocity V,
pipe diameter D, and pipe length L by the empirical equation ilL =
116
are determined experimentally and are typically represented in friction
factor diagrams offas a function of pipe Reynolds number R pUDfl1=
• The friction factor for laminar pipe flow is given by f= 64fR. Combining
this with the head loss equation gives an equation for the mean velocity
for laminar flow in a pipe (Poiseuille's law): U -{dpidx)([j1f32J.1).
=
{Section 3.7.1}
• For turbulent pipe flow, friction factor must be obtained from the friction
factor diagram or determined experimentally. The friction factor for
turbulent flow increases with increa3ing roughness of the pipe walls.
{Section 3.7.2}
D. What depth (m) of mercury, with a unit weight of 133 kN m-3 , would
be required to produce a pressure of300 kPa?
117
Problem 3. Observations show that flow in a circular pipe of diameter D
remains laminar up to a Reynolds number Rpipe pUD/� 2000 (Figure
= =
3.10). What about flows in other flow in channels or pipes with different
geometries? Consider the example of flow between two flat plates shown
in Figure 3.1. In this case, the appropriate length scale for the Reynolds
number is L = d. I t also makes sense to use plate speed rather than mean
flow velocity speed as the characteristic velocity in R, giving Rp1me =
Pllp/ale d/�. Changes in length and velocity scales can alter the upper limit
for laminar flow, but for the case of flow between two parallel platcs, flow
is again laminar up to Rp1ate = 2000. (A parameter called the hydraulic
radius [Chapter 4.5] can be used to find values of R corresponding \0 the
laminar-turbulent transition for different flow cross-sections. This is
explored further in an example problem in Chapter 4.)
Problem 5. A tank like the one pictured in Figure 3.6 is filled to a constant
level of 0.70 m. The center of the outflow opening near the bottom is 0.10
m above the bottom of the tank. What is the velocity (m S-I) of flow
exiting from the outflow opening?
118
the density of the water is 1000.0 kg m-3.
B. What is the friction factor and the head loss per unit length for this now
(both are dimensionless)?
C. What is the change in pressure (Pa) over a 10-m length of the hose?
Problem 8. Lava, with a density of 2700 kg m-3 and viscosity of 1.0 x 103
Pa-s, nows through a conduit that is circular in cross section. The diameter
of the conduit is 1.0 m. Flow of lava through the conduit is driven by a
pressure gradient of -2.0 kPa m-I. Assuming the now is laminar, what is
the discharge of lava (m) S-I)? Is the assumption of laminar flow valid?
Fox, R. W., A.T. McDonald, and P.J. Pritchard. 2011. II/troductiol/ to Fillid
Mechanics, 8th ed. New York: John Wiley & Sons.
Middleton, G.V., and P.R. Wilcock. 1994. Mechanics in the Earth and
Environmental Sciences. Cambridge: Cambridge University Press.
Chapters 9 and II, pp. 296--336, 365-394.
119
4 Open Channel Hydraulics
4.1 Introduction
4.2 Specific Energy
4.2.1 Flow over a vertical step
4.2.2 Flow criticality
4.3 Discharge Measurements Using Control Structures
4.4 The Effect of Bed Roughness
4.5 Channel Flow Equations
4.5.1 Velocity distribution in open channels
4.6 Measuring Flow in Natural Channels
4.7 Concluding Remarks
4.8 Key Points
4.9 Example Problems
4.10 Suggested Readings
4.1 Introduction
120
when a train car containing a hazardous chemical derailed and spilled into
the river upstream of a city in California. It was imperative in this case to
know when the city water intakc had to be shut down to avoid drawing the
contaminant into the city's water-supply system. It turns out that, for river
channels, there is generally a useful relationship between mean velocity of
flow in a channel and the discharge in the channel (Figure 4.1).
The relationship between water depth and discharge is one of very
broad importance in hydrology. We often need to know how the depth of
water in a channel is related to thc quantity of watcr the channel is
carrying. Unlike the pipes and hoses that we considered in the last chapter,
channels do not have "lids." The water depth varies along with the other
hydraulic quantities. In the following discussion of hydraulics in stream
channels, the rclationship between watcr dcpth and discharge will be seen
to be important for flow mcasurement. Our primary goal in this chapter is
to explore the physical basis for relationships among the hydraulic
variables that we can measure in channels. In Chapter 5 these principles
will be applied to stream hydrology.
10 ,-------�--__,
-
�
,
.,
E
�
:;, 0
>-
""
0 0
0
Qi
>
c:
co
.,
:0
0.1 �--�����o�--�------�----------�
0.1 1 10 100
Discharge, Q (m' S-')
Figure 4.1 Velocity and discharge in rivers usually are related by a power
121
function, which plots as a straight line on logarithmic axes. The equation
of the straight line shown in the graph is U= 0.28(11-62.
Dala from Miller el a1. (1971).
To begin, consider how velocity and water depth are related in open
channel flow. An open channel flow differs from flow in a pipe in that the
channel flow is only partially enclosed by a solid boundary. The upper
"boundary" of an open channel flow, between the water and thc
atmosphere, is called a free surface. To keep thc analysis simplc we will
consider the flow to be steady and frictionless so that the Bernoulli
equation can be used.
Our first step will be to adapt this equation to a foml particularly suited
for application to channel flow. First consider flow in a short section of
channel (Figure 4.2). In a very short channel section, or reach, frictional
losses often c(ln be disregarded, (IS can any chrmge in Ihe elevation oflhe
bed. The Bernoulli equation for a streamline in this short segment of
channel is:
122
II' 1
+_
1 _+: = H (4.1)
2g pg
_
1/2 P
-+-+(=.+"-d)� fI,
2g pg
f
", water depth d
Streamline
FIOW --...
I
I
Zb' elevation z
t� atum (z = 0)
j of bed
Figure 4.2 Horizontal flow in a short channel section. The flow can be
considered uniform and frictionless over this section. The total energy per
unit weight of the water along the streamline is lJ2J2g + Ii + Zb = E + Zh,
where E is the specific energy.
where zb is the elevation of the stream bed above datum [l], It is total
water depth [L], d is the depth of the streamline below the free surface [L],
and H is the total energy per unit weight or total head [ll The "datum" is
just a convcnient rcference elevation; often mean sca level is used. As
discussed in thc derivation of the Bernoulli equation in Chapter 3, H is a
constant. The tenn zb + II d is the elevation head. If the flow is essentially
-
123
equal to Zb + h, where h is the total water depth and zb is the elevation of
the stream bed above datum. This would be tme for any streamline
because it is independent of d (e.g.. the gage pressure at the water surface
Cd = 0) is zero and the total elevation head is zb + h).
For frictionless flow the velocity is assumed constant over the depth. In
this case, the velocity on each streamline is the same and is equal to the
channel mean velocity V (L .1], so that Equation 4.1 can be written:
u'
-+,,+=� = H. (4.2)
2g
The first two terms of this equation, V2/2g + II, can be interpreted as the
energy per unit weight of the flowing water relative to the stream bottom.
This combined term is defined as the s pecific energy, E [L]:
u'
£=-+11. (0)
2g
From Equation 4.4 we see that irthe elevation of the channel bottom
remains constalll, then because H is constant, E must also be constant.
The continuity equation for the steady, uniform flow of a constant
density fluid states that the discharge Q [L3•1] remains constant so that
VI A I = V2A2, where A is the cross-sectional area of flow [L 2]. To simplify
the analysis even further, leI the channel have a reclangular cross section.
Then the cross-sectional area is equal 10 the channel width, IV [L], times
the water depth, II:
Q '" Vwl!.
If, for the moment, we also assume that the width of the channel is
constant, Equation 4.5 implies that:
Q
q.., = = VII= constant. (4 .6)
w
(4.7)
124
Equation 4.7 provides a means for detennining the depth of a
frictionless channel flow of known discharge and specific energy. That is,
given numerical values for E and for qw' h is the single unknown quantity
in Equation 4.7. Although it is possible to solvc Equation 4.7 directly, a
graphical solution is more convenient. A specific energy diagram is a
graph of specific energy, E, versus depth, II, for a given value of specific
discharge, qw (Figure 4.3). From the specific energy diagram, we see that
for any specific discharge there are three depths for each value of specific
energy. (Equation 4.7 is a cubic equation in II, so there are three roots).
Only two of these possible depths, hi and hz, are positive; the third depth,
113, is negative and can be disregarded in that it has no physical meaning.
The two positive depths that are possible for a single value of the specific
energy and the given specific discharge are called alternate depths.
The specific energy diagram (Figure 4.3) exhibits a minimum in E,
which corresponds to the condition (called the critical condition) in which
the solutions hi and 112 of Equation 4.7 coincide. Thus, under critical
conditions, a flow with specific discharge, qw' has the minimum specific
energy. The corresponding depth is known as the critical depth, he.
Critical flow is a unique condition in which, for any given value of specific
discharge, thcre is only one possible water dcpth (i.c., the critical dcpth).
For values of E with two possible (alternate) depths, one will be larger
than the critical depth and the other will be smaller (Figure 4.3). The flow
depth hi corresponds to II flow that is deeper and slower than the critical
flow and is called subcritical flow. llle flow with depth il2 is shallower
and fastcr than the critical flow, and is termcd supercritical flow.
The original problem poscd at the beginning of this discussion was to
detennine the water velocity and depth for a given discharge and stream
cross section. Application of the Bernoulli equation and the continuity
equation led to the specific energy diagram. This suggests that an answer
to the question dcmands additional knowledge (the specific cncrgy of the
flow) and that two combinations of depth and velocity are possible. Thus,
we will have to examine the question further to delennine how specific
energy and depth are related.
125
1.5 -,------"71
E
-
'"
£0.5
a.
�
c
-O.5+�----_r---��_,-----___l
o 0.5 1 1.5
Specific Energy, E (m)
I
Figure 4.3 A specific energy diagram for q", = 0.5 m 2 S- . The three
depths, "1 ' "2' and 113, are those allowed by theory for a given specific
discharge and a given specific energy (0.75 m in this figure). The three
depths arc the solutions to the cubic equation relating E, II, and q",.
Because negative depths have no physical meaning, the upper pan of the
diagram, with positive values of It, is used in practice.
A comparison between now through a hose with a contraction and the free
surface analog, a stream with a shallow risc, or step, in the bottom
illustrates the role of specific energy in understanding problems of open
channel flow (Figure 4.4 ). From the equation of continuity and Bernoulli
equation, it is easy to compute the velocity or pressure in the narrow
section of the hose if we know these values for the larger section. As found
in Chapter 3, V2> VI and P < PI for a steady, homogeneous, frictionless
2
now through a pipe contraction such as that illustrated in Figure 4.4.
Unlike the now through the hose contraction, the stream would seem to
have a "choice" as it nows over the step. To constrain the now we will
126
stipulate that the width efthe stream is eenstant and that flew remains
within the ehannel. What will the surfaee an as the flew eresses the step:
rise.. fall. er remain unehanged? Assuming the transitien is essentially
frietienless. the Berneulli equatien {4.2} ean be written:
A \ -
._ . fl- ______ ? ————————
_...
U1
_....
U:
_...
U]
.........
{_f1
U /
if: if;
j"+Jn: 1 ' +hg+hm [+8]
"fg "' I?
where the stream hettem at. the upstream leeatien is ehesen as the datum
and he is the height ef the step. Restating Equatien 4.3 in tenns e'fspeeifie
energyt
El = E3+.-i't:= H. {4.1};
Heeause this is a frietienless Ilew. the tetal energy H must be eenstant.
Hewett-er. Equatien 4}} shews that speeifie energy will net he eenstant if
the elevatinn et" the ehannel hettem ehanges.
We begin by eensidering the ease in whieh the flew upstream efthe
step is suheritieal. Presenting that its flew yeleeity and depth are knewn,
we earl eempule the upstream speeifie energy.
1'.
a; = at +gi~ teen
as
and we ean lee-ate El en a speeifie energy diagram heeause the speeifie
diseharge alse is knewn (Figure 4.5]:
(4.11)
From Equation 4.9 we next can find £2 ' the specific energy over the step;
it is equal to EI minus the height of the step, Liz. As the flow moves over
the step it gains potential energy at the expense of specific energy. We can
make Ihis subtraction directly on the energy axis of the diagram. We now
have the specific energy over the step, £2 ' but we still have to determine
which of the two depths possible for the single values of £2' h2 or hi. is
corree!.
1 ,----0
E,
0.9
0.8
0.7
Subcriticallimb
0.6
-
E
-
""
£ 0.5
C.
� 0.4
-.6.z-
0.3
0.2
0.1
Supercriticallimb
o
o 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Specific Energy, E (m)
2
Figure 4.5 Specific energy diagram for now over a step; q", = 0.5 m 5-1,
The specific energy at station 1 , 0.8 tn, is represented by the vertical line to
the right. The specific energy at station 2 is 0. 1 m less than that at station I
because a step up i n the bottom has increased the potential energy, thereby
decreasing the specific energy. £2 is shown as a vertical line on the
diagram, 0.1 m to the left of (less than) £1.
128
establishes over the step, the change must proceed along the same specific
discharge curve, because q", is constant. Thus, as the flow moves up the
step with decreasing specific energy, the depth decreases, first to h2, and
then perhaps to hi.
To go from hi to 112 requires only a rise in the bed of elevation z. For the
flow finally to reach II; it must pass through the minimum specific energy
at depth 11o, and then procced to hi. But to go from ho to hi would require
an upward step of a certain critical height followed by a downward step
such that the final height of the step is less than the initial height (Figure
4.6). It is obvious then, because our step does not include this hump, that
the ollly depth accessible to Ihe flow is the first choice, "2'
.. .. ......................
:: '
"
" '.
" "
.
'.
•
.
"'" 112
'
1
............... ----.--
t,
",
6.Zaitiwl
[
Figure 4.6 A critical step is rcquircd to reach an alternate depth. If the
upstream flow (shawl/to file left ill the diagram) is suberitieal, the
hi
alternate depth can be reached only if the flow follows the specific
energy diagram around the "nose" of the curve. This means that the flow
must be critical at some point. The only way that critical flow can be
created under the conditions we are considering is by having a step
sufficiently high to cause the specific energy to drop to Eo.
We now know the water depth over the step, "2' and the height of the
step, z, but we still don't know if 112 + D.z is larger or smaller than hi; that
is, whether the water surface elevalion increases or decrcases over the step
(see Figure 4.4). To resolve this question, Equation 4.8 can be rearranged
in the fonn:
129
We already have found from Figure 4.5 that 112 is less than hi; therefore,
U2 must be grea\er than VI if the discharge is constam, and so,
U,' U'
1 >0.
2g 2g
-'-'-_
(4.12)
So, in a subcritical flow the water surface drops and velocity increases as
water flows over the step. This conclusion explains why canoeists are wary
of sudden changes in stream elevation, as they often mean that large rocks
are below.
Notc that if the upstream flow is supcrcritical with specific cncrgy E1
and depth 11; the specific energy will still decrease as the flow crosses the
step. However, in this case the decrease in E is accompanied by an
increase in depth from 11; to II;, and, therefore, a decrease in velocity
because the discharge remains constant (see Figure 4.5).
It is also important to stress that if the step height, 6z, is greater than the
differcncc bctwccn E 1 and the spccific cnergy of the critical flow (Figure
4.5), EI would not be sufficient to allow for the flow of the specific
discharge qw across the step. [n these conditions it is observed that the step
modifies the flow and induces an increase in E in the surroundings of the
step, thereby allowing the flow to occur across the step with minimum
specific energy (i.e., in critical conditions). [n this case the step functions
as a broad crested weir, as explained in Section 4.3.
130
either the depth or the velocity is at a minimum value. We can derive a
relationship between the velocity and depth at this condition of minimum
specific energy as follows:
(4.\3)
At E equal to a minimum,
(4.14)
or
(4.14)
and
(4.15)
This last equation states that when the mean velocity is equal to JgJ'o' the
specific energy is a minimum.
The line defining critical flow conditions divides the specific energy
diagram into two regimes (Figure 4.5) . As noted in the previous section,
for any value of specific discharge, the regime above this line has slow,
deep flow relative to the flow below the line. This upper regime is called
subcritical flow and the lower regime is referred to as supercritical flow.
Critical flow occurs along the dividing line of these two regimes. For any
value of C/o. the specific energy will be at a minimum when the flow is
..
(4.16)
I3 I
How does a flow come to be either subcritical or supercritical? One of the
obvious ways that flow can be controlled is by using some type of
structure. Wc intuitivcly expect that a dam will cause a river to back up
and that the resulting flow within the rescrvoir will be deep and suberitieal.
On the other hand, the rapid, shallow flow of water down the face of a
steep spillway must be supercritical. For the case of frictionless flow
considcred so far, subcritical flow can be thought of as controlled by a
structure downstream (e.g., the dam is downstream of the suberi{ical
region) and supercritical flow controlled by a structure upstream.
The control that a structure, such as a dam, has on flow through a
channel can be exploited for making measurements of channel discharge.
Upstream of the dam the flow is subcritical. The flow over the spillway is
supercritical. Therefore, somewhere over the dam crest the flow must
reach the critical condition where U = J
gh, that is, the Froude number
equals 1 . In fact, from the discussion of critical flow in the previous
section, we know that we can force any channel flow to become critical by
placing in the flow a properly designed step across the channel. An
artificial obstruction like a step or dam over which all the water in a
channel must flow is referred to as a weir. There are various types of
weirs, ineluding sharp-crested, broad-crested, and notched weirs. Here we
will focus on broad-crested weirs.
Consider the flow over the broad-crested weir shown in Figure 4.7.
Because we are interested in only a short reach of flow we can expect that
Frictional energy losses will be small enough that we can use the Bernoulli
equation. We also will assume that the flow is steady on the time scale
required for making a measurement. Using the Bernoulli equation to relate
the flow over the weir crest (position 0) to the flow upstream oFthe weir
(position I), we obtain:
(4.t7)
The upstream flow is subcritical, so UlIghl < I. In fact, generally the flow
just upstream of the weir will be relatively slow and deep, in which case
we can make the assumption that the Froude number is small enough that
UlI2K« II. In this case, we can neglect the velocity tenn on the left side
of Equation 4. 1 7. lfwe choose our datum to be at zl, the level of the
channel bottom upstream of the wcir, Equation 4 . 1 7 reduces to
132
h lI'eir
"0
I ,
/ 1\,
Zo
.
Figure 4.7 A broad-crested weIr. The structure causes water In the channel to
• .
"back up" until the flow over the weir is critical. Because we know how velocity
and water depth are related for critical flow, measurement of the height of the
water in the pool behind the weir can be used to detennine the discharge in the
channel.
U6
IIt==o+lIo+-=/-I
2g
or, if we define h eir as the height of water upstream of the weir relative to
...
(4.18)
At some point over the weir crest, the flow must be critical. At this point
U6 gila or UJl2g
= hr/2
Substituting this into Equation 4. 1 8 gives:
=
"0 3 o
110+"2
h..
·nr=
"2h .
=
or
]fwe then substitute Ihis expression for 110 into Equation 4. 1 8, we find that:
133
Rearranging to obtain an expression for velocity,
(4.19)
(4.20)
134
4.4 The Effect of Bed Roughness
dN to
"/ '
_
(4.21)
_
dx -
where Sj is the slope of the total energy line as indicated in Figure 4.8 . Sjis
known as the friction slope becausc it is frictional head loss that causcs H
to decrease. Note that Equation 4.21 can be writtcn as:
S friction slope
f
=
- / =
"L/ L
Ui,{I, 2g -
-L
h,
- I•
".
H,
1
S = slope of bed
I H,
1-
-t-
L
-
"
-
,
-
Datum (: 0)
=
. . .
•
.
Figure 4.8 Fnctlon along the bed causes head loss. The schematic diagram
•
1 35
constant in the downstream direction and the elevation of the bed
decreases, the Bernoulli equation indicates that energy is lost to friction as
the watcr flows along thc channel.
dE + d:=_S.
(4.22)
d... dt' J
(4.23)
constant and that friction slope is equal to the bed slope means that
uniform flows dissipate as much energy as is provided by the bed slope. In
other words, if the flow is unifon11 the energy dissipation in a channel is
thc same as thc encrgy added by the changc in bed elevation between the
upstream and downstream sections of the channel. Another way to look at
uniform flows is that they exhibit a balance between friction forces and the
component of gravity in the flow direction.
Equation 4.23 can be used to infer how flow is controlled in a river
channel. To picture this, imagine a laboratory experiment in which water is
caused to flow ovcr a bed of sand and creatc a channel. Thc inflow might
be controlled by a reservoir of water and the upstream specific energy
would then be detennined by the height of the water in that reservoir. We
might suppose that initially the friction slope would exceed the bed slope
and thus that dEldx < O. This means that the specific energy would
decrease in thc downstream direction through bottom friction. If the flow
possesses sufficient energy to move the sand, we might expect the bed to
erode and, therefore, increase the bottom slope, S. As S approaches Sf'
dEldx approaches zero and the flow becomes unifonn, the slope and the
roughness detenn ining whether the flow is subcritical for a given value of
specific energy.
The above analysis is really just an intuitive examination of the
meaning of Equation 4.23. In actual practice, this equation can be used to
dctcrmine how depth and velocity vary in channels but the proccdurcs arc
rather involved (see Chanson, 2004, for a more complete discussion). A
136
simpler approach is to investigate the relationship between velocity and
head loss in a manner similar to that used for flow through pipes.
Friction slope is the head loss per unit length and can be represented as
II,JL (sec Figure 4.8). In Chapter 3 , an equation for the head loss in a pipe
was presented:
L u'
111-[--,
D2g
(.3.30)
III. _ [ 1 U2
' (·U4)
L - 4R1/2g
where RI/ is the hydraulic radius and is defined as the ratio of the cross
sectional area of the flow to the channel wetted perimeter. The numerical
factor 4 is introduced because for a pipe with a circular cross section, the
hydraulic radius is equal to the diameter divided by 4. For a rectangular
channel (Figure 4.9) the cross-sectional area is wh, and the wetted
perimeter is 211 + w. Thus,
wh
RI, = 0;;"'-- (4.25)
211+ 11'
If the width is much greater than the depth, w» II, then
wll
RI/ .. -=11. (4.26)
W
1 37
intuitively might be thought to be "laminar" when, in fact, it is not, a
mistake made quite onen. As a concrete example, suppose water at 20°C
nows in a wide channel with a depth of 0.20 m and a mean velocity of
0.30 m S-I . This now has a Reynolds number of about 6000, well into the
turbulent range.
"
I
. . . .
FIgure 4.9 Cross sectIOn of a rectangular channel. The hydraulic rad1Us IS
•
the ratio of the area of flow (wh) to the wetted perimeter (w+ 2h). For a
very wide rectangular channel, the wetted perimeter is very close to the
width of the channel so the hydraulic radius approximately equals the
water depth.
(4.27)
13 8
C with a parameter that is a function of roughness only. In 1 891, another
Frenchman, A. Flamant, proposed,
kRI16
C = 1/ . (.US)
"
U=�RJPS'I2. (4.29)
"
Note: For natural streams, the report by Barnes (1967) is an excellent reference.
139
and for the other /I equals 0.075. Assume that the slope is 0.0006 mlm, and
the hydraulic radius is 3.0 rn. In this example,
0
(.4030)
I
V,. 1. (3.0)!Il(0.OOO6)"2 =0.79 rn S-
1 oms
V=�,,2J3SI12.
(4.31)
"
140
turbulent channel flow is logarithmic rather than parabolic (Figure 4. 1 0) .
The logarithmic velocity distribution of a turbulent channel flow is
given by the Kam1an-Prandtl equation:
(4.33)
Often we don't know enough about a stream to make even good estimates
of flow velocity or discharge without making some direct measurements.
Channel cross section and slope can be measured using meter sticks and
surveying equipment. Channel roughness cannot be measured directly.
Instead, it must be detennincd by measuring all of the other variables in
Manning's equation or the Karman-Prandtl equation and rearranging the
equations to provide an expression for 11 or k,.. To do this requires
measurements of velocity in the channel using a current meter. Current
meters measure velocity at a point. How can we most efficiently use point
measurements to estimate discharge?
Mean velocity in a channel is the sum, or integral, of the velocities at
each point in a cross section divided by the cross-sectional area:
u= -
[ f·f'u(y.:)d:dy.
A • • (4.3")
141
1 ,-------,-
0.9 ,
,
Turbulent ,
flow ,
0.8-
,
,
,
,
'" 0.7-
,
.
,
."
,
I
,
,
,
;: 0.6-
-
,
,
e.
,
,
"
J//
'tJ 0.5-
"
>
:; 0.4
"
-
a: 0.3-
-
-
-
-
-
0.2- -
Laminar /"
0.1-
o ¥ --
�/ ��
�
�........
-
..
o 0.5 1 1.5
Relative velocity, it/V
Figure 4.10 Velocity distribution in a channel. For laminar flow (almost
never observed in natural channels), the velocity profile is parabolic,
similar to the case examined for pipe flow in Chapter 3. For turbulent
flow, the velocity near the bed changes much morc rapidly than it does in
laminar flow. The velocity profile for turbulent flow generally is described
using a logarithmic equation.
1 1
V= -
11'11 I'
\II
0
lI(z)dz =
110
-
I'u(z)dz.
If we substitute the Karman-Prandtl equation for lI(z) and carry out the
integration, we obtain:
(4.35)
142
4.10). The mean velocity is some intennediate value and thus there will, in
general, be some level Zu at which the velocity u(zu) is equal to U. Setting
Equations 4.33 and 4.35 equal to each other and solving for the value ofZu
gives zu= 0.37h (Figure 4. 1 0). This means that all we have to do if we
want to estimate mean velocity is to measure the velocity at a level
roughly 0.411 above the bottom or 0.611 below the surface.
Of course, most natural channels are not rectangular in cross section so
that estimating U with Equation 4.33 letting II = average depth cannot be
expected to be very accurate. However, we still can make use of the results
for rectangular cross scctions by dividing a more complicated channel
cross section into a number of rectangular subsections as illustrated in
Figure 4. 1 1 . For each subsection, the width and depth can be measured
with a tape and meter stick and the mean velocity can be approximated as
the measured velocity at a depth 0.611 below the surface. Taking Ihis
approach, we can re-writc Equation 4.34 as:
U = �L N
,. ,
II',II,U" (4.36)
Q UA = L 1Il,II,U,. (4.37)
,.,
=
Once we know mean velocity, slope, and depth (or hydraulic radius),
we can determine the value of Manning's 11 using Manning's equation. If
we did this for a number of different discharges in a given stream reach we
could estimate a characteristic value of 1/ for the re<lch that could then be
used to predict the mean velocity and discharge for depths other than those
for which direct mcasurements wcrc madc. Leopold and Wolman (\957)
derived an empirical relationship between the friction factor, water depth,
and DS4' the size of bed material of which 84% of bed material is finer (see
also Leopold, 1 994, for other examples of data supporting the empirical
relationship). In terms of Manning's 11, the empirical relationship is
(assuming that RII = II)
143
I W'I
B 9
7
,
2 --- -'
3 6
h, 5 --- ,
- --
-- -- -
�-
II
- :EL[, (-"-) 1
"
116 _.Olog. lo
DS4
+ 1.0 . (4.38)
144
"
'"
• 0.2 i---- --- -;======;---
+ Measured
o Predicted
o
0.1
+ -
0.06 o
o
+
0.04
There are many important applications of the theory developed above and
some oflhese will be explored in subsequent chapters. Here we briefly
consider examples of how the relationships derived above can be useful in
interpreting infonnation about flows in channels.
145
Recall that our consideration of Manning's equation suggested that
discharge should be related \0 the 5/3 power of stream depth for a stream
with a rectangular cross section (Equation 4.32). Streams rarely have
strictly rectangular cross sections, of course, but our analysis could be used
to suggest that a power relationship, perhaps with an exponent different
from 5/3, might be useful for describing stream data. In fact this often is
the case. For Brandywine Creek near Comog, an equation Q = 28.26h2.26
fits the observations of discharge and depth reasonably well (Figure 4. 1 3).
As we will see in the next chapter, such power relationships are used
routinely in making continuous measurements of stream discharge,
measurements that are essential for solving many practical and theoretical
hydrological problems.
10
�
r
"'
,
"
E
�
Ol
Q) 1
"
'"
�
co
.c
"
"'
C
,-
0,1
0.1 1
Water depth, " (m)
Figure 4.13 Stream discharge is related to a power of stream-water depth
for many streams and rivers. For Brandywine Creek at Comog, the
exponent in the empirical relationship between Q and h is 2.26.
146
variations along the longitudinal profile ofa river, that is, on the changes
in slope, width, depth, and velocity in the downstream direction. Leopold
et al. ( 1 964) and Lcopold ( 1 994) present data of this type for a number of
rivers. Relationships tend to be straight lines on logarithmic axes (i.e., to
be power relationships) for most river systems; the Brandywine Creek
basin is an example (Figure 4.13). Of course, slope decreases as a river
progresses from the headwaters to the sea and discharge generally
increases bccause of the steady increasc in drainage basin area. The
observed increase in width and depth also seems intuitively plausible.
Velocity also increases in the downstream direction.
There are many interrelated changes that occur among hydraulic
variables in streams and these must be considered collectively in
explaining the observed changcs in mean velocity of a river system. Smith
(1 974) assumed that Manning's equation held for a hypothetical river
system with constant roughness (Manning's /I constant in the downstream
direction), that discharge increased directly with drainage area, and that the
basin was eroding at a constant rate. He used bank filII discharge, the
discharge in the stream when it is filled right to the top of the channel, in
his analysis. Using empirical equations for sediment transport in the river,
he then showed that slope should be proportional to discharge to the
negative 0. 1 8 power, that velocity should be proportional 10 the 0.09
power of discharge, that depth should be proportional 10 Ihe 0.27 power of
discharge, and that channel width should be proportional to the 0.64 power
of discharge. The corresponding powers for the Brandywine basin (the
slopes of the lines in Figure 4. 1 4) are -0.69, 0.30, 0.20, and 0.43. The
magnitudes of the powers for this particular basin do nOI correspond all
that closely with Smith's theory, but again the fonn of the relationship is
similar and the hydraulic interpretations help provide a consistent
conceptual basis for explaining natural phenomena. These ideas also are
useful in assessing the potential impact of modifications on a stream
channel (e.g., how stream channelization affects all of the pertinent
variables in a channel as a result of changes in slope).
147
1 100
8 0
E E
0
- -
0
- -
�
00
•
"
� � 10
"
c- 0
-
-
"
•
c �
0.1 1
1 10 100 1 10 100
Bankfull discharge (rna 5-1) Bankfull discharge (m3 s-')
0.1 10
"
-
>-
-
E
,
-
0 'i3 .::--
E 0'.
- 00 -
0
'" 0.01 0 � E 1,
"
• 0 c- o
c- o;,
0 0 •
-
<J) "
0.001 0.1
1 10 100 1 10 100
Bankfull discharge (m3 5-') Bankfull discharge (m3 5-')
Figure 4.14 Hydraulic parameters in the downstream direction for the
Brandywine Creek basin.
Data from Miller ct a1. (1971).
• Specific energy E [L] is the energy per unit weight of water in the stream
relative to the stream bottom: E= U2/2g + h. Total energy, H [L], is
therefore E + zb' {Section 4.2}
148
among spccific energy, specific dischargc, and water depth follows from
the definition of specific energy: E q�.!2 gh2 + II, where qw is the
=
specific dischargc, the discharge per unit width of channel [L211]. For
given valucs of E and qw' this equation is cubic in II giving three
mathematically allowable depths of water for the given conditions, only
two of which are positive. These two physically meaningful depths are
callcd thc altcmatc dcpths. {Section 4.2}
slow and tranquil. For Froude numbers greater than one, the flow is
supercritical, rapid and torrential. {Section 4.2.2}
weir to the water surface in the pool behind the weir and we is the width
of the weir crest. {Section 4.3}
• For steady, unifonn open channel flows, the friction slopc is cquallo the
bottom slope, and depth, velocity, and specific energy are constant.
{Section 4.4}
• The hydraulic radius R" [L] of a stream channel is defined as the ralio of
the cross-sectional area of flow to the wetted perimeter. {Section 4.5}
• Using the friction factor discussed in Chapter 3, the Chezy equation can
be derived: U �8g /f JSR" CJSRI!' where C is Ihe Chezy number
= =
velocity to slope and hydraulic radius that is very widely used. Values of
the roughness parameter, II, have been tabulated for a range of channel
conditions. {Scction 4.5}
149
• The vertical profile of velocity in a river channel often is assumed to be
logarithmic by virtue of a derivation due to Prandtl and von Karman. The
mean velocity for flows with a logarithmic velocity profile occurs at a
distance of 0.411 above the bottom. The common method of gaging
streams using current meter measurement relies on this approximation.
{Sections 4.5. 1 and 4.6}
A. Calculate the specific discharge (m2 S- I ) and specific energy (m) at the
upstream station.
B. Calculate the specific discharge (m2 S-I) and specific energy (m) at the
downstream station.
150
w
w/2.S
.
Figure 4.15 Canal cross secllon for Problem 2.
15 1
D. Use the equation you developed in 3C to find the critical Reynolds
number for the transition from laminar to turbulent flow in a relatively
wide, rectangular channel (such that RH :::; II).
Problem 4. Use Equation 4.32 to estimate the depth and mean velocity of
a flow in a channel with a slope S 0.003, width )II = 15 m, discharge Q =
=:
Dunne, T., and L.B. Leopold. 1978. Water in environmental planning. San
Francisco: W. H. Freeman. Chapter 16, pp. 590-660.
Leopold, L.B. 1994. A view ofthe river. Cambridge: Harvard University
Press. Chapter 9, pp. 148-167.
152
5 Catchment Hydrology: Streams and Floods
5. 1 Introduction
5.1 Introduction
153
As an example, consider the situation for Sacramento, California. The
American River basin lies to the east of Sacramento. The basin area is
about 5000 km2 with headwaters in the Sierra Nevada. Three forks of the
river flow through scenic canyons in the mountains, valued by white-water
sports enthusiasts and hikers, and merge just above Folsom Dam. Below
the dam, the American River flows between levees through downtown
Sacramento. More than one million people live in the Sacramento
metropolitan area, within the floodplains of the American and Sacramento
rivers.
.
154
to the reservoir and later releasing this water at a rate much lower than the
highest flood rates. The problcm of deducing how outflow from a reservoir
is relatcd to its inflows is known as flood routing. As will bc described in
this chapter, flood routing depends on some of the principles of fluid flow
that were covered in earlier chapters.
In this chapter we discuss some selected elements of surface hydrology
and show how the fom1Ulation and solution of surface-water problems are
developed from fundamental hydraulics. We consider the propagation of
flood waves through reservoirs, flood routing, and the frequency with
which a flood of a given magnitude would be expected to occur. There are
many facets of surface-water hydrology that are not included in this
chapter. The material presented here is intended to introduce the subject,
particularly with regard 10 river floods.
8,0.2 8,0.6
-
Sl � 0.4
li:
0.1 -
o
0.2
o • • O+---�-
0
-�r----r--
o 6 12 18 24 6 12 18 24
Time (hr) Time (hr)
Figure 5.1 The hydrograph. River stage (a) and discharge (b) as functions
of time.
155
The discharge hydrograph is not measured directly, but is inferred from
the stage hydrograph. In the United States, the United States Geological
Survey is responsible for the measurement of river stage at over 7000
gaging stations. At many of these locations, data are continuously recorded
by a float gage installed inside a stilling well (Figure 5.2). As the river
rises and falls, the float moves with the water level and the motion of the
float is recorded. The stilling well serves both to protect the float
mechanism and to dampen the rapid fluctuations of the water surface due
to local disturbances and turbulence. In recent years, the float mechanism
has been replaced by a pressure transducer, a device that records the
pressure at the bottom of the stilling well. The hydrostatic Equation 3.6 is
used to calculate water depth from the measured pressure. Radar is also
used to measure stage in some rivers.
The record of water depth versus lime measured at a gaging station is
the stage hydrograph (Figure 5 . 1 a). To construct a discharge hydrograph
from these data, determinations must be made of the river discharge for
various values of flow depth or stage. By measuring stream velocity at
numerous positions across the width of the stream, the discharge at a given
stage can be detennined as described in Section 4.6 . These measurements
are generally made with a current meter either attached to a wading rod or
suspended from a cable, depending upon the depth and discharge of tile
fiver.
.
After discharge has been measured for a range of flow depths, a rating
curve can be constructed that relates stage to discharge. Rating curves
typically are nonlinear and often are approximated using a power function
(Figure 5.3). To be useful, a stream gage should be located at a point
where the stage-discharge relationship will not change from year to year;
such a location is called a control section. An example of a control section
might be a place where the flow is constricted by a natural rock fonnation
or by bridge abutments. A station where the cross section changes almost
continually because of streambed erosion or deposition would not provide
good control.
156
Stilling well
Recorder
---' - ;....-----,
Inta ke
pipe
After the rating curve for a gaging station is detennined, it can be used
to convert each value of stage to discharge. Consider, for example, the
stage hydrograph shown in Figure 5.1 a. The stage peaks at 0.35 m (at time
= 6hr). Using the rating curve shown in Figure 5.3 , the corresponding peak
157
seasonally in response to changing weather conditions. In contrast, the Frio
River is an ephemeral stream in Texas that is dry for long periods with
occasional large floods. Little Blackfoot River lies in a snowmclt4
dominated catchment. The hydrograph from a strcam like this is
characterized by a single, extended period of high discharge in the late
spring and early summer when the snow melts. Even within a single
climatic zone, differences in catchmcnt shape, geology, and vegetation will
produce differing hydrographs in response to similar precipitation events.
, ,5
(a) Arithmetic axes (b) Logarithmic axes
"----"
lO-i- -- ------r----,
, ,
- -
•
• ,
,
, ..
• • "
E E
.'
" "
....
•
- -
CI 0.5 CI 0.1
"
o 0,01 +':...-----�
o 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.1 0.5 1
Stage (m) Stage (m)
Figure 5.3 Rating curve (stage-discharge curve) for the Snake River above
Montezuma, Colorado. Thc equation for the lines is Q =
76.5 (stage)4. 1 .
Data from Boyer (1994).
E
- -
•
•
E
,
E
" -
0
c
•
-
,
-
'"
0.
-
,
•
-
•
u u
"
•
,.
•
is 0.
,
158
The floods shown in Figure 5.4 may be thought of as waves that propagate
downstream. The way in which flood waves propagate-their speed, peak
height, and peak discharge-is an important area of study in hydrology.
Planning for the response of rivers to both normal and extreme rainfall
depends on this understanding.
E 75
•
•
'" 50
"
•
"" 25
0
.•
-
0 0
OcI ' , St Jan l , S2 Apr t , 82 Jul l , S2 Sep 30, 82
(b) Frio River below Dry Frio River near Uvalde, Texas
� 1 00
:.,
E 75
•
-
l!'
50
•
""
��-----1
25
� 0
5 J,. , . ro . , . ro J." . ro Sep 30. 76
� '.7
� (c) Little Blackfoot River near Garrison, Montana
50
•
E
-
•
25
l!'
•
""
o
.•
-
o 0
Oct 1 , 90 Jan l , 91 Apr l , 91 Jul l , 91 Sep 30,91
Figure 5.5 Annual hydrographs for three rivers in catchments in three
climate zones illustrating the differences among streamflow in perennial
(a), ephemeral (b), and snowmelt (c) streams.
159
wave to be reduced, or attenuated, as it travels downstream. This
attenuation will depend on several factors: the ability of the channel
system to store and release walcr, the channel roughncss, the amount of
vegetation in the channel and its floodplain, the relative straightness of the
channel, and the number of bridges and other obstructions in the path of
the flood wave.
The picture is complicated further because the lateral inflow of water
from tributaries and groundwater will increase the volume of water in the
channel as onc moves downstrcam. In gcncral, dischargc increascs
downstream as drainage area increases. An example may serve to illustrate
the relative effects of drainage area and flood wave attenuation in a river
or stream network. Hydrographs for the South Fork of the South River at
Waynesboro, the Shenandoah River at Front Royal, and the Potomac River
at Washington, D.C., for the same event are quite different (Figurc 5.6).
The South River is a tributary of the Shenandoah and the Shenandoah
flows into the Potomac weU above Washington. Perhaps the most
noticeable difference among the hydrographs is the highest peak flow at
Washington, the lowest peak at Waynesboro, and the intennediate peak at
Front Royal. This seems to contradict the "attenuation by friction and
storage" hypothesis stated above, but as we have already indicated, the
behavior is complicated by the increasing drainage area in the downstream
direction. The effect of increasing drainage area can be taken into account
by dividing discharge by drainage area to "normalize" the discharges
before they are compared. The nonnalized hydrographs indicate thai the
flood wave is indeed attenuated as it moves through the channel (Figure
5 .6). The normalized hydrographs show clearly the time delay between the
flood peaks and the reduction in peak discharge per unit area in the
downstream direction.
160
5 (aJ Discharge hydrograph
000
-
4000
/ Potomac
1n 3000
•
§. 2000
'"
1000 � '/" �enandoah
0
/" South
0 2 4 6 8 10 12
Time (days from April 13, 1987)
�
-
:.
E
•
•
•
• 0.1
,;
'"
o 2 4 6 8 10 12
Time (days from April 13, 1987)
Figure 5.6 Hydrogmphs for three nested basins. Discharge increases
downstream in the stream network due to increased drainage area (a).
Dividing discharge by catchment area (b) reveals both attenuation and
delay of the flood wave as it moves downstream.
161
equations we need to solve for the two unknowns, depth and velocity.
If a section of river channel, or reach, is taken as the control volume,
then the inflow is given by the upstream hydrograph and the outflow by
the downstream hydrograph. We will assume that the lateral inflow of
water is small relative to the change in storage associated with the flood
wave and, therefore, can be neglected. This assumption may not be valid if
the reach is too long. Analyzing the movement of a flood wave is different
from analyzing he t steady flows considered in Chapter 4. Depth and
velocity change quickly with time during a flood and, therefore, we must
solve this as an unsteady flow problem.
In simplest tenns, neglecting any lateral inflow, the continuity equation
for unsteady flow can be wrinen as
dV
/ - 0.. (5.1)
d,
-
=
162
flood routing problems.
The first step in obtaining a numerical solution to the flood routing
problem is to rewrite Equation 5 . 1 in tenns of finite differences. To do
this, we choose a computational time interval, !J.I, separating two instants
in time, I" and f" + I ' Lei the subscripts /I and 11 + 1 refer to values of each
!-v. Iw + 1n;!
variable at these times. Equation 5.1 can then be approximated as:
v••
(5.2)
III 2
-
At each time step, the terms at time 1/ are known; the terms at time /I + 1
are the unknowns we are solving for. At the initial time step (1/ = I),
volume and outflow must be specified. We refer to the values of VI and 01
as initial conditions. In addition to the initial conditions, we have to
specify the inflow hydrograph during the interval of interest. That is, In
and lIP + I are known at each time step. The unknown terms being solved for
at each time step are V,, + I and 0,, + I ' Rearranging Equation 5.2 to put the
unknowns on the left side of the equation gives:
(5.3)
Given conditions at time I", the quantity on the left-hand side of Equation
5.3, (2 V" + I /!J./) + 0" + I ' can be calculated. We need a second equation, as
described below, to separate this into values of V" .. I and 0" .. I ' Once these
are detennined, all of the quantities on the right-hand side of Equation 5.3
are known for time t" I ' The procedure is then repeated to solve for the
..
163
in magnitude and delayed in time compared to those on the same stream
upstream of the reservoir. Understanding how a flood wave is modified as
it passes through a reservoir makes it possibJc (0 design and manage
reservoirs to protect downstream areas.
Reservoirs are formed when a river is dammed. The simplest dam is
one designed so that when reservoir depth exceeds a certain value, water
will flow over a spillway (Figure 5.7). The spillway is like a weir (see
Section 4.3). There is a fixed relationship between the depth of water
above a weir crest (h ei�) and discharge over it. This relationship provides
....
Reservoir
Q = Cq �2g"�t�r 11',. ,
where lVe is the width of the weir crest and Cq is an empirical coefficient.
With Cq = 0.5, typical for a dam spillway, the weir equation gives us an
expression for reservoir outflow
(5.4)
with 0 expressed in mJ S-I and both h ej� and We in m. Water depth in the
....
reservoir h is given by the sum of hh"tir and the height or the weir
164
(spillway) crest, hspilll<'lIY" Therefore, Equation 5.4 allows us to relate water
depth in the reservoir to the outflow over the spillway.
Water depth in a reservoir is also related to the volume of water in the
reservoir V in a manner dictated by the geometry or topography of the
reservoir basin. For example, a 20 I I bathymetric survey was used to
establish the relationship between depth and volume in Totten Reservoir, a
small reservoir in southwest Colorado (Kohn 2012; Figure 5.8). This
reservoir has a spillway that is roughly 20 III long at a height of 1 0 . 1 III
above the reservoir bottom. Reservoir depth and volume above the
spillway arc provided in Table 5. 1 .
The application of these relationships to the reservoir routing problem
is best illustrated by example. The reservoir in the example has the
geometry indicated in Figure 5.8. The computational time step D.( is
dictated by the resolution of the input hydrograph. For this example we
will use the input hydrograph given in Table 5.2, which has a time step D.I
= 6hr or 2 . 1 6 x 1 04 seconds. We also must specify the initial values of 0 1
and VI before we begin the calculation. The spillway height h''Pi/hl'll), = 10 . 1
m (Figure 5.8) and we will assume h""ei� = 0.0 m initially, corresponding to
a reservoir volume VI = 3.7 x 1 06 m3 and initial outnow 0 1 = 0 m3s-1
(Table 5. 1 ). In other words, the water level of the reservoir is just al lhe top
of the spillway at the initial time I I '
12 2
. S.9SE+06 2.13 137 687
NOle: Assuming III = 2 1 6005 (= 0.25 days), outflow 0 is given by Equation 5.4,
165
14 --
--
12
10
E
_
------------ - - - - - - - \- - - - - - - - - - -
- -
\
Spillway height
li
• B
."
-
--
�
•
6
•
•
a:
4
�----�,----�2�--�3�--�4�--�5�--�6�--�7
Reservoir volume (m3) x 10e
Figure 5.8 Relationship between reservoir volume and depth for Totten
Reservoir in Montezuma County, CO. Such relationships are typically
obtained from topographic and bathymetric surveys of the reservoir area.
Data from Kohn (2012).
166
The solution to Equation 5.3 can be found by constructing Table 5.3.
The quantities in each column are identified by the expressions at the lOpS
or thc columns. To begin, values for In and In + ,'' + 1 from the inflow
hydrograph (Table 5.2) arc entered into columns A and B, respectively, for
each time I". Based on the initial values VI and 01 given above, the first
entry in column C of Table 5.3 is [(2 * 3 .7 x 106 m3)/(2. 1 6 x 1 04 s)] - 0 m3
S-I = 342 6 m3 S- I . The entry in column D is computed using Equation 5.3:
.
2 V, 211:
=--
c
· + 02 = 'I + '2 + 1 01 = 344.2 m ) S- I .
!>J !>J
-
That is, 0 = B + C.
Now we need to figure out how to separate the teon in 01 (column 0,
row 1: 2 V2//::..1 + O2) into values of V2 and O2, From Table 5. I (and Figure
5.8), we know reservoir volume V for a range or values of water depth, h.
We can convert the values of II to corresponding values of 0 using
Equation 5.4 with Iiwe'r = II - Ii'1'''/lI"I1)" This gives us values of V and 0 for
each value of h. From these we can calculate 2V/61 + 0 (column D) for a
given /::..r. The resulting values of "wei,., V, 0, and 2 V//::..r + 0 are listed in
Table 5. 1 for !:J.r = 2. 1 6104 s and Ii.,pill"''')' = 10.1 m. A graph of2 V/!:J.r + 0
x
versus 0 (Figure 5.9) allows us to detenlline the value of 0 for any value
of2V/!:J.t + O. In practice, it is often more convenient-particularly when
doing the calculation in a spreadsheet-to fit a curve to the relationship
shown in Figure 5.9. Since we don't expect outflows thai are larger than
inflows, we focus on getting a good fit to small values of outflow, 0
(Figure 5.9):
167
Time t. lV
e 2V_1
'. 1. +'.�1 - 0. 0.�1
+0"'1
Step n f.�1
Idays) '" M
A • C D ,
Note that Equation 5.5 may not be a good fit for values of 0 > 50 mJ S-I.
Returning to Table 5.3, we find from Equation 5 .5 (or Figure 5.9) that
the value of O2 (E 1) is 0.06 mJ S- I for a Dl value of 344.2 m3 S-I . We
enter 12 in the last column to keep track of the proper time for the outnow
hydrograph.
At time step 11 = 2, we can calculate the tenn C2 in Table 5 .3 as the
term 01 - 2 )( the tcrm E l . That is,
2V., 2V., J 1
�-
.:c...c 02 = � + °2 - 2°2 = 344.1 m s- .
AI M
repeat these steps for each 11 until the table is complete. The calculated
outflow hydrograph is tabulated in the last two columns of Table 5 .3. To
summarize, the series of steps for completing Table 5.3 (using i to
represem the row number and letters to represent thc variables in each
168
column) is:
50
45
::::
I �
-
::: �
Values
��l-----/
from
- - - Equation 5.5
Table 5. 1
40
35
'"
-
;:;- 30
E
-
� 25
0
'"
-
" 20
0
15
10
gOO
� -
�3�
20
--�
�-�
0��
����
�0--
�4OO��
�2�
4 -
0 �
� 0--
�4�
60
--��
48�
0 �
50a
2V/dl + 0 [col O] (m'/s)
Figure 5.9 Outflow versus (2 Vldt) + 0 for the reservoir routing example
bascd on low outflow valucs in Table 5. 1 . Equation 5.5 providcs a fit to
the observcd relationship useful for completing the calculations in the
reservoir routing example.
I . Bi + Ci = Di.
2. Di + Equation 5.5 or Figure 5.9 produces Ei.
3. Di - 2 x Ei = C(i + 1).
4. return to step 1 for next i.
The predicted outtlow hydrograph shows how the reservoir has worked 10
reducc thc efTect ofthc inflow flood (Figure 5. 1 0) . The rcservoir stores a
portion of the inflow during high inflow periods and releases the water
when the inflow subsides. Again, note that the effect of this storage is to
delay the peak discharge and to reduce its magnitude.
It is instmctive to consider how the outflow hydrograph for the
reservoir routing problem changes with reservoir characteristics. First, if
the reservoir werc larger, it would store morc water. For the same inflow,
169
the rise in water level in the reservoir would be smaller and, therefore, the
increase in outflow would be smaller. The outflow hydrograph would have
a lower pcak valuc and would be spread over a longer period of time.
Second, we might conceive of an outflow mechanism that released water
more slowly than the spillway (or ofa spillway with a lower weir
coefficient). This also would have the effect of decreasing the peak
discharge because the water would have to build up to a higher level to get
thc same outflow. If thc watcr level in the reservoir were initially below
the level of the spillway, then there would be no outflow over the spillway
until the inflow could raise the water level in the reservoir to the level of
the spillway. In the example shown in Figure 5 . 10, 9.6 x 105 m3 of water
enters the reservoir over thc eoursc of 3.5 days, an amount roughly
equivalent to the volume of a 1 .0 m-thiek slab of water in the reservoir just
below the height of the spillway. Therefore, if water depth in the reservoir
were initially > 1.0 m below the spillway, the water surface would never
reach the level ofthc spillway during this flood.
10
170
located on relatively small rivers or streams. However, large dams, such as
the Grand Coulee Dam on the Columbia River, Glen Canyon Dam on the
Colorado River, and Folsom Dam on the American River, have outlets that
are more complex than simple spillways. Water is released through
carefully controlled outlets and over spillways. This allows the volume of
water in the reservoirs behind the dams to be varied seasonally and the
discharge of water through the dams to be adjusted according to the
operational policy for the dam. By gaging inflow and considering the
trade-offs between volume of water in the reservoir and outflow rates,
reservoirs can be managed for maximum benefit. For example, during
periods of drought, extnt water can be stored. However, if water levels in a
reservoir are kept high in amicipation of drought, the reservoir capacity for
storing floodwaters is decreased.
Some of the difficulties in optimally managing a reservoir are
illustrated by an example from the American River. In 1986, before the
flood on the American River, the operating procedures for Folsom Dam
called for a flood-storage capacity of 4.93 x 1 08 m3 beginning on
November 1 7 and extending until February 8. Subsequcm to February 8,
the allowable storage was varied according to how much the accumulated
precipitation for that season had been. The operating policy also called for
a maximum controlled release from the dam of3.26 x 1 03 m3 S- I up until
the reservoir level reached full pool. Subsequently releases were dictated
by emergency spillway procedures designed to protect the structure from
failure.
When the flood began in February 1986, Folsom Reservoir's flood
storage capacity was about three-fourths of that called for in the operating
policy; that is, the amount of water stored in the reservoir exceeded the
permissible storage (Figure 5. 1 1 ). Outflows reached 3.68 x 1 03 m3 S-I ,
some 13% above the maximum controlled release. Fortunately, disastrous
nooding was avoided because the storm abated and because the levee
system downstream of Folsom Reservoir stood up to the "excess" releases
from the reservoir. The effect of the storage reservoir on the flood is clear
-the peak inflow discharge, for example, was 5.44 x 1 03mJ S- I , compared
3 3 l
10 the outflow peak of 3.68 x 1 0 m S- (Figure 5. 1 1 ).
I7 I
2.5>:10'
Actual storage
E 2.0>:10'
,
-
g
8, 1.5>:10'
� 1.0>:10"
----
� ---=-
-c -
Permissable storage
, , , , -�"""1,---f
12 14 16 18 20 22 24 26
6000
,
-
"
•
E 4000
.
J£ Inflow
o
-
•
�
'"
-
2000
\ '\Outllow
u
.
• "-
0
-
/
0 • • •
12 14 16 18 20 22 24 26
Oay of the month, February, 1986
Figure 5.11 Flows and storage in Folsom Reservoir for the storm of
February 1986.
Data from NRC ( I 995a).
finite difference equivalent (Equation 5 .3), serves as the starting point for
river routing. The difference is that rather than using a weir equation to
relate storagc and outflow, we need an cquation relating the outflow from
a river rcach to the volume of water stored in thc reach. A complcte
analysis of the relationship between storage volume and outflow in a river
reach is beyond the scope of this book. Inste<ld, we adopt a simpler
approach in which observed inllow and outflow hydrographs are used to
develop an empirical relationship between storage and outflow.
172
The Muskingum method of flood routing in rivers assumes that the
volume of waler stored in a river reach can be related to the inflow and
outflow for thc rcach by thc equation
v= K,[xI + (1- x)O], (5.6)
mean velocity through the reach, U. Note that a rapid flow through a long
reach could have the same K, as a slower flow through a shorter reach.
Substituting Equation 5 .6 into Equation 5 .3 gives
(5.7)
where
The values of the coefficients Co, CI , and C2 must be positive (for x = 0.2,
0.75M < K, < 2.0.11) and the sum Co + C1 + C2 = I . Like the reservoir
problem, once the initial outflow 01 and the inflow hydrograph are
specified, the calculations can be stepped through to obtain the outflow
hydrograph.
The flood routing proccdure presented above allows one to compute an
outflow hydrograph given a known inflow hydrograph. To illustrate the
method, the inflow hydrograph used in the reservoir routing problem
(Table 5 .2) is routed through a river reach using the Muskingum method.
The value of x is taken to be 0.2. Outflow hydrographs for K, = 6hr and
12hr are shown with the inflow hydrograph in Figure 5. 1 2. For a given
flow rate, larger values ofK, correspond to increasingly long reach lengths.
For example, if U = I m S- I , the reach length for K, = 6 hr is approximately
173
22 km; the reach length is 44 km for K, = 1 2 hr. Alternatively, for a single
value of reach length, say 40 km, the two values ofK, correspond to flow
rates of approximalcly 1 .9 and 0.9 m S-l, respectively. Because we are
disregarding lateral inflow to the reach, the choice of reach length for the
river routing calculation would depend on the distance between major
tributaries and the limitations on values of tJ.t and K, necessary to keep the
coefficients in Equation 5.7 positive.
The hydrographs pictured in Figure 5 . 1 2 show that the lag in the timing
of the outflow peak relative to peak inflow is approximately equal to the
travel time constant. As the travel time constant increases, the flood wave
aucnuation incrcases. For K, 12 hr, the peak outflow is about 70% of the
=
7
0
-
0;-
E S
-
�
'" 5
"
�
"
u 4
0
C
.-
o
o 0.5 1 1.5 2 2.5 3 3.5 4
Time since start of flood (days)
Figure 5.12 Outflow hydrographs calculated using the Muskingum
174
method for the same inflow hydrograph used in the reservoir routing
example (Figure 5 . 10). As the travel time constant Kt incrcascs, the time
lag and attenuation increase.
175
This is another way of saying that the probability of the house being
flooded is 1/50 in any year. This situation docs not guarantee that the
company will make money on each individual policy. The structure may
be flooded in the first year of the policy or it may be flooded three times in
the first 1 0 years. On the other hand, it may not be flooded at all in 75
years. The probability statement does not specify that floods will occur
exactly 50 years apart but only that one is expected to occur an average of
once in 50 years (scc also Section 2.2.4.)
Thc same notion of probability is needcd in the analysis of flood·
protection measures. At the beginning of this chapter, the problem of
flooding on the American River system at Sacramento was introduced.
Mention was made that Folsom Dam was originally thought to offer flood
protection from a I-in-SOO-year stonn. Analysis of the level of protection
in this case required: an estimate of the I-in-SOO-year inflow hydrograph
and a routing procedure for this inflow hydrograph through the reservoir.
The probability of a flood equaling or exceeding a given magnitude can
be approximated using discharge records from a gaging station. For
example, if there were 20 years of record for a certain station we might
intuitively expect that the largest recorded flood is an approximation to the
h20-year flood"-a flood that occurs, on average, once in 20 years. The
fonnal procedure is summarized in the following steps:
I . The highest discharges recorded in each year for the II years of record
are listed, similar to the case ofrainf:111 extremes discussed in Chapter
2.2.4. These peak discharges are called floods and the series of the
largest flood in each year is tenned thc allllual series.
2. These floods are ranked according to magnitude. The largest flood is
assigned rank I , the second largest rank 2, and so on.
3. An initial detennination of the flood statistics can be found by plotting
the logarithm of discharge for each flood in the annual series against the
fraction of floods greater than or equal to that flood; this fraction,
tem1ed the exceedance probability, is given by r/(II + I), where r is the
rank of the particular flood.
4. If the data conform to a lognonnal distribution (i.e., the logarithm of the
data are nonnally distributed), they will plot along a straight line if a
nonnal probability scale is used on the y-axis as in Figure 5 . 1 3 .
5. If a log-normal distribution does not adequately fit the data, another
distribution, such as a log Pearson type III extreme value distribution,
might provide a better fit to the data (see Haan, 2002.) By fitting a
suitable distribution to the data, the exceedance probability and rctum
176
period for a flood of any magnitude can be estimated.
6. The best fit to the data, whether lognormal or othenvise, defines the
exceedal1ce probabilities for floods of any given discharge. The retllm
period, the average span of time between any flood and one equaling or
exceeding it, is calculated as Trcruril = I/(exceedance probability).
Table 5A. Annual flood peaks for Powell River, Big Stone Gap, VA,
1945-1994
177
rln + 1,
Waler year, 0_" ma.imum
r, flood
Oct 1 to discharlle for In(O_,)
rank year'
Sep 30 the year, m] s-'
178
1980 93.73 <1.54 42 0.82
0.01
0.02 o
0.05
0.10
.-
.-
L>
� 0.25
5.
cu 0.50
u
o
1l
� 0.75
u
•
W
0.90
0.95
0.98
0.99
179
5.6 Concluding Remarks
Dams, such as Folsom Dam on the American River, offer important flood
protection to people and property located on floodplains. Evaluation of
protection measures requires the routing of hypothetical floods with
specified exceedance probabilities. The U.S. Anny Corps of Engineers
(USACE) is responsible for the analysis of flood-protection altematives
for the American River system and many others. The USACE derives the
"design hydrographs" for various exceedance probabilities by using a
flood-frequency analysis similar to that described in this chapter.
Technical evaluation of dams or alternative flood mitigation measures,
such as natural wetland restoration, demands that we know how to "route"
rainfall through thc appropriatc storage zonc, be it a rcscrvoir or a wctland.
A comprehensive analysis of these routing problems requires that all
hydrological processes that can affect streamflow be understood, including
the movement ofwater in the subsurface. Chapters 6, 7, and 8 explore flow
in the saturated and unsaturated ponions of the subsurface. Wc rcturn to
consider thc dynamics of watcr movemcnt in catchments in Chaptcr 10.
Dams servc many valuablc functions, including watcr storagc and flood
protection, but they are not without problems. A dam regulates the
downstream flow of water in such a way that the discharge is more
constant than it would be naturally. Damming of streams can alter water
tcmperaturc and strcam chcmistry and fragment thc strcam itsclf. While
regulation of flow in strcams and rivers is bencficial in many respects for
the human populations they affect, it may have far-reaching consequences
for other pans of the river and riparian ecosystems. For example,
fragmentation and reduced flow rates in dammed streams impede the
upstrcam spawning migrations of salmon. Thcy also affect the transpon of
sediment in the stream. One consequence of this is accumulation of fine
sediment in reservoirs and increased water clarity downstream of dams.
The latter effect is pronounced in the Colorado River downstream of the
Glen Canyon Dam. Where the water was once warm and muddy, it is now
clcar and cold. This has transfonncd this part ofthc Colorado Rivcr into an
excellent trout fishing stream, to the detriment of the native fish
populations. Stream regulation also affects the riparian zones. For
example, reduced overbank flooding allows time for less flood-tolerant
plants to become established and for greater stabilization of the stream
banks by vcgetation.
Analysis of the world's largest river systems (292 rivcrs) revealed that
over half ( 1 72) are strongly to moderately affected by fragmentation by
dams and flow regulation by reservoirs (Nilsson et al., 2005). The large
180
impacts of such fragmentation and regulation on river hydrology and
ecosystems have prompted efforts to understand the scope of the changes
and, in some cases, to mitigatc thc changes by introducing controlled
floods (periods of high release intended to mimic naturally occurring
floods) or even removing dams entirely. For example, the Elwha and
Glines Canyon Dams on the Elwha River in the Olympic Peninsula,
western Washington, were removed in 201 1-2013, representing the largest
dam removal proj ect in U.S. history. A primary goal for removing the
dams was habitat restoration, particularly spawning grounds for salmon
and trout that used to be abundant in the Elwha River, but had become
reduced in numbers or even locally extinct. Other goats included
increasing the supply of sediment to the coastal area ncar the mouth of the
Elwha River-sediment that had been trapped behind the dams-and
renewing cultural traditions of the Lower Elwha Klallam Tribe, whose
native home is along the Elwha River.
I8 1
be found by numerically solving the finite difference fonn of the
conservation of mass equation. {Section 5.4 }
• Reservoirs are effective for flood control bccausc they can store a large
volume of water, thereby producing a significant decrease in the
magnitude of the outflow from the reservoir relative to the inflow at peak
flood conditions. { Section 5.4. I }
{ Section 5.4.2 }
• Estimates of return periods for time spans longer than the existing record
or probabilities of floods exceeding the largest recorded value can be
obtained by fitting a suitable probability distribution to the data, which
can then be used to estimate the return period of discharges larger than
those found in the record. The farther the data are extrapolated (e.g., for
re!"Urn periods much longer than the record length), the larger the
potential error will be in the resulting values ofretum period or flood
magnitude. Errors also result if the data are not well fit by the chosen
probability distribution. While a lognormal distribution is relatively easy
to fit 10 an annual series of peak discharge values, an extreme value
distribution is often needed to provide accurate estimates of floods with
return periods much longer than the record. { Section 5.5 }
Problem I. If the flood used in the reservoir example delivered the same
182
volume of water in a shorter amount of lime (shorter duration with higher
peak discharge), as given by the inflow hydrograph in the table below,
how would the outflow hydrograph changc? Complcte the tablc below,
assuming the initial conditions and other reservoir parameters remain the
same.
Time t. 2V� _ 0 2V
/hJ+ 0
Step N '. 1. +1..., 1
'" 6t
...
(days)
.
A • C 0 ,
0000 10 10
0600 SO
1200 130
1800 110
2400 70
Problem 3. In Figure 5 . 1 4, a 40-year annual series (1950- 1989) of floods
on the Eel River, California, is plotted against the fraction [r/(1/ + 1)J of
floods with discharges greater than or equal to each value.
A. Fit a line through the data and detemline the return period of an 8000
1
m3s- flood.
183
0.Q1
0.02 o
0.05 o
�
.
-
0.10
.0
.-
.. 0.20
.0
o 0.30
�
Q. 0.40
., 0.50
u
c 0.60
..
" 0.70
� 0.80
u
><
W 0.90
0.95 o
0.98 o
0.99 ..'I . .. .. • ••• • •• •
1 0' 1 02 1 03 1 04 1 0'
Discharge, Q (m3 S-' )
Figure 5.14 Probability plot for Eel River peak annual discharge, 1 950--
1989.
Dunne, T., and L.B. Leopold. 1 978. Water ill environmental planning. San
Francisco: W. H. Freeman. Chapter 1 0, pp. 279-39J.
Haan, C.T. 2002. Statistical methods ill hydrology. New York: Wiley.
Mays, L.W. 20 I I . WaleI' resources engineering, 2nd ed. Danvers, MA:
John Wiley & Sons. Chapter 9, pp. 3 3 1-358.
Nilsson, c., C.A. Reidy, M. Dynesius, and C. Revenga. 2005.
Fragmentation and now regulation of the world's large river systems.
Science 308:405-408.
184
6 Groundwater Hydraulics
6. 1 Introduction
6.1 Introduction
185
agriculture in many areas is based on withdrawal of groundwater. For
example, Shah et al. (2000) estimate that 60% of irrigated grain production
in India depends on the usc ofgroundwatcr. Halfofthc world's mcga
cities (population of 1 0 million or more) are dependent on groundwater
(Giordano 2009). Even in areas where surface water is available,
groundwater often is a preferable source because of water temperature,
quality, or accessibility. For example, in Dhaka, a city of 1 0 million in
Bangladesh, groundwater is the predominant source of freshwater even
though the city is located in the delta region of the Ganges and
Brahmaputra rivers and is bordered by major rivers.
The study of groundwater is motivated partly by practical
considerations of water supply. In addition, an understanding of the
hydrological cycle for a catchment requires a quantitativc description of
how the groundwater reservoir functions. For example, recall that
streamflow (baseflow) is maintained in perennial streams between
precipitation events. We can hypothesize that this is in large part the result
of the discharge of groundwater into stream channels. In this chapter and
in Chapter 7 we discuss subsurface pathways through which precipitation
eventually may reach a surface water body. We retum to these ideas again
in Chapter 1 0, when we examine mechanisms of runoff generation.
A number of environmental issues involve groundwater, especially the
remediation of sites that have been contaminated by poorly controlled
dumping practices and the identification of and planning for sites to safely
dispose of hazardous wastes. An example is planning for the disposal of
radioactive wastes.
The Waste Isolation PilO( Plant (WIPP) in New Mexico, United States,
is an underground repository built for certain radioactive wastes that were
generated during the construction of nuclear weapons. These wastes,
including work gloves and laboratory glassware, have been stored in 55-
gallon drums at facilities of the Department of Energy. The wastes are not
high-level, but they contain isotopes that remain radioactive for very long
periods of time (tens of thousands of years). These wastes must be isolated
for millennia to ensure that they do not pose risks to human health. In the
U.S. these wastes are being buricd in a repository in a 600-m thick salt
formation (thc Salado Formation) some 700 m below the ground surface.
Rock salt is thought to be a good host rock for radioactive wastes because
it flows and over time will seal the wastes off from the environment. The
only path that would lead to the release of radioactivity to parts of the
cnvironmcnt wherc it might adverscly affect people or ceosystcms is
through dissolution of the waste by water and transport of the dissolved
constituents by groundwater. The Culebra dolomite, a regional aquifer,
186
overlies the Salado Formation. Assessment of the suitability of the WIPP
site for disposal of radioactive waste requires knowledge of how
groundwater flows in the aquifer above the repository. Several questions
might be raised. What causes or drives the movement of groundwater?
What physical characteristics of subsurface fluids and porous media
detemline the rate of fluid movement?
Because we are again interested in the flow of water in relation to
imposed forces, the equations of fluid mechanics provide the basis for the
quantitative description of groundwater flow. At thc scale of the porcs in
rocks and soils, however, the paths along which groundwatcr flows are
complex, with many twists, turns, contractions, and expansions. So we
recognize immediately that simplifications will have to be made to enable
useful equations to be derivcd.
Let us try to picture the flow of water in a porous medium, for example,
likc sand. Thc path a "parccl" of watcr might follow in moving through a
material containing pores or void spaces is convoluted (Figure 6. I ). Not
only does the water follow tortuous paths, but the geometry of the
channels of now is extremely complex and cannot be specified completely
(i.e., the position, size, and shape of all of the sand grains cannot be
known). Finally, we recognize thai the opcnings in which water flows arc
very small. Therefore, we might expect that frictionless flow is totally
meaningless in this situation and that head losses will play the
predominant role.
187
Figure 6.1 Schematic of a thin section of a porous medium and the
tortuous flow path of two water "parcels." (Note that a real medium is
three-dimensional with flow through the open spaces within the three
dimensional matrix.)
188
flow in the capillary tube is laminar. Moreover, because the velocities are
very small, we can neglect the effect of changes in velocity and velocity
head through the capillary tube, and assume that the diameter is constant.
In Section 3.7. 1 , we found an expression for the average velocity,
_!!....(..E.... + _]
medium.
u D2pg
(3.38)
-
-
dl pg - 32jJ
(6.1)
(6.2)
189
Equation 6.2 is a form of Poiseuille's law for the flow of a viscous fluid
through a capillary tube. Discharge is directly proportional to the hydraulic
head gradient (or hydraulic gradient), inversely proportional to the fluid
viscosity, and directly proportional to the fourth power of the radius of the
tube.
The equations derived from this conceptual model have several
important implications. First, Equation 6.2 indicates that for a given fluid
and given hydraulic gradient (dhldl), the discharge varies as the fourth
power of the radius of the tube. For example, the discharge from a tube of
radius 1 0 mm will be 104 times greater than that from a tube of radius 1
mm, if all other conditions are the same. Because the size of the capillary
tubes in our conceptual model is related (on an intuitive basis) to the
texture or grain size of the porous medium, flow rates will depend on the
texture. In addition, Equation 3.3 8 specifies that the average velocity, V, is
proportional to the hydraulic gradient, dhldl,
U
=
_ D2 pg!!.!!. (6.3)
32 Jl dl
Equation 6.3 is the fonn ofPoiseuille's law that provides the most useful
analogy for flow through a porous medium. As we will see, an equation
almost identical to Equation 6.3 is the basis for studies of groundwater
flow. Because Equation 6.3 was derived in Chapter 3 using well-known
principles and stated assumptions, we can usc the analogy between the
conceptual model and the actual porous medium to great advantage in the
sense that we can apply, in qualitative lenns at least, knowledge about the
factors that control laminar flow in tubes directly to the flow of
groundwater.
(Figure 6.3). The sum of elevation head (z) and pressure head (plpg) is
190
represented as II and varies from hi to h2 along the column. The hydraulic
head at each end of the column is measured with an open tube, a simple
manometer, as shown in the diagram. By varying L and the hydraulic head
difference across the column (hi - h2 = I1h), Darcy found that the total
discharge Q varies in direct proportion to A and to I1h and inversely with
L . That is,
(6A)
== _ K dh
q dl .
(6.6)
191
•
t,},
I
Cross-sectional
area, A
z,
Datum "
- = 0
Figure 6.3 Schematic diagram of an apparatus for illustrating Darcy's law .
• 92
200 .-------�
-
-
.'"
-
E
150
E
E
-
"
� 100
..
.t:
U S lope = K
'"
.-
"
U
;;:
. 50
u
-
1i
'"
5 10 15
Hydraulic gradient, -dhldi
Figure 6.4 Darcy's ( 1 856) original data showing a linear relationship
between specific discharge and hydraulic gradient for two different sands.
We can envision the hydraulic conductivity as the slope ofa line relating
specific discharge, q, with hydraulic gradient, dhldJ (Equation 6.6) .
Imagine a sel of experiments using a given sample of material and a fluid
of constant density and viscosity. By varying the hydraulic gradient and
measuring the discharge, q can be plotted against dhldl. According to
Darcy's law a straight line should be the result, the slope of which will be
K (Figure 6.4). Using a fluid with a constant density and viscosity, the
slope of the relationship between q and dh/dl will depend only on the
material and generally will increase with the coarseness of the material.
Now imagine repeating the experiments with another fluid having different
properties (e.g., a greater viscosity). We would expect the more viscous
fluid to move more slowly if everything else remains constant.
The example above suggests that hydraulic conductivity depends on the
nature of both the fluid and the porous material. The way in which K
depends on these properties can be inferred by reference to the conceptual
193
model discussed earlier. We already have remarked on the similarity
between Equations 6.3 and 6.6, which are reproduced below (for one
dircctional flow in the I dircction) to facilitate dircct comparison.
dll .
q : _K ( Oarcy's law) (6.6)
dl
This analogy suggests that pglJI should describe the variation of K with
fluid density and viscosity, and that D2/32 should describe the variation of
K with pore diameter. Of course, the "diameter" of the pores is not
measurable (or even well defined), so the analogy is not perfect. However,
qualitatively we might expect the grain size of the material to give an
indication of the size of the openings (i.e., we would expect larger pores in
a boulder field than in a silt deposit) and we can define and measure the
average grain size of a granular material such as sand. Experimental
evidence supports this analogy, at least for simple granular porous media.
Based on experimenlal results, the following empirical relationship for
hydraulic conductivity can be written:
(6.7)
2];/1 = the viscosity of the fluid 1M L- 1 t1]. From this we see that K is
composed of two factors, one represcnling fluid properties and thc other
representing properties of the medium. Darcy's law can be written in a
manner that clearly separates these two influences:
(94
values of K or k resulting in faster transport. Measuring or estimating these
properties is a fundamental step in applying Darcy's law to a natural
setting. There arc a variety of techniques and methods for either directly or
indirectly detennining the penneability of a sample of porous material. A
small sample can be placed in a device called a permeametcr, not unlike
Darcy's original column. The flow rate through the sample can be
measured for a known hydraulic gradient and the permeability can be
calculated directly using Darcy's law. [n some situations, for example in
looking at flow deep underground, samples may be difficult to obtain.
There are a variety of methods, known as aquifer tests or "pump" tests, for
detemlining penneability in such cases. By withdrawing or adding water
to a well, and measuring the water level in that well or other wells nearby
as a function of time, we can ealcu[ate the penneability. Finally, there are
indirect methods that are based on measuring some other parameter, such
as grain size, that is related to the permeability.
Literally thousands ofpenneability measurements have been made in
different materials. The results show that the range of permeability of
natural materials is quite large (Figure 6.5). Recalling the discussion of the
WIPP site in the introduction, one reported range of hydraulic
conductivities for salt deposits, such as the Salado Formation, is 10-12 to
1 0-10 m S-1 (Domenico and Schwartz, 1 990). Relative to the other types of
materials shown in Figure 6.5, this range is near the low end and is similar
to shales and unfractured crystalline rocks. The resulting slow rate of
groundwatcr flow is one rcason that salt fonnations have bcen considercd
for waste repositories. For example, a typical hydraulic gradient of 1/100
in a salt fonnation with a hydraulic conductivity of 10-10 m S-I will
produce a specific discharge of 10-12 III S-I, or less than I mm per 30
years !
Darcy's law indicates that for a given value of the hydraulic gradient
(dhldf), the specific discharge will be greater for a penneable material,
such as a sand or gravel, than for a granite. The difference can be several
orders of magnitude (Figure 6.5) . As mentioned earlier, the sped fic
discharge, despite having dimensions of velocity [L JI], is not a velocity.
We could not use specific discharge to detennine how long it will take a
parcel of water to move from one point to another. The cross-sectional
area available to the water is smaller than the actual cross-sectional area,
such that the solid portion of the porous medium acts as a constriction.
This constriction means that a tagged parcel of water, or, better, many
tagged parcels that are averaged together, will appear to move through a
porous medium at a speed that is faster than the specific discharge. The
195
effect is similar to the constriction in a pipe discussed in Chapter 3; the
constricted Ilow has a greater mean velocity for the same value of
dischargc.
Intrinsic permeability, Ii.
, 0-" 10-19 H)-17 H)-15 10-13
, , I t !
Gh&1 till
'"'"
I i i i i i i
10-1' 10-'2 10-'0 10-11 10-11 1 0.... 10-2
Hydraulic conductivity, K (m s-t)
Figure 6.5 Ranges of intrinsic permeability and hydraulic conductivity for
a variety of rocks (gray bars) and sediments (blue bars).
Data from Freeze and Cherry (1979).
We can detcrminc thc mcan pore water velocity ifwc havc some idea of
the amount or "constriction." For porous materials, this property is the
porosity [dimensionless], which is simply the fraction ofa porous material
that is void space:
v,. .
¢= (6.9)
V,
where V" is the volume of void space (L3] and VI is the total volume [L3].
lfwe have measured the porosity, which by definition must be between 0
and 1 , we can detennine the mean pore water velocity, or average linear
velocity, v [L I I]:
- q
I' =-. (6.10)
�
196
Darcy's law is widely used for almost all situations involving motion of a
fluid through soil or rock in the natural environment. Although the
conceptual modcl we used to introduce our discussion of Darcy's law can
apply (to a limited extent) only to granular materials such as sand,
Equation 6.6 is applied in materials ranging from clay to limestone to
fractured crystalline and metamorphic rocks. The openings through which
fluid flows in most of these materials cannot be envisioned as capillary
tubes. Nevertheless, Darcy's law usually can be applied with success.
There arc some general limitations on the usc of Equation 6.6, and
because of the analogy we developed we can infer how these restrictions
arise by examining the assumptions implicit in the derivation of Equation
6.3. First, Darcy's law has been found to be invalid for high values of
Reynolds number. Experiments using very high flow rates in very
penneable materials have found that when values of the Reynolds number
exceed the range 1 - 10, Darcy's law does not describe the relationship
between flow rate and hydraulic gradient. For flow through porous
materials, the characteristic length in the Reynolds number is the mean
grain diameter, d:
(6. 1 1 )
197
mineral structure and at low hydraulic gradients these may contribute to a
force balance.
198
profile, saturated conditions prevail (saturated zone); groundwater by
definition refers to water in the saturated zone of the subsurface. The
water table or phreatic surface is dcfined as a surfacc ofzcro gage
pressure within Ihe subsurface, and separates the saturated and unsaturated
zones. Water will flow inlo an excavation or well up to this level; the
water table is equivalent to a free surface. An aquifer with the water table
as thc bounding surface of the lOp of the aquifer is an unconfined aquifer.
The second type of aquifer is called confined or artesian. This type of
aquifer is found when permeable material (the aquifer) is overlain by
relatively impermeable material (an aquiclude; Figure 6.6). The water in a
confined aquifer is under pressure and, in a well penetrating the aquifer,
will rise above the top of the aquifer (see Figure 6.6). The height to which
water rises i n a well defines the piezometric or potentiometric surface.
Note that a well penetrating a confined aquifer can be thought of as a
piezometer, or single tube manometer. The water level in a piezometer is
a measure of waler pressure in the aquifer. Thus, the elevation of the
potentiometric surface above an arbitrary (horizontal) datum is the sum of
pressure head and elevation head, or the hydraulic head, h, of Darcy's law.
Potentiometric surface
-
-
____
_
__
l in confined aquifer
'.
- -
'.
' .
or water-table
Confined aquifer
199
the piezometer (jar right) indicates the height of the water table; in
confined aquifers, the water level in the piezometers (left and cell1er) rises
abovc thc top ofthc aquifcr and indicatcs thc position ofthc potcntiomctric
surface.
200
concrete cap is generally installed to complctc thc scal ofthc casing and
prevent surface water and contaminants from entering the well (Figure
6.7b).
Kcy mcasuremcnts that wc derivc from wells and piezomcters arc thc
elevations of the potentiometric surface or the water table. It is these
measurements that allow us to establish head gradients that can be used
with estimates of hydraulic conductivity to calculate groundwater flow
rates. Once the elevation of the top of the well casing is established by
survcying or the use of satellite positioning, we can dctermine thc
elevation of the water surface in a well by measuring the distance from the
top of the well casing. One instrument that we routinely use to make these
measurements is a pressure transducer. This is a device, as the name
suggests, that measures water pressure. The transducer is lowered a known
distance from the top of the well casing and below the watcr level in the
well and the pressure is measured. The hydrostatic Equation 3.5 can be
used to compute the pressure at the depth of the transducer. Pressure
transducers can be left in place for extended periods and connected to a
dalR logger to record well hydrographs (Figure 6.8).
""-
• Clay seal
• Gravel pack
Screen
201
6.4.3 Geophysical techniques
There are other, indirect techniques that are used to estimate subsurface
conditions that cannot be "seen" in the conventional sense of the word.
Geophysics is a field that uses measurements of physical processes and
properties to infer information about the structure of the Earth, and in the
case of hydrogeology, the relatively shallow solid Earth in particular. That
is, we usc measurements involving the propagation of sound (seismic
vibrations), electricity, or radar, for example, to infer the distribution of
Earth properties that affect transmission. These include density for seismic
vibrations, electrical resistivity for electricity, and the dielectric constant
for radio waves (radar). These physical properties often can be correlated
with hydraulic properties because the differences in rock type and structure
that affect the physical properties also affect the hydraulic properties such
as porosity and hydraulic conductivity. In a sense, use of geophysical
techniques can be considered to be a way of "seeing into the Earth" (NRC,
2000). The array of quite sophisticated techniques used in geophysical
investigations is extensive and a description of methods is beyond the
scope of this book. Suffice it to say that lhe visualization of subsurface
structure can provide significant information for understanding
hydrogeology. Even a subtle underground feature may have important
implications for the flow of water in soils and rocks (Figure 6.9).
202
2
�
�
Measured in feel from ground surface
0
=;
"
; j>-2
" ..
,,
o -4
-
C!J
� --------�--�-
Feb Mar Apr May
Time. 2012
Figure 6.8 Well hydrograph for USGS Well 351428085003600, Hamilton
County, Tennessee, USA.
Provisional data from USGS.
, dl1
q �-I\. - . (6.12)
d,
We can rearrange this expression and integrate over the length of the
confined aquifer to obtain an expression for hydraulic head, II:
203
"I"" .. I.... "I'" HI""
.'."..., .'&
�� OO� E.� ��
." '- . ....,
•
__ Potentiometric surface
� •
in confined aquifer
•
•
-
I Confined
h, � b h,
Flow
aquifer
'-'
x=o x= L
Datum, ::. = 0
Figure 6.10 Horizontal flow in a confined aquifer.
204
q
dh::--dr
K
(6.13)
Equation 6.13 specifics that head decreases linearly with distance from the
left boundary; that is, the potentiometric surface is a plane sloping from
left to right. This is shown as a solid gray line in Figure 6.10. Water flows
in the downhill direction of this surface.
On a plan view of the channel-aquifer system considered above, the
contour lines of the potentiometric surface arc parallel to the channels
(Figure 6.11). The spacing of the contours indicates the slope oflhis
surface, just as do the contour lines of a standard topographic map. Once
these lines of equal hydraulic head, or equipotentials, have been
established, lines that indicate the direction of flow can be sketched in by
constructing perpendiculars to the cquipotemials because this reprcsents
the downhill direction of the potcntiomctric surfacc, the direction of flow
specified by Darcy's law. These lines are called streamlines. Together, the
equipotentials and the streamlines constitute a flow net. Flow nets can be
applicd to great advantagc in actual field problems whcre groundwater
flow patterns are to be cstablished based on the measured watcr levcls in a
series of wells (see Fetter, 2000, for examples).
We will discuss the use of flow nets in greater detail in a later section of
this chapter, but at this point we need to consider the physical reasoning
behind the idea that streamlines must be perpendicular to equipotentials.
Darcy's law indicates that flow should always be from high valucs of thc
hydraulic head to low values. In fact, flow must follow the path ofsteepest
descent. Picture the potentiometric surface or the water table as a
topographic surface, with hills (high II) and valleys (low II). Now imagine
a drop of water moving along this surface in response to the pull of
gravity. We havc to imagine this watcr drop moving rather slowly, without
the momentum we might expect a ball bearing or other object to have on a
steep surface. The water drop moves down the slope and perpendicular to
the "topographic" contours rather than following some arbitrary path down
the slope; that is, it follows the path of steepest descent. The rate of
desccnt is proponional to the slope, just as the specific discharge is
205
is not horizontal (as one might expect intuitively) but has a significant
upward component Groundwater flow is not concentrated near the water
table with a large volume of stagnam water at depth-the flow patterns are
actually such that flow occurs throughout the saturated zone. The fact that
water flows along "U-shaped" paths is of practical and theoretical
importance. The U.S. Geological Survey conducted a study on the
presence and the movement of agricultural chemicals in shallow
groundwater on the Delmarva (Delaware-Maryland-Virginia) Peninsula
(Hamilton and Shedlock, 1 992). They found nitrate (partly from applied
fertilizers) at almost all depths sampled in the groundwater and attributed
the patterns of contamination to land-use practices and to groundwater
flow paths.
The construction of accurate flow nets by hand is a task requiring
considerable practice. More commonly today, flow nets arc constructed
using numerical solutions to the equations governing groundwater flow.
Regardless of how a flow net is constructed, it provides not only a
visualization of the groundwater flow paths, but also infon11ation on the
rate of groundwaler flow in a particular region.
net I a
207
hillslope.
Redrawn from Hubben, 1940, fig. 45.
208
detennine the total discharge through this box:
Equipotentials
;v �f
'. 't
-j-
· ..
·
, ,
•
..
•
,
•
.
.
·
• •
·
,
·
,
.
.
-+ · ,· ,
, tis
•
.' .
..
. ·
,, ,
'
,
· ·
·
,
.
.. ..
tim
" "
·
·
. .
·
..,.
.
•
.
",
·
·
',
, .
71· ..
S ream-
t
lines
,
·
"< >
•
.,
, • · ·
.
'.
•
•
•
· •
· . ·
' • ,
·
·.
..
. · · ··,
·
,
·
"
·
..
•
·
I"
•
' ,
I·. ·
·
· ' •
· ·
..
".
.
•
..
.. ,
·
· .
. .. ,
·
· ·
·
·
· ,
•
•
I I
·
where Q, refers to the total discharge [L3r-1J through a streamtube and dll
refers to the head difference across the box, or in this case "4 113, Because -
209
transmissivity, T= Kb), we can simply look at our flow net to see what
contour interval is used for hydraulic head (dh) and multiply that value by
thc transmissivity to detcnninc thc amount of water moving through cach
streamtube. If the aquifer is bounded at upper and lower ends, then we can
count the number of streamtubes and multiply by Qs to detennine the total
amount of water flowing through the aquifer.
We now look at a somewhat more complicated example. One effect of
placing a dam in a stream or river is that a hydraulic gradient is created
beneath the dam. At the upstream end, the hydraulic head (relative to the
basc of the dam) at thc bottom of the reservoir is equal to the depth of
water in the reservoir (Figure 6.14). On the downstream side, the hydraulic
head is equal to the height of water in the river. A flow net can be
constructed, as shown in Figure 6.1 4, to detennine the panem and rate of
steady groundwater flow beneath the dam. Note that a low-penneability
laycr exists at depth, which rcpresents the bottom of the aquifer beneath
the dam. We can envision this boundary, as well as the base of the dam
itself, as streamlines, because there will be relatively little flow across
them. For this example, we assume that the dam is 100 m wide (in the
direction into the page), and that the hydraulic conductivity of the material
beneath the dam is 10-10 m S-I. We can use Equation 6.17 to calculate the
total discharge beneath the dam. We use the length of the dam (100 m) in
place of the aquifer thickness (b). The contour interval for hydraulic head,
dh, is 2 m. Then,
We can multiply this value by the number of streamtubes (3) to obtain the
total discharge, 6 x 10-8 mJ S- I, which is approximately equal to 1.9 mJ yr
I. In this case, the flow beneath the dam is fairly small, which is what we
would hope for; the dam would be fairly inefficient if water was constantly
leaking around it! A larger value would result if the hydraulic conductivity
was higher, or the differcncc in head across the dam was greater.
Before we conclude this chapter with a discussion of some of the
complexities that enter into our evaluation of groundwater flow, we
mention several other points regarding the use of flow nets as we have
described it. First, we have considered only steady groulldwaterjlow. It is
not possible to construct a flow net for unsteady or transient groundwater
flow. Also, we can usc a flow net to describe thc variation in specific
discharge, and, therefore, flow velocity, in an aquifer. You can imagine
each streamtube as a "pipe," because water cannot cross a streamline.
Therefore, from the principle of conservation of mass described in
210
Chapters 3 and 4, we can state that within a streamtube, the specific
discharge will be greatest where the streamtube is narrowest. The total
dischargc through the streamtube must be the same at any cross section.
Referring to Figure 6.14, the flow velocity will be greater at point A than
at point B. Finally, we have considered only homogeneous porous media,
that is, materials with a spatially constant k. In the next section, we discuss
the implicat'ions for flow through materials thaI are not homogeneous.
-
It= 26m
11=10 m
Concrete dam -
>- � 12
B , \
, '4
•
16
20
24 22
I
Figure 6.14 Flow net for groundwater movement beneath a concrete dam.
The equipotentials (gray lilies) are labeled with values of hydraulic head.
21 1
the pattern and rate of groundwater flow.
Another complication often arises when measuring the penneability of
natural materials. Rocks, sediments, and soils often have textural features
that cause the penneability at a point to vat)' with the direction of
measurement. Materials that display this trait are referred to as
anisotropic, whereas isotropic refers to the condition in which the
penneability does not depend on the direction of measurement. Consider a
fractured rock aquifer, in which the fractures are predominantly horizontal.
In this case the penneability will be higher when measured in the
horizontal direction than in the vertical direction. You might picture this
situation by referring to Figure 6.2 and imagining that the capillary tubes
are aligned in the horizontal direction with relatively little hydraulic
communication between them. Other features that produce anisotropy
include (but are not limited to) the orientation of: platy minerals and small
scale layering in sedimentary rocks, such as clays; cooling cracks and lava
flow tubes in basalis; large pores due to animal burrowing and plant roots
in soils; and schistosity and fractures in metamorphic and igneous rocks.
Anisotropy, like heterogeneity, makes the job of constructing flow nets
difficult, so that once again we must rely on numerical models that are
capable of including this aspect of natural porous media. In the case of
anisotropy, because water will tend to flow in a "preferred" direction, that
is, in the direction of maximum penneability, Darcy's law must be
modified to include this preference. As a result, in an anisotropic medium,
the streamlines and equipotentials may not be perpendicular to one another
at all points.
In introducing the concept of a flow net in Section 6.5, we described
streamlines as paths of steepest descent in the "topographic landscape" of
hydraulic head. We can extend this analogy to anisotropic media if we
now imagine a series of ridges running at some angle with respect to the
equipotentials. These ridges cause the water drop to move at some angle
other than straight "downhill," since the water wants to follow the ruts
between the ridges at least part of the time. (The flow will be straight
downhill if the ridges are either straight downhill or follow the
equipotentials.) These ridges in the hydraulic "topography" have the same
effect as anisotropy in porous media. The flow direction will be altered
from the nonnal direction parallel to the hydraulic gradient toward the
direction of maximum intrinsic penneability.
212
Darcy's law (and the law of conservation of mass implied in the
construction and use of flow nets) provides the basis for computing steady
groundwater flow patterns and rates. In some cases a simplc "back-of-the
envelope" calculation is sufficient to gain a rough estimate of flow rates.
In other cases, the aquifer may have a complex geometry or be
heterogeneous and anisotropic and simple calculations will not suffice.
Today hydrologists routinely use groundwater models (by which the
groundwater flow equations are solved with the assistance of computers)
to address environmental issues (e.g., Konikow et aI., 2006). The
computational techniques used in groundwater models are much more
sophisticated than those we have used in this chapter. However, the ideas
behind the computation are essentially the same. The geometry, conditions
at the boundaries, and hydraulic parameters (intrinsic penneability,
porosity) of the aquifer must be specified before predictions can be made
with the models.
One use of groundwater models is in making assessments of
compliance with environmental regulations. For example, before the WIPP
site was licensed to begin operation, a set of regulations issued by the U.S.
Environmental Protection Agency (EPA) had to be met. For WIPP, the
U.S. Department of Energy had to show that the probability of significant
releases to the accessible environment over 10,000 years into the future
will be very small. The "accessible environment" for WIPP means
groundwater in the Culehra fonnal'ion "down gradient" of WIPP, that is, in
the direction of decreasing hydraulic head. The demonstration of
compliance with the EPA standard is done using a perfonnance assessment
analysis (Helton et aI., 1997). In short, performance assessment uses a set
of scenarios-sequences of hypothetical events that might occur in the
future-and models to predict the impact of these scenarios. For WLPP,
one scenario that is of concern involves someone in the distant future
drilling a well into the repository and allowing some of the radioactive
wastes to flow up the well and into the Culebra dolomite. A groundwater
model must be used to "route" the hypothetical contaminant through the
aquifer to decide whether significant quantities of contaminant might reach
a human population in this scenario. These and many other environmental
problems require the kind of knowledge of groundwater hydraulics that we
have introduced in this chapter.
213
saturated porous media. {Section 6. 1 }
• Darcy's law may be used for many situations of flow through porous
media, except when values of the Reynolds number are high or other
forces (accelerations, electrostatic forces) are significant. {Section 6.3.2}
214
{Section 6.4}
• Aquifers may be either unconfined (bound at the top by the water table)
or conlined (bound at the top by an overlying aquitard). {Section 6.4}
Problem 1. You are charged with designing a very simple filtration system
for a community water supply, using cylindrical sand columns (K = 5.0 m
day-I ). The filter needs to be 3.0 m long to adequately trap particulates in
the water, and since the system will be driven by gravity, the pressure
heads at the top and bottom of the (venically-oriented) filter will be zero.
B. Consider each of the alternatives and how you might modify your
design:
215
i. Lengthen the sand filter (how long?)
ii. Raise the hydraulic head at the inflow (how high?)
iii. Use several filters (how many? what size?)
A. Which way is water flowing in the eolumn? [s water flowing from high
to low hydraulic head? From high to low pressure?
Piezometer n Piezometer #2
Pressure, p (Pa)
216
Problcm 4. Consider the flow net for a drainage problem shown in Figure
6. 1 5. Drains such as pipes and culverts placed in a wet field may be used
to remove groundwater by creating a "sink" or area of low hydraulic head.
In the figure, a cross section through such a field is shown. The hydraulic
conductivity, K, of the surficial material is 1 . 0 x 10-5 m S-I. The thick
black lines represent impermeable boundaries; a constant head is assigned
to the top and lower left side. The cross section is 20 m long by 10m deep.
The gray lines are cquipotentials and the blue lines arc streamlines.
,,= 1 m
, . ' ' .' .
' ..
.. :. ' . " C,
' '
.. . .
. .
.'. .
.
..
'
.
. :' . .
.. .
.
. . ' . . , . . . '
. .
: :
.
'. . .
.
. ,
."
. . ., .
. . .
. '
"
. " , ':. :: .. . :
:. .... . : . : ." .';:� ::' . : '.': .<.:.:- .
'
. , : :
. "
.
. . . .
. '
' .
. .
.... . . . .
..
. ' .
. '
. . . . . .. .
.. , . . . : .
.
•:
..,
. .
. . .
.
. .
. .
.' ' ',
.
.
"
:..
. . . . . • . . . . . . .
.
'. '
.
. .
.
. .
.
" . .
.
.
.
. .
.
. .
. .
. . '
. .
. . .: .
.
.
,
.
'
.
, ':'.» :::';:.'
"
'
, , ,
. . :'.
" "
,
:
" "
: : < : .: . .
' .. . ... .. .'
:
.
. . .
.
. .
.
.
'
.
. .
'
" . .
. .' .
, . ' - ..
.. . ' . . .
.
'
: " ,
: .
. . .
B. Calculate the discharge through each streamtube, and the total discharge
or rate at which the field is drained (m3 day-l per width of material).
Fetter C.W. 2000. Applied hydrogeology, 4th ed. Upper Saddle River, NJ:
Prentice Hall, Chapters 3 and 4.
217
7 Groundwater Hydrology
7. 1 Introduction
7. 1 0 Key Points
7. 1 1 Example Problems
7. 1 2 Suggested Readings
7.1 Introduction
218
the United States, for example, more than 40% of the water used by people
in Orange County south of Los Angeles comes from groundwater. Because
precipitation and surface runoff vary with thc seasons, Orange County uses
spreading basins (gravel-lined areas) over which they spread water so it
can infiltrate to recharge the groundwater that is in high demand. During
most years there is not enough natural flow in the rivers to accommodate
the demand for recharge. In those years, reclaimed wastewater may be
used for recharge. For example, in the city of Anaheim the Santa Ana
River is diverted into spreading basins to allow the water to infiltrate and
recharge the underlying aquifer. In the summer over 90% of the flow in the
river is treated wastewater that is returned to the river upstream of the
spreading basins (NRC, 2012c). Trade-ofTs between amount and quality of
water are inevitable and careful plans must be made to sustain the
resource.
One example of a difficult problem in the use and protection of
groundwater resources can be found in Mexico City, which is now one of
the world's "megacities." Water must be supplied-and wastewater treated
-for about 20 million pcoplc. Usable surface water is scarce in thc
Mexico City Basin, so groundwater from the Mexico City Aquifer is a
primary source of freshwater. Near Mexico City, the aquifer is a sequence
of alluvial fill sediments interstratified with basalt deposits and overlain by
clays. The principal aquifer is from 100 to 500 m thick and the overlying
clay aquitard is approximately 100 m thick. The water table bclow Mcxico
City has been declining at a rate of I to 1.5 m per year; the rate at which
water is being withdrawn by pumping wells is much greater than the
natural rate of replenishment. The practice of withdrawing groundwater
faster than it can be replenished, referred to as "overdraft," has been
common since about 1900. Furthennore, poor waste-disposal practices
have adversely affected water quality in the aquifer. Clearly, if the water
resource for Mexico City is to be made sustainable, changes in the pattern
of use based on an understanding of the basin hydrology will be required.
In the United States, the High Plains Aquifer (often referred to as the
Ogallala Aquifer) has experienced similar problems (McGuire, 2011). The
High Plains Aquifer is an important source of water for much of the
central United States, including parts of Colorado, Kansas, Nebraska, New
Mexico, Oklahoma, South Dakota, Texas, and Wyoming. About 20
percent of the irrigated land in the United States is i n the High Plains, and
about 30 percent of the groundwatcr uscd for irrigation eomcs from thc
High Plains Aquifer. Overdrafts of groundwater have resulted in a decline
in water levels throughout much of the aquifer. Between 1950 and 2009,
the decline was more than 4 m (on average) and exceeded 45 m in some
219
places. Declining water levels (which are expected to continue into the
future) increase the cost of water, because pumps require more energy to
lift the water a greater distance.
In places such as the High Plains, Mexico City, northern India, the
North China Plain, and the southeast of Spain, groundwater is a limited
resource and is being depleted (Wada et al., 20 I 0). In other words, the rate
of groundwater withdrawal exceeds the rate at which it is naturally
replenished. As a result there is a net reduction in groundwater storage. In
some cases, modem societies are using groundwaler that accumulated in
aquifers over geological time scales and under different climate conditions.
This unsustainable use of non-renewable groundwater resources is
sometimes called groundwater mining. The water removed from the
ground is ultimately released to the ocean, and contributes-in addition to
the effects of climate change-to sea-level rise (Wada et aI., 2010;
Konikow, 201 1). Global estimates of groundwater depletion vary between
27 km3 yr-I (Margat et al., 2006) and 283 km3 yr-1 (Wada et al., 2010).
Table 7.1 shows an intennediate estimate (145 km3 yr-I) along with the
major hotspots of groundwater depletion in the world. As a tenn of
comparison, globally, the irrigation water used by crops is about 545 km3
yr- 1 (Siebert et aI., 2010).
To understand limitations on the sustainable use of groundwater, we
must examine the water balance for the subsurface. How and at what rate
is groundwater replenished or recharged? Hydrological basins (the rocks,
sediments, and soil underlying catchments) exhibit natural rates of
groundwater flow as a result of the infiltration of precipital'ion, exfiltralion
or discharge of groundwater, and evapotranspiration. Flow nets provide a
tool for evaluating the natural rate of groundwater flow. How is that flow
altered by perturbations such as pumping? Groundwater pumping in the
Mexico City Basin has changed the panern of flow as well as the water
balance. Because the flow pattern detennines how pollutants move within
the subsurface, flow nets are also important in evaluating the risk of
contamination.
220
Net depletion in Averase annual
Aquifer
2001-200S{kml) rate (kml yr")
In the USA
Flow nets can be constructed for any selling where the approximations of
steady and two-dimensional flow are valid, as long as the conditions at the
boundaries and the distribution of hydraulic conductivity are known. Later
in this chapter we explore cases where flow is not steady. For the present
discussion, we consider steady flow to be an approximation to the real
system, such that temporal variations in the height of the water table are
small relative to the thickness of the flow system.
221
I I
I I
I I
I I •
I I
I I
I I
I I
I I
Q;n I I QOfll
II dx I
I
I I
I I
I I
hi h2
I I
I I II
I I
I I
I I
Figure 7.1 Definition sketch for the derivation of the steady groundwater
flow equation.
QOut - Q", = dQ =
0
d\· dr
,
where Q [L3r-l] refers to the total discharge at any section through the
streamtube. For uniform now in the x-direction (both streamlines are
horizontal), the specific discharge, q�, also must be a constant:
dq,
=0.
dx
(7.1)
222
dh
q� dx '
K (7.2)
_
_
where dh is equal to 112 " 1 ' Substituting Equation 7.2 into Equation 7. 1
-
and assuming that K does not val)' in space (i.e., the material is
homogeneous):
dq,
= K.!!... :!!!. =
0
dt dt
_
dt
or
d2h
=0. (7.3)
d,'
-
(7.4)
223
bottom (for example, the bedrock beneath an unconfined aquifer) and two
impermeable sides, called groundwater divides. The upper boundary is
the water table. Often, the water table is a subdued replica of the land
surface topography, such that the water table is higher beneath hilltops
than beneath a valley. A flow net shows a balance between recharge in the
upland portion and discharge in the valley (Figure 7.2a).
Flow nets must obey certain rules at the boundaries. For example, it is
probably apparent that streamlines cannot cross impcnneable, or "no
flow," boundaries. Equipotentials must be perpendicular to impenneable
boundaries, because otherwise some flow across the boundary would be
implied. A constant-head boundary is represented by an equipotential, and
streamlines must be perpendicular to a constant-head boundary. At the
water table, the hydraulic head is everywhere equal to the elevation,
because the gage pressure is zero. Equipotentials and streamlines may both
intersect a non-horizontal water table, and the orientation of these lines
provides a means of delineating recharge and discharge areas (Figure
7.2b).
Discharge area -1
,
,
,
-·----- Recharge area ------ :
,
Recharge area
(b)
Discharge area � UphiU
Figure 7.2 The natural pattern of groundwater flow in a simple basin (a).
Equipotentials are gray and streamlines are blue. Recharge and discharge
areas may be distinguished by looking at the angle that an equipotential
224
makes with the water table (b).
Piezometer 60
'----- 55 50
50 ---1
30
45
-1
20
__
40 _
__ _
_
Open portion 10
of piezometer
o
Figure 7.3 The hydraulic head along any equipotential is equal to the
elevation of its intersection with the water table.
Qs = Kbdh. (6.17)
Equation 6.17 is valid as long as the flow net has been constructed
using curvilinear squares. For a horizontal flow net, b is the aquifer
225
thickness (see Chapter 6). The total volume of flow through a vertical
cross section depends on the cross section width, however, and so b refers
to that width. If we let b I unit [L], the discharge calculated from
Equation 6. 1 7 is simply the discharge per unit basin width.
=
Consider how we might use the flow net shown in Figure 7.2a. We
would have to know the length of the basin, for example L = 100 m, in
which case the hill is 30 m (0 . 3L) high on the right side. The difference in
hydraulic head between adjacent equipotentials is equal to 0.005L, or 0.5
m. If the hydraulic conductivity is 0.5 m day- I (approximately 5.8 )( 10-6
m S- I ), the discharge through each streamtube is:
The simple flow net above (Figure 7 .2a) provides a template for
understanding regional groundwater flow. However, flow patterns are
conditioned by variation in the shapc ofthc basin and the watcr table, and
spatial patterns of hydraulic conductivity. This section explores the
primary controls on the pattern and rate (natural basin yield) of steady
groundwater flow in a basin bounded by divides at the sides, the water
table at the top, and a low-permeability unit at the base . The terms
introduced in thc previous section (recharge and discharge areas, natural
basin yield) provide some keys to exploring the importance of each of
these controls.
226
7.3.1 The effect of basin aspect ratio
227
(a) Large basin aspect ratio (shallow basin)
o. ,
I"I< I 0.0
-
I
0.0 o. , 0.2 0.3
I
0.4 0.5 0.6 0.7 0.8
= O.003L
0.9 , .0
x/[
I- >\
I -'( 0.2::::
"
X \ o. ,
'f-
till = O.OO3L
0.0
0.0 o., 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 ' 0
xlL
Figure 7.4 The effect of basin aspect ratio (length to depth) on natural
patterns of groundwater flow. The water-table profile is the same for the
shallow (a) and deep (b) basins.
There is some difference between the patterns of flow in the two basins
in Figure 7.4. For the deep basin (Figure 7.4b), vertical hydraulic gradients
(and, therefore, vertical flow) exisl over a large portion of the basin. In the
shallow basin (Figure 7.4a), the flow is essentially horizontal over most of
the basin.
228
produced by regional water-table topography (Figure 7.4). These flow
systems appear to have a lower depth limit, below which flow is very slow
or non-existent, as may be seen particularly well in Figure 7.5b. However,
the flow doesn'l actually become zero, but only slows as slreamlubes
become wider and the hydraulic gradient along the streamlines decreases.
It could also be shown that the apparent "depth" of these local flow
systems depends on the local hydraulic gradient.
0.3
0.2::::
..
,
,
0.1
_�
It:"
__ ���-;�� �����
I&�I�O�.O�O�3�
LI 0.0
__
T6th, in the same studies cited earlier, referred to these flow cells
(dashed boxes in Figure 7.5) in which water flows from topographic high
to adjacent low as local flow systems. Larger flow systems, from regional
high to low, h e referred to as regional flow systems. The similarity in
appearance between thcse types of flow systems suggests that the use of
different tenns to describe them depends only on one's definition of
"regional." However, the distinction becomes clearer when we
superimpose local hill-and-valley topography on top of a regional slope
(Figure 7.6). For the case of a deep basin (Figure 7.6b), both local and
229
regional systems develop (streamtubes labeled "L" and "R" in the figure),
as well as what Toth referred to as an intermediate flow system
(SlreamlUbe labeled "I" in the figure). If the basin is relatively shallow
(Figure 7.6a), a regional system may still exist, but may be attenuated as a
result of the dominant influence of the local flow systems.
The conclusions that we can draw from these flow nets are as follows:
(a) Regional and local watertable topography, large basin aspect ratio
(b) Regional and local watertable topography, small basin aspect ratio
xlL
Figure 7.6 Groundwater flow patterns for the case of combined regional
and local water-table topography. Streamtubes labeled "L," "I," and "R"
indicate the local, intermediate, and regional flow systems, respectively.
The same equipOlential spacing (dh) ofO.003L is used for both the large
(a) and small (b) basin aspect ratios.
230
Complex water-table topography will necessarily produce complex
patterns of groundwater flow, but the general observations discussed above
can be applied to other settings. Referring to Figure 7.7, can you identify
local, intermediate, and regional flow systems?
231
heterogeneity on flow nets. Onee the model is downloaded
(http://waler.usgs.gov/nrp/gwsoftware/tdp£ltdpf.html), it ean easily be used
to explore a variety of diffcrent flow conditions to gain insights about
recharge and discharge areas of groundwater flow systems.
Mexico City lies within the discharge area for the Basin of Mexico,
which is a "closed basin." This means that there are no natural surface
water outflows for the basin, and all outflow occurs through
evapotranspiration. The Basin is ringcd by volcanic mountains that fonn
the recharge areas for the Basin. Because of the heterogeneity of the
materials filling the basin, an aquifer-aquitard system exists, and the
pattern of flow from the mountains to the center of the basin looks
something like Figure 7.8b. Prior to heavy pumping during the past 100
years, groundwater flowed from thc rccharge areas (mountains) into
discharge areas in the Basin center through the aquifers, and discharged
upward across the overlying aquitards. Surface springs and lakes resulted
from this natural pattent of groundwater flow.
Potable groundwater under artesian pressures (i.e., heads that caused
water flow to the surface in wells without pumping) was discovered in the
Basin of Mexico in 1846, promoting the rapid development of this
resource (NRC, I 995b). Since that time, the large overdrafts of
groundwater have reduced the pressures within the confined aquifers and
altered the natural pattern of groundwater flow. Now most of the water
within the Basin is moving downward toward the heavily pumped
aquifcrs; surfacc springs and lakes have dried up. In 1983, systematic
monitoring of the water levels in wells began. These records of water
levels in wells have provided important information on the water balance
in the basin.
232
1<111 = o.oo2L1
(a) Homogeneous
O. 1
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
xlL
� ,-t
-
, ,
:!: , tt,
0.0
0.0 0.1 0.2 0.3 0.4 0.5 06 07 08 0.9 1.0
xlL
� ...
,
"
xlL
Figure 7.8 The influence of geological heterogeneity on patterns of
groundwater flow. The homogeneous case (a). An aquifer at depth (b;
hydraulic conductivity of the aquifer is an order of magnitude greater than
the overlying unil). An aquitard at depth (c; hydraulic conductivity an
order of magnitude less than the overlying unit). A discontinuous aquifer
at depth (d). The equipotential spacing (dh) is the same for all, O.002L.
233
allow us to gain some understanding of these effects. As we will see in the
next sections, confined and unconfined aquifers behave somewhat
differently, and we consider each scparately.
234
1
>
-
..
�
E
-
u
.-
f!
Q.
0 r------,
-5
E
-
235
can be implemented by introducing water into wells (recharge wells) or by
routing water into infiltration basins in permeable material (the idea
mentioned in regard to Orange County in thc introduction).
-0.76
-0.77
E -0.78
�
� +-
"
.c -0.79
-
l!!
"
�
-0.80
'"
-
;: -0.81
0
-0.82
-
.t:
Co
-
"
C -0.83
-0.84
-0.85
'" '" '" '" 0> 0> 0> 0> 0> 0>
0 0 0 0 0 0 0 0 0 0
.., 0 '" "' "- 0>
- - -
<0 '"
- - - - - -
0 0 0 0 0
� �
N N N '" �
Time (day/month)
Figure 7.10 Water levels for a latc summcr period in 2010 in a shallow
well adjacent to White Clay Creek at the Stroud Water Center near
Avondale, Pennsylvania, USA.
236
characteristic of a given aquifer and allows determinations to be made of
the change in the volume of water stored over time. Values of specific
yield typically are less than the porosity, and most range from 0.0 I to 0.30.
Not all of the water in an aquifer volume will drain under the influence of
gravity. Some will be retained by forces causing adhesion of water to
particle surfaces. This is why the specific yield is less than the porosity.
Area =
1 m2
.
I
_ Original
. " . ..
,
, .
water table
· - : · 1 ·.. ·
I
I
. .1
1 m r-,-
-
I _ Water table after
I
I water removal
=---
-+-=
I
I
I
Aquifer
) ,
- - -
,
/
Volume removed = V m3
Figure 7.11 The specific yield of an unconfined aquifer. For a I-m decline
in the water table, the volume of water produced per unit aquifer area is
the specitic yield, Sr
237
which increases the hydraulic gradient toward the well. The change in
water level in the pumping well, or in observation wells nearby, is referred
10 as a drawdown. The amount of this drawdown will decrease as one
moves away from the pumping well, and the pattern that is produced is
called a cone of depression because of its characteristic shape (see Figure
7.14). The shape and extent of the cone of depression within an unconfined
aquifer depend on the pumping rale, the transmissivity and specific yield
of the aquifer, and lime. From the definition of the specific yield, however,
we know that the volume of the cone of depression is equal to the volume
of water removed (by pumping) divided by the specific yield:
I""'f"'''
V
v.
-
=
S,
(7.6)
238
solids, called the effective stress (ae):
Area =
1 m2 "
1-
/1 _
_ _
Original
// 1 //
/ 1 / potentiometric
.:: J
1 surface
_ _ _ _ _ ....
1
1
1
1 m
1 '-_ Potentiometric
1
1 surface after
1
•
water removal
Aquitard
1
,
1
1
,
Aquifer
1-
•
./
//
_ _
.
//
Volume removed = V m3
Figure 7.12 The storativity of a confined aquifer. For a l�m decline in the
potentiometric surface, the volume of water produced per unit aquifer area
is the storalivity, S. The aquifer material is not drained and remains
saturated.
This idea may be slated in another way: the weight of lhe overlying
material on a horizontal plane is supported in part by the fluid and in part
by the solid. The total stress won't val)' over time:
da,. = dp + da, = 0.
239
Therefore, any change i n pressure must be offset by a change in effective
stress:
dp = -d(J,. (7.8)
plane
Aquifer of area A
240
within the aquifer and none within the confining layers. If an aquifer is
only semi-confined, with some water coming from surrounding aquitards,
then it is referred to as a leaky aquifer. Because of this additional source
of water, the storativity will often be higher in leaky aquifers.
The extent of the eone of depression depends, among other things, on
the value of the storage parameter (either specific yield or storativity).
Consider two aquifers, one unconfined and one confined, that are being
pumped at the same rate. Because the specific yield of the unconfined
aquifer will be greater than the storativity of the confined aquifer, the
drawdown in the unconfined aquifer will be less than that in the confined
aquifer (Figure 7.14). That is, the volume of the cone of depression in a
confined aquifer is equal to the volume of water pumped divided by the
storativity [change Sy to S in Equation 7.6], a very small number. Thus
removal of water from confined aquifers produces substantial drawdown
of the potentiometric surfacc. In confined aquifers that arc heavily pumped
over long time periods, the cone of depression ean be quite extensive,
approaching tens of kilometers in lateral extent, with tens of meters of
drawdown in the vicinity of the well (Figure 7. 1 5) and a cone of
depression that extends a long distance from the major pumping center
(Figure 7.16).
241
(a) Unconfined aquifer
r
•
Water table
Unconfined
aquifer
- - - - - - -
' "
�'-" - -
"
/' -- -
--
-
Potentiometric '\,
surface
Aquitard
Confined
b
aquifer
242
Figure 7.14 The cone of depression in an unconfined (a) and confined (b)
aquifer. The wells are being pumped at a constant rate (Qp), and hydraulic
heads are being monitored at two piezometers at radial distances 1"1 and 1"2
from the pumping well.
0
"
> -5
E
-
.-
.. -
-
-10
" "
"
-
..
�
.. "
" '"
-20
"" "0
" "
. " -25
" 0
-
-
.. �
�
'" -30
"g, o
,,
\.
-
-35
-40
a � a � a � a � a � a
� � <D <D � � <D <D m m a
m m m m m m m m m m a
� � � � � � � � � � '"
Year
Figure 7.15 Water levels in a well in a confined aquifer in Calvert County,
Maryland, USA.
Data from DePaul et a1. (2008).
Other things can affect the balance between water pressure and
effective stress in aquifers and so affect the water levels in wells.
Atmospheric pressure varies as low-pressure and high-pressure systems
move across the landscape. An increase in atmospheric pressure creates an
additional force on the top of a confined aquifer that is distributed between
water pressure and effective stress. That is, only a portion ofthe force
from the increased atmospheric pressure will be transmitted to the water in
the aquifer. The increased atmospheric pressure also acts on the water
standing in the well. In the well, the full increase in the atmospheric
pressure is transmitted to the water. Thus, for a confined aquifer, an
increase in atmospheric pressure causes a decrease in the water level in a
well and, conversely, a decrease in atmospheric pressure results in an
increase in water level. For an unconfined aquifer, changes in atmospheric
pressure are transmitted equally to the water table and the water in a well,
and so do not produce changes in water levels.
243
7.5 Well Tests to Estimate Aquifer Properties
The rate at which water levels change in wells depends in part on the
aquifer properties. For example, if a quantity of recharge water is suddenly
added to an unconfined aquifer, the water levels will decline as the water
drains out to streams. The rate of decline will depend on the hydraulic
conductivity and the storativiry of the aquifer. Thus, if the water table or
potentiometric surface in an aquifer is purposefully disturbed, by removing
or adding water for example, measurements of the time variation of water
levels in wells can be used to back-calculate the aquifer properties. Below
we briefly describe two well tests that are llsed in the field.
244
--
�-50
� --
--
�� - --
-30
-
--
- -
�
-- -
�
'"
----_tIl: --,-6() -
-
--
-
-
- -
- --
-70
()
Slug tests use measurements made on a single well. The water level in the
well is suddenly perturbed by removing a "slug" ofwaler. This lowers the
245
water level in the well, creating a head gradient toward the well.
According to Darcy's law, water will then flow radially into the well until
the water lcvel in thc well is restored. The level of the watcr in the wcll
relative to the steady-state water table, H, is measured as a function of
time.
The initial perturbation inside the well is a depression of the water level
by Ho and subsequently the level varies with time, H(l). The well casing
has radius rc whereas the radius of the borehole is R. The length of the well
screen (the non-impervious part of well casing through which water flows)
is L (Figure 7. 17).
The rate of water flow into the well, Q, can be calculated using Darcy's
law.
dh
Q � 21CrLK- .
dr
where r is the distance from the center of the well and II the saturated
thickness ofthc unconfincd aquifer at distance r (Figurc 7. 1 7) .
This equation can bc rearranged as:
Q dr = ,111.
2trKL
(7.9)
r
246
I n it ial
H yd raulic co n dition s
head at t ime h
T. after slug
is removed 2R
R. ---�)
Figure 7.17. Schematic for a slug test. Re is the radial distance beyond
which the perturbation by the slug has no effect. Note that the right side of
the figure depicts conditions initially, before the slug is removed, and the
left side depicts conditions that evolve through time after the slug test is
initiated.
Equation 7.9 can be integrated across the step change in H caused by the
removal of the slug, lelting r vary from R to Re, the effective radius beyond
which there is no effect.
(7.10)
Because after the initial abrupt withdrawal no other water is extracted from
the well, a mass balance in the well itself gives the following relationship.
dH Q.
dl
- : -
(7. 1 1 )
JCrl
•
247
2KL
- =- ,
dH
dl. (7.12)
H r,,- ln( R"I R)
which can be integrated from time zero (when the slug is removed) to a
later time, t. to give:
H, 2KL
In- = - I or
Ho r/ln(R,.IR) ,
2 KL
InH(I) : - , 1+lnHo. (7.13)
r; In( R"IR)
-(sfope)xr/ In(L! R)
2L
K =
. (7.14)
For example, data collected for a slug test on well near Oyster, Virginia
(Figure 7.18) are fit with a straight line with slope -0.19 S-I . The well
casing has an inside diameter of 7.6 cm and the borehole radius is
estimated to be 12.7 cm. The length of the well screen is 5 m. From
Equation 7.14, we estimate the hydraulic conductivity for this shallow
(sand) aquifer to be
248
1.0 --------,
0.5
0.0
_ -0.5
�
c:
-
-1.0
Slope = 0 1 9 S-I
-1.5
- .
-2 .0
-2.5 :---;:-
a 2 -;-
-
4 -;:-
6 --:
8:--
-
--::
1 :--
0 -:-;:-
12 :-;-
1 4 --:
-;:-
1 6 --
1 8-!.
-
K � (0 . 1 9 s- ' )(0.076 m) ' ln(5 mlO. 127 m)l( I O m) � 4. l x I 0-4 ms- ' .
Slug tests to determine hydraulic conductivity in the field are quite useful
in many ways. They are easy to carry out, the duration of a test is short,
and only one well is involved. The measurements reflect a relatively small
volume of aquifer material around the well, however. In cases where an
average response over a larger area is the aim, tests that reflect the aquifer
properties between two wells may be preferred. The basic idea is that one
well is pumped and the hydraulic head is measured as a function oftime in
observation wells, say at distances rl and r2 from the pumping well (Figure
7.14). The heads hi and h2 will change with time as the pumping proceeds.
The equation that describes thc variation of head with respect to distance
fr om the pumping well and time is an extension of the Laplace equation
(Equation 7.4) to account for time-varying conditions. (Details can be
found in hydrogeology texts such as Fetter, 2000.)
249
°.:;
'"
,Ph +'0-
(7. t 5)
s ail
T al axl ay!
=
QL
= -,---
"' '
I'---,- - 5 1 40 m day- ' = 1 1 30 ' d ' .
T m ay-
41r(slope) (4 x Jt x 0.36 m)
( )
2.2ST (2.2SX 1 130 m2 day- 1 )
S
=
r2exp (intercept)
41fT - ("
(100 m)�exp -
�I .8:::
2 ;;:
: " :.: ,,, ,,;-
4If",
..'" X" X" m '.::
3 0:..
I ' ::: d;:,
'Y
,- ')
-..2.
Q" 5140 m 3 day- I
= 0.0010.
Notice that, unlike the Cooper-Jacob method, the slug test provides an
estimate only of the hydraulic conductivity and not of the storativity. This
is due to the fact that the slug test is based on a water balance equation for
the well (Equation 7. 1 1 ), in which the flow from the aquifer is calculated
without accounting for the unsteady character of groundwater flow. In fact,
Equation 7.9 is integrated assuming that Q is constant with respect to r.
This assumption (i.e., 8Qlor = KfYhla?- = 0) corresponds to setting to zero
the left-hand side of Equation 7. 1 5. Because the transient character of the
250
slug test's response is expressed by the water balance of the well and not
by the unsteady groundwater flow, this test does not provide any
information on the "inertia" of groundwater flow (i.e., of storativity or
specific yield).
Among the vel)' important environmental issues today are how to prevent
contamination of groundwater and, should contamination occur, how to
clean up aquifers. Groundwater flow is relatively slow in human terms, so
once an aquifer is contaminated, clean up can be extremely difficult or
even essentially impossible using reasonable resources. Knowledge needed
to determine effective protective measures and to explore the feasibility of
clean-up options includes insights from physical hydrologists regarding
rates and directions of groundwater flow.
1.8 ,-------:
1.6
1.4
E
-
- 1.2
•
-
"
� 1.0
l 0.8
c
.5 = 0.36In(I)+1 .82
0.6
0.4
0.2 L--
-5 -4 -3 -2 -1 a
In(/), time in days
Figure 7.19 Pumping test results for the Madhupur aquifer in Kapasia,
Bangladesh, for the Cooper-Jacob method.
251
of the water in which they are dissolved or suspended. The contaminants
"go with the flow" and travel along groundwater flow paths. For example,
a contaminant injected into the groundwater at the water table will move
by advection along a path detennined by the flow net if the flow is steady
(Figure 7.20). The speed at which the contaminant moves can be estimated
using Darcy's law provided the hydraulic conductivity, K, and the
porosity, ¢, of the aquifer is known. Although the specific discharge, q,
given by Darcy's law has dimensions of velocity, it is not a true velocity.
The specific discharge is discharge per total cross-sectional area of aquifer,
but the flow occurs only through the openings in the rock. Recalling that
porosity represents the fraction of the total volume of rock that is available
to water, the average linear velocity of the groundwater flow can be
estimated as v = ql¢. If we take the schematic in Figure 7.20 to represent a
local flow system in a sandy aquifer with L = 1 0 km, ¢ = 0.25, K = 1 0-6 m
S-I, and a head gradient along the flow path A _ B of 0.033 mJm, the time
for a contaminant to travel 100 m would be
�-, o ..
0.3
0.2:::::
"
0.1
I:---01�
00
--= ..L,O'::___'� :"___--;,-
03
,-""":T:
0. '
--' 0.:'--
5
__:+:
O'
--07::-
---: � ::-'-----:T
0.'
:----4:1 .00 . 0
O'
xlL
Figure 7.20 A contaminant entering the groundwater at point A will be
carried by advection toward point B.
Even with this relatively high head gradient and hydraulic conductivity,
we would estimate that the time for a contaminant to travel the length of
the domain would approach 25 years. For lower gradients and longer
paths, the times would be longer. Flushing a contaminated aquifer by
natural recharge of fresh water can require very long time periods.
Advection is not the only process that occurs as contaminants move in
groundwater. The flow paths through pores, cracks, and fractures in the
252
aquifer are convoluted and water moving along various paths mixes as the
paths converge, diverge, join, and splil. These mixing processes result in
spreading of the contaminant by dispersion. A pulse of contaminant will
spread out to fonn a plume as it moves along the flow path (Figure 7.2 1 ).
Groundwater contamination can originate from many sources, for
example from landfills that were constructed before we understood how
contaminants can leach from them and move to groundwater. Over time,
extensive plumes can fonn, threatening natural ecosystems and the quality
of streams into which the groundwater drains. Hydrologists need to
understand how transport occurs to address potentially serious
environmental problems. The U.S. Geological Survey has been studying
hydrological and biogeochemical process at a landfill near Nonnan,
Oklahoma (Christenson and Cozzarelli, 2003). The leachate plume has
spread several hundred meters from the landfill and has affected a nearby
wetland. Detennination of flow paths and rates of flow require estimates of
hydraulic conductivity. A t the Nonnan landfill site, slug tests were used to
characterize the heterogeneity of the sediments and gage the impact on
groundwater contaminant transport (Scholl and Christenson, 1998).
7--, 0. '
0.3
0.2:::!
"
0.1
Contaminant disperses
to fo rm a plume
Figure 7.21 Dispersion spreads a contaminant out as it is advected along a
groundwater flow path.
253
clays, shales) may contain large amounts of water that can move given
enough time. Figure 7.8b illustrates this possibility. In instances where
overlying units (or interlayers within the aquifer system) are clays,
pumping the confined aquifer has two effects. The first is to increase the
effective stress on the aquifer and the second is to de-water the clay units.
What happens when clays are de-watered? You may have observed the
cracks thai appear on the bottoms of mud puddles that have dried up.
These cracks reftect the shrinkage of clays upon dl)'ing. The same thing
can (and does) happen when clay units above or within aquifers have part
of the water in them removed: they shrink. The result is called land
subsidence.
As discussed, the cone of depression in confined aquifers can be quite
large (e.g., Figure 7 . 1 6), and land subsidence due to pumping can be
significant ovcr considerable areas. [n the San Joaquin Basin, California,
subsidence has affected an area of at least 4200 square miles; maximum
subsidence approaches 9 m in some areas (Helm, 1 982).
In Mexico City the lowering of the potentiometric surface in the
Mexico City Aquifer has resulted in the removal of water from the
overlying clays. The land surface has subsided by some 7.5 m in the
central part of Mexico City (Figure 7.22; NRC, I 995b). Old well casings
in the city extend several meters above the ground surface because the land
has subsided around the casings. The result of this subsidence has been
extensive damage to the city's infrastructure, including building
foundations and the sewer system. Anothcr serious problem in Mexico
City relates to flooding. The city is bordered to the east by Texcoco Lake
-the natural low point of the southern portion of the Basin of Mexico. In
1 900, the lake bottom was 3 meters below the median level of the city
center. By 1974, the lake bottom was 2 meters higher than the city! This
change is thc rcsult of greatcr land subsidence within the city and has
aggravated flooding problems. A complex drainage system (including
excavations to [ower Texcoco Lake) has evolved to control flooding.
254
2241
2240
e � 2239
-
-
0 >
.- �
.. .. 2238
- -
> �
� �
" e
..
2237
"
u �
-l!! E 2236
��
<II > 2235
' 0
"
e .0
..
..
...J E 2234
-
2233
2232 •• •
1900 1910 1920 1930 1940 1950 1960 1970 1980 1990
Figure 7.22 Land subsidence in Mexico City.
Data from Onega-Guerrero et a1. (1993).
255
the base-flow recession. The fonn of the curve for the recessions is:
..
'" o 0.3
; 0.4 o
.r:::
u o
" a> 0.2
is 0.3
0.2
'lr"
Ufo
..
.. .. .
o. , o. ,
-
May Jun Jul Aug Sep May Jun Jul Aug Sep
Month in year 201 0
Figure 7.23 Groundwater recession for Pescadero Creek, California,
showing both arithmetic (a) and logarithmic (b) axes.
(7.17)
where Q (L3 11] is the discharge at time I after recession begins, Qo is the
discharge at the beginning of the recession period, I is time measured from
the beginning of the period, and c [I I] is a recession constant.
The appropriateness of this equation can be investigated by considering
the conservation of mass. During recession, the inflow to the entire
groundwater reservoir can be assumed to be zero and we would expect that
(he outflow would be a function of some measure of the groundwaler
elevation, Ii [L]. In other words, the hydraulic head (and hydraulic
gradient) should decrease with time, and, thereby, reduce the driving force
for groundwater discharge. We could write this as:
q, = [(II), (1.18)
where qr is the discharge per meter width of stream [L2 11]. Conservation
of mass requires the outflow to be balanced by the change in groundwater
256
storage. The time rate of change of groundwater stored should depend on
;;, the storativity S or specific yield �v (depending on whethcr the aquifcr is
confined or unconfined), and the length of the aquifer L (Figure 7.24). The
appropriate conservation equation would then be:
dh -
SL- =-q, = -f(h). (7.19)
d1
Thc Icft sidc of this cquation is thc timc rate of change of watcr stored
in the aquifer per meter stream width. Now to integrate Equation 7.19 we
need a specific fonn of/(/1). The simplest relationship that might be used
is a direct proportionality:
(7.20)
dh
=-eh. (7.21)
d1
II = "oe-�', (7.22)
where 110 is the average hydraulic head at the initial time. Then, from
Equation 7.20 we have:
257
-
0'
(7.23)
258
concern. Numerical models can be used to construct flow nets describing
past or present groundwater flow patterns. Several kinds of information are
required to model groundwater flow accurately, some of which havc been
described in this chapter and Chapter 6, namely: porosity, intrinsic
penneability, and storativity or specific yield of basin materials; basin
geometry and water-table configuration; and a record of water levels thaI
can be compared with model predictions. Once the prcsent conditions have
been established, these models can be used to predict the conscqucnccs of
future or continued changes or management practices. The High Plains
Aquifer in the central United States was mentioned in the introduction as a
heavily exploited aquifer system. There is obvious concern for this
resource because of the dramatic changes in water levels that have been
observed over thc past 50 years. A numerical modeling study ofthc impact
of future (1 998-2020) management strategies for this aquifer system
predicted water-level declines of over 30 meters for some portion of the
region (Luckey et aI., 1 999). The extent of the predicted region of
declining watcr lcvels depends on the particular pumping rates assumed.
Despite the vastness of the region underlain by this aquifer system, the
groundwater resource does have a limit, and significant depletion remains
a concern for the millions of people who depend on it.
• Flow nets may be used to determine the natural basin yield-the quantity
of water, on average, that discharges from a basin. If the flow is steady,
this quantity will be balanced by the recharge rate. {Section 7.2}
• The pattern and rate of groundwater flow are controlled by the basin
aspect ratio, the water-table topography, and the distribution of
permcability. {Section 7.3 }
• The basin aspect ratio is the ratio of basin length to basin dcpth or
thickness. Shallow basins are characterized by mostly horizontal flow,
and deeper basins by significant vertical flow components and relatively
259
higher natural basin yield. { Section 7.3. I }
• Variations in the height of the water table drive groundwater flow. The
topography of the water table may produce patterns of local,
intcnnediate, and regional flow. { Section 7.3.2}
• Withdrawing water from a well reduces the hydraulic head and fluid
pressure at that point. As a result, the effective stress (the total stress
minus the fluid pressure, or 0" = 0r - p) is increased; the effective stress
is that portion of the weight of the overlying material that is borne by the
solid material of the aquifer. Water is then "released from storage" in two
ways: through fluid expansion as the pressure decreases and by expUlsion
from the compressed aquifer material. {Section 7.4.2 }
• Well tests are used to detennine aquifer hydraulic properties in the field.
Slug tests perfonned in a single well give information about the
260
hydraulic conductivity of the aquifer material surrounding the well.
{Section 7.5. 1 }
• Tests involving more than one well usc a pumping well to remove water
from an aquifer and observation wells to measure changes in hydraulic
head that occur as a result of the pumping. The Cooper-Jacob ( 1 946)
method can be used with the observation well measurements to estimate
the transmissivity and storativity of the aquifer in the vicinity of the area
of the wells used in the test. {Section 7.5.2}
Problem I. Determine the natural basin yicld (m3 yr- I pcr metcr basin
width) for the following cases.
C. For the basin in Figure 7.5a with L = 200 m and K = 100 m yel . You
might want to begin by calculating the discharge through each local
flow system (dashed box).
Problem 2. Answer the following questions for the basin in Figure 7.25.
261
B. What fraction of the total basin yield passes through the lower unit?
x (m)
Figure 7.25 Flow net for Problem 2.
Problem 3. Consider the flow net shown in Figure 7.26. The sides of the
region are groundwater divides, the top boundary is the water table and the
bottom is an impermeable boundary. Streamlines are blue and
equipotentials are gray.
D. Indicate at least one area within the flow net where flow is relatively
fast, and one area where flow is relatively slow.
262
4Oj ------
30
E
-
20
"
10
o +-....:!,- '-,-.1.-.,-'--
�-'--,
- ___.,.J.
- _- ,_!_�-
o 10 20 30 40 50 60 70 80 90 100
x (m)
Figure 7.26 Flow net for Problem 3.
A. What is the change in water-table level during this time period (m)?
Docs this indicate an increase or a decrease in water stored within the
aquifer?
B. If the specific yield of the aquifer is 0.25, and the aquifer has an area of
600 km2, what is the change in water stored over the same time interval
(m) ?
Fetter, C.W. 2000. Applied hydrogeology, 4th ed. Upper Saddle River, NJ:
Prentice Hall, Chapters 5 and 7.
National Research Council (NRC). 1995. Mexico City's water supply:
Improving the olltlookfor slistainability. Washington: The National
Academies Press.
263
8 Water in the Unsaturated Zone
8 . 1 Introduction
8. 1 1 Capillary Barriers
8 . 1 2 Concluding Remarks
8. 1 3 Key Points
8. 14 Example Problems
8 . 1 5 Suggested Reading
8.1 Introduction
In most areas, with the exception of bogs and swamps, the water table is
some distance below the ground surface. Between the ground surface and
the water table is a region in which the pore spaces of the rock or soil may
be partly filled with air and partly with water. This region is referred to as
the unsaturated zone or vadose zone and water in this zone is referred to
as soil moisture. Hydrologists want to be able to describe the flow of
water in the unsaturated zone to deal with a number of important issues.
264
For example, in the last chaptcr we discussed the concept of recharge to
subsurface aquifers. Recharge takes place most often through the
unsaturated zonc, either ovcrlying an unconfined aquifer or in the recharge
zone of a confined aquifer. Changing climate and land use practices have
strong effects on recharge through the vadose zone, and understanding
these effects is particularly important in arid and semi-arid areas (Scanlon
et aI., 2006).
One important aspect of water flow in the vadose zone is the water
balance of plants. Most terrestrial plants extract water from the vadose
zone. Plants wilt when soils become too dry because the forces holding the
water in the soil are too great to allow the plants access to the water (see
also Chapter 9). Related to the water balance of plants is the practice of
irrigation in agricuhure. Agriculture accounted for 90% of freshwater use
over the past century and water used for irrigation accounts for 90% of
consumptive use (return of water to the atmosphere by evapotranspiration)
of water that is withdrawn from streams, rivers, and aquifers (Scanlon et
aI., 2007). Given that water resources are already stressed in many regions
and that food demand is expected to increase by 50% as the global
population increases to 9 billion by 2050, there is a real need to improve
the efficiency with which we use water to grow crops (Rockstrom et aI.,
2007). Understanding the movement of soil water, and its uptake by plants
and "loss" through evapotranspiration and recharge to the groundwater
system, is essential in this regard.
Another example of an important problem in vadose-zone hydrology is
the usc of scmiarid locations with deep unsaturated zones for the disposal
of wastes. The Low-Level Radioactive Waste Policy Act was passed by
the United States Congress in 1980. This act provides for the fonnation of
regional compacts by states to supply sites for the safe disposal of low
level radioactive waste (LLRW). LLRW includes test tubes, rags, rubber
gloves, tools, and so forth used in medical research and treatment, in other
research (for example, in research in environmental sciences to study the
biodegradation of organic wastes), and in nuclear power plants. The major
concern about finding a suitable site is groundwater contamination,
because the pathway through groundwater is the one that is most likely to
be the one that places human populations at risk of exposure to
contaminants. Given the concern for transport of contaminants to
groundwater, it has been recognized for some time that disposal of wastes
in the unsaturated zone in a desert environment should be a safe alternative
for such disposal (Winograd, 1 98\).
In response to the LLRW Policy Act, Arizona, California, North
Dakota, and South Dakota fonned the Southwest Compact and selected a
265
site in Ward Valley in the Mojave Desert for the first disposal site for the
Compact. Before any site can be used for the disposal of wastes, a license
must be obtaincd. The liccnsing procedurc rcquires earcful study ofthc sitc
and estimates of rates at which radionuclides might leach into the
groundwater. One of the key parts of the analysis is the determination of
flow rates in the unsaturated zone. Even with careful study, siting waste
disposal facilities is a contentious issue and opposition can be expected to
any sitc. Opponents oftcn challenge the scientific assumptions regarding
hydrological processes. The relationships between the forces on and the
flow of water are ingredients of arguments presented by both proponents
and opponents of waste-disposal facilities. We will return to the case of
Ward Valley at the end of this chapter after we have developed the ideas
useful in describing the flow of soil moisture.
Finally, as we will see in our discussion of catchment dynamics in
Chapter 1 0, the storage within and release of water from the vadose zone
are quite important in detennining the stonnflow dynamics of a catchment.
The infiltration and movement of water in the unsaturated zone represents
one potential pathway for precipitation entering a stream. Variation in the
ability of water to infiltrate a soil or rock is, therefore, an important aspect
of catchment water dynamics.
The early literature recognized three divisions within the unsaturated
zone: the capillary fringe, the intennediate belt, and the belt of soil water
(e.g., see Meinzer, 1923). According to Meinzer, the capillary fringe is "a
zone in which the pressure is less than atmospheric, overlying the zone of
saturation and containing eapil1ary intcrstices some or all of which arc
filled with water that is continuous with the water in the zone of saturation
but is held above that zone by capillarity acting against gravity." That is,
the capillary fringe is a saturated zone above the water table where water is
affcctcd by capillary forccs. Thc uppcnnost belt, or bclt of soil water, is
"that part of the lithosphere immediately bclow the surface, from which
water is discharged into the atmosphere in perceptible quantities by the
action of plants or by soil evaporation." This definition recognizes that
plants, for the most part, extract water from a portion of the soil (the "root
zone") ncar thc surface. The intennediate bclt is "that part of a zone of
aeration [i.e., the ul/saluraled zone] that lies between the belt of soil water
and the capillary fringe." The intermediate belt is distinguished mainly by
the fact that something must be between the root zone and the capillary
fringe. The distribution of moisture abovc the water table is what
motivates the definition suggested above. The volumetric moisture content
(or simply moisture content) in the capillary fringe is the saturation value
(Figure 8. 1 ); that is, the pores are completely filled with water.
266
Volumetric moisture content, or more precisely volume wetness, is
defined as the volume of water per bulk volume of soil sample. We will
usc the symbol () to represent volumetric moisture content. After a rather
rapid decrease from saturation, the moisture content in the intermediate
belt may remain fairly constant. Field capacity is a term used to represent
this "constant" moisture content. The moisture content in the soil water
belt decreases rapidly from the field capacity due to the extraction of water
by plant roots and to direct evaporation at the soil surface.
z
Soil water
zone
Unsaturated Intermediate
or vadose zone
zone
- - - - - - - - - - - - - - - - - - - - - - - - - - - _ .
Capillary
fringe
v
8
Saturated
zone
Figure 8.1 The distribution of moisture in the vadose zone and the
classification of waters according to Meinzer ( 1 923). Water near the
surface of the soil is available for uptake by plant roots. After several days
affair weather, the moisture content in this belt of soil water (or rool zone)
decreases substantially due to evapotranspiration. Directly beneath the root
zone, the moisture content tends to be fairly constant over a depth of up to
a meter or more. The relatively constant value of moisture content i n this
region is referred to as the field capacity of the soil. Near the water table,
the pores of the soil act as ';capillary tubes" and remain saturated even
267
though the pressure head in the water is negative. This saturated zone
above the water table is the capillary fringe.
The divisions of the vadose zone often are useful i n describing general
observations of soil moisturc. Of coursc, there arc no sharp dividing lines
marked off in the field. The physical principles that we use to quantify
flows in the vadose zone do not change in moving, say, from the root zone
to the intennediate zone. The terminology introduced above is used
widely, however, and we will encounter these descriptive terms later. Our
introduction to the physics of soil moisture will hold for all of the zones.
If water is withdrawn from a rock or soil matrix that does not shrink upon
drying, air enters the pore space, and air-water interfaces (menisci) are
present in the pore space. Such curved interfaces are maintained by
capillary forces. Surface tension acting in the interfaces provides a
mechanism of soi I-water retention against externally applied suction. This
phenomenon is seen in the rise of water in capillary tubes, for example,
268
and is explained by the attractive forces between the glass walls of the tube
and the water. The glass attracts the adjacent water molecules more
strongly than do other water molecules themselves. The water is therefore
"pulled" up the inside of the tube (Figure 8.2). This "pull" is a tension
which, in the terms that we are using, is a negative (gage) pressure. That is,
the gage pressure in unsaturated soils is Ilegative. Negative pressure heads
are developed in unsaturated rock and soil matrices. The height of rise in a
capillary tube (a measure of the ncgative pressure head) is inversely
related to the diameter of the tube. Water will rise higher in a tube with a
small diameter than it will in a tube with a large diameter. This observation
translates to soil physics in that smaller diameter pores retain water against
higher suctions than do larger pores (cf. the large capillary tube and the
small capillary tube sketched in Figure 8.2). Thus, when water drains from
a soil or rock, large pores empty first because it takes relatively less
applied suction to pull water out of larger pores. The negative pressure
produced by capillary forces, when divided by pg, is referred to as the
capillary-pressure head.
AI first glance, the idea of negative pressure head does not seem
intuitive to most people. Yet there are commonplace examples with which
most of us are familiar. If asked about the pressure head at the surface of a
pan of water in your kitchen, you should be comfortable with the answer
"zero." But what happens if you bring a dry sponge just barely into contact
with the water surface? Water rises into the pores of the sponge. The
sponge pulls water up from a state of zero pressure head. The "pull" is
capillary tension, a negative gage pressure. Conversely, if the sponge
subsequently is withdrawn from the water surface, all of the water does not
drain out. The sponge retains water against the downward action of gravity
by counteracting the downward gradient in elevation head with an upward
gradient in (negative) pressure head.
269
Height of
capi lla ry Pair
rise is
inversely A
related
to the
radius
of the tube
-1 ---., PB Pc Pair 0
h, = = =
PA ; PB - pghz
Pcapillary = Pair PA-
Figure 8.2 Surface tension "pulls" water up into capillary tubes. Water
pressure within the tubes is less than atmospheric pressure, or is negative
in gage units. The height of water above the free surface in the tube is
equal 10 Ihe negative of the capillary-pressure head. The amount of
negative pressure head with which a capillary tube can " hold" water is
inversely related to the diameter of the tube. That is, small-diameter tubes
(and by analogy, soil pores) hold water at a more negative pressure head
than do large-diameter hlbes (or soil pores).
For areas where there are moderate fluxes ofwatcr through the vadose
zone, the two major driving forces on soil water are the gradients in the
negative capillary-pressure head and the gradient in elevation head. This
situation is exactly analogous 10 that for flow in the saturated zone, with
the only change being that the positive pressure heads encountered in
groundwater are replaced by negative capillary-pressure heads. Thus, the
hydraulic head for the vadose zone is defined to be the sum of the head
due to gravity and the (negative) capillary-pressure head. (In general,
several forces act 10 create the negative pressure heads in the unsaturated
zone. The treatment given here remains valid, but an "equivalent" negative
pressure hcad that incorporates all important forces, rather than just
capillarity, is used. See Childs ( 1 969) or Guymon ( 1 994) for further
explanation. The material that we present is strictly valid for relatively
270
moist soils and rocks and is valid for almost all circumstances with the
extended definition of negative pressure head.)
A suction (or negative pressure relative to atmospheric pressure) must
be applied to withdraw water from the unsaturaled zone above the water
table. The greater the applied suction, the more water is withdrawn, and
the lower is the soil-moisture content when the soil has reached
equilibrium with the applied suction. Our example of a sponge may
provide insight. In the kitchen, we don't have devices to exert suction to
remove watcr from a sponge. We can make an analogy bctwcen exerting
suction and "squeezing" the sponge, however. lf we exert small effort in
squeezing a sponge, a relatively small amount of the water will be forced
from the pores. To paraphrase the sentence referring to exerting suction on
a rock, the greater the applied " squeezing" to the sponge, the more water is
withdrawn, and the lower is the amount of water remaining in the sponge
when "squeezing" has ceased. The relationship between the external
suction applied to a rock and the amount of water per bulk volume (the
moisture content) that the rock retains against that de-watering suction is
called the moisture characleristic. The applied negative pressure is a
measure of the water-retaining forces of the soil and represents the
capillary-pressure head, W. The moisture characteristic generally is
presented as a plot of II' versus (}.
The moisture characteristic for a porous material can be detennined
using a pressure plate apparatus (Figure 8.3). The rock (soil) sample sits
on a porous plate made of a fine-grained material (e.g., a ceramic) that
remains saturated even at high negative pressure heads. The sample is
allowed to equilibrate at a given negative pressure head and the moisture
content associated with this capillary-pressure head is determined. The
experiment is repeated at different negative heads to obtain other points on
the moisture characteristic. The locus of all such points then defines the
moisture characteristic (Figure 8.4). The moisture characteristic is one of
the important curves that define the relationships among hydraulic
variables in the soil-water system. Another is the relationship between
moisture content and hydraulic conductivity.
271
Soil or rock sample
/
I
I
Figure 8.3 A pressure plate for measunng capli\ary-prcssurc head. The
• •
rock sample is placed in contact with the ceramic plate which is saturated
with water under a negative pressure head that is set by the distance of the
free water surface to the right of the diagram below the ceramic plate. The
moisture content of the sample is recorded after the sample has come to
equilibrium with the selected capillary-pressure head. This measurement
gives one point on the moisture characteristic curve.
As water is drained from a saturated soil, the large pores fill with air but
the smaller spaces remain filled with water. The water in the smaller
capillaries is held by surface tension more tightly than is the water in the
large pores. In a moist but unsaturated soil, water flows through the water
filled pores while avoiding the larger, air-filled spaces. For a given
moisture content, we might conceptualize this flow as if the air-filled pores
were filled instead with a solid (e.g., wax). We would thus have flow in an
equivalent, saturated porous medium and would expect Darcy's law to be
valid, but with the hydraulic conductivity appropriate for the equivalent
medium. This is, in fact, found to be the case; Darcy's law is valid for
unsaturated soil but each moisture contcnt corresponds to a different
equivalent saturated medium, and hence a different value of hydraulic
conductivity. We therefore write the hydraulic conductivity as K(f})
indicating that it is a function of moisture content.
272
The hydraulic conductivity decreases very rapidly as the medium
becomes unsaturated (Figure 8.5). This can be explained by recalling that
Poiscuille's law (Equation 3.36) indicates that mean velocity is
proportional to the second power of the diameter of a cyl indrical tube.
Hence, as we have seen, the intrinsic penneability of a porous medium is
proportional to the square of the pore size (Equation 6.7) . Because the
discharge is equal to the mean velocity multiplied by the cross-sectional
area of now (7tlJ2) i n the case ofa cylindrical lUbe), the discharge will be
proportional to the fourth power of the tube diameter. The large pores of a
soil become air-filled first as a suction is applied to a soil, thereby
relegating now to the smaller pores which (according to Poiseuille's law)
can conduct water at milch lowerflolV rates than could be handled by the
emptied (larger) pores. This reasoning also can explain why the hydraulic
conductivity of an unsaturated clayey soil may be appreciably higher than
that of an unsaturated sandy soil (see Section 8.10).
0.4
", 0.35
"
"
-
" 0.3-
.
"
-
0
"
" 0.25-
"
�
'"
-
0.2
0
.-
E
"
.-
0.15
�
"
-
E
"
0.1
0
-
> 0.05
o
0.60 0.55 0.50 0.45 0.40 0.35 0.30
Negative pressure head, -\jI (m)
Figure 8.4 Moisture characteristic for a fine sand detennined by starting at
saturation and draining the sample. Note that for this sand, saturation is
maintained for capillary-pressure heads between 0 and -0 .36 m . The
capillary fringe in such a material would be 0.36 m high. As the capillary
pressure head is reduced from about -0.40 m to -0.45 m, the moisture
273
content drops sharply from the saturation value of 0. 35 to about 0. 1 5 . This
steep drop is Iypical for sandy soils. Much of the pore space is in large
porcs that drain once a critical suction is exceeded. Thc moisture contcnt
therefore drops abruptly. On the other hand, moisture content drops by
only about 0.08 as capillary-pressure head drops from -0.45 m to -0.60 m.
l D� ,-
---,
-
-
UI
,
10<
E
"
�
z:.
-
>
.-
10·
u
.-
"
..,
I::
0 10·
u
u
.-
"
-
..
.., 10-1-
�
,.,
:I:
lD·i-�---r-----'--r---�--,--1
0.05 0.1 0.15 0.2 0.25 0.3 0.35
Volumetric moisture content, e
Figure 8.5 Variation of K with () for a fine sand. Note that the scale for K
is logarithmic. The hydraulic conductivity decreases by orders of
magnitude as capillary-prcssurc hcad drops from -0. 35 m to -0.55 m
(moisture content decreases from saturation to about 0.06 as seen in Figure
8 . 4).
q= -K(8) d (
dh
(8.2)
274
8.5 Vertical Water Movement
dh
q.
. =_K(8) (8.3)
d,
or,
(8.4)
(8.5)
The left side oflhis equation represents the rate of change of mass in a
small control volume and the right side is the difference between the
inflow rate and the outflow rate, each expressed on a per unit volume
basis. That is, the equation has the same conceptual basis as the continuity
equation that we used in earlier chapters. Combining Equations 8.4 and 8.5
gives an equation that, along with infonnation on the relationships among
0, '1', and K, describes the flow of water in unsaturated rocks . The resulting
equation, referred to as the Richards' equation, is:
Q8 �[K(8)(Q'I' Ill.
(){
=
(); a;
+ (8.6)
For steady flow, the time derivative of moisture content is zero and a
single integration of the right hand side of Equation 8.6 recaptures
Equation 8.4. That is, for steady flow the specific discharge, q_, is constant
275
and can be calculated from Equation 8.4. Of course, Equation 8.4 can be
used to calculate the flux in unsteady flow as well. Under unsteady
conditions, q_ changes with time so the calculation can be taken to give a
"snapshot" of the flow at a givcn timc. Solutions to the Richards' equation,
which accounts for time variation explicitly, yield a complete time history
of heads and fluxes for specified conditions.
A simple example illustrates the kind of calculation that can be done
using Equation 8.4. Suppose that, at noon on a day in September, moisture
content in a fine sand is measured to be 0.25 at an elevation of 3 m above
the local watcr table and to be 0. 1 5 at an clevation of 3.5 m above the
water table. Assuming that the hydraulic relationships in Figures 8.4 and
8.5 hold for the sand in question, we can estimate the direction of flow of
water and the magnitude of the flux. Consider first the direction of flow.
From Equation 8.3 we see that, if iJhliJz > 0, then the flow will be
downward (because the calculated q� will be negative). Conversely, if alii
Oz > 0, then the flow will be upward. We recognize that this conclusion
merely reiterates the main point of Darcy's law, that water flows down a
gradient in hydraulic head. So, for our case, we approximate the derivative
by a finite difference:
276
Now consider how we would expect the moisture content profile to look in
a homogenous material above a static water table under conditions of zero
vertical flux. This would be the situation we might expect from the
capillary fringe through a good part of the intennediate zone after a
prolonged period without rain . In this case, Equation 8,4 becomes:
(8.7)
Dividing through by K (we can assume that K is not zero) and rearranging,
we have:
d'l' = -1
d=
0'
(8.8)
(8.0)
That is, the capillary-pressure head balances the head due to gravity so
there is no hydraulic gradient and thus (by Darcy's law) no moisture flux.
Equation 8.9 can be used to infer the moisture distribution if the
moisture characteristic (rp as a function or8) is known. Clearly, the
equilibrium moisture profile above a water table has exactly the same
shape as the moisture characteristic because of Equation 8 . 9 . Thus, for a
sandy soil with a very steep moisture characteristic at the dry end of the
curve, the field capacity represents the "nearly constant" moisture content
associated with the steep portion of the curve (Figure 8.6). The moisture
characteristic for a soil having a higher clay content, however, may not
exhibit such a steep characteristic and i n such cases the unambiguous
definition of a field capacity is problematic.
277
1
EE 08
l> "
- -
, :;; .
"tI _
-
'"
"' �
" " 0.6 �
"' 16
!! �
" "
"' >
"' 0
" .c
5. '" 0.4
" c:
> .2
",
.- -
"' >
-
C»
,, "
Z w
-
"
0.2
o
o 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
Volumetric moisture content, 8
Figure 8.6 The profile of moisture conlent above a water table under
conditions of zero venical flux of water. The equilibrium profile under
these conditions has the same shape as the moisture characteristic. For fine
sand, the moisture content drops rapidly above the capillary fringe and
remains relatively constant in the range of 0.04 to 0.08 over a sizable range
of elevation. This relatively constant moisture content is referred to as field
capacity.
278
cup mounted on the cnd (Figure 8.7). The cup, typically made of ceramic
or Teflon, is porous but with rine pores that remain saturated under the
water tensions (i.e., capillaJy-pressure heads) to be measured. The tube is
inserted into the soil, ensuring that a close contact is established between
the porous cup and the soil. The tube is filled with water and tightly
capped . A pressure gage is used to measure the pressure in the water.
Suppose the tensiometer is placed in a soil in which the capillary-pressure
head is -0.3 m. Immediately after the tube is filled with water and capped,
the pressure head inside the porous cup can be assumed to be zero.
Because the porous cup is in contact with the soil, water is drawn out of
the (otherwise sealed) tensiometer through the porous cup, lowering the
pressure in the water inside the tube. Only a small amount of water needs
to be withdrawn from the tensiometer to lower the pressure due to the low
compressibility of water; as the water pressure inside the tensiometer is
lowered, the water expands (slightly) because the porous cup remains
saturated and air cannot enter. Waler will eease to flow OUI oflhe
tensiometer when the (negative) pressure inside the tensiometer reachcs -
0.3 m, the tension in the surrounding soil water. Thus, by allowing a
tensiometer to equilibrate with the surrounding soil and then reading the
pressure gage, we get a measure of the capillary-pressure head in the soil.
Pressure gage
Porous cup
279
Figure 8.7 A tensiometer, consisting of a closed tube with a porous cup at
the end. Ncgative prcssurc, or tension, in thc unsaturated zone is rccordcd
by the pressure gage as watcr in the tube is drawn into the partially
saturated soil.
280
seil muisture ean be rapid and large. One ufthe inest impertant efthe
surtaee hydrelegieal preeesses is infiltratienj the meyement et‘ rain and
melting snew inte the sail. The preeess elitttiltratien at the sail surfaee is
inherently unsteady and a rigureus treatment en" the preblem usually is
apprrtaehed thrdttgh the use dt'Riehardsi equatimt, Equatimt 8.6. A
simplified treatment etitlte preblem. first presented by Green and Ampt
(It‘ll 1}. illustrates hew seil physieal prineiples ean be used te understand
the infiltratiun preeess. The Green—Ampt equatien will he eutlined after a
physieal deseriptinn et‘the int‘iltratidn preeess is sketehed.
The maximum infiltratiun rate ledmmdnly ealled the infiltratiun
eapaeity} dt'a sell is the rate at whieh water will mnye yertieally
duwnward when a supply uf water at aeru (gage) pressure is maintained at
the surlaee. The inliltratien rate in a dry seil is initially yery large. Eyen
theugh the hydraulie eenduetiyity is law. the eapillary pressure gradient is
estretnely large heeause the sail unly a few millimeters belew the surfaee
has a yery law {and negatiye) pressure head. Fer example, if the eapillary-
pressure head 10 mm beluw the surtaee is initially —2.0 m. the finite-
dil‘t‘erenee appresimatien ef the gradient at the surfaee [assuming that the
surt‘aee is at the saturatien meisture eentent} weuld he (0.0] m I 2.0
n't}.-"'U.Ol m = Ell]. Eyen with a hydraulie ee-nduetiyity at l [l ti m s", the
intiltratinn rate writtld he ahnut 0.0002 ttt s" 2 TED mm hr". A rainfall rate
dt‘TEfl mm hr‘i wmtld be quite intense! The intiltratinn rate deereases itt
time as infiltrating water muistens the surfaee layers and red tie-es the
gradient in eapillary-pressure head. Ultimately. the nearwsurt‘aee pertien el‘
the sail appreaehes saturatien, the eapillary—pressure head gradient
appruaehes aere and Equatinn 8.4 indieates that the infiltratidn rate
apprttaehes the saturated ednduetiyity {Figure 51.9]. The infiltratidn-
eapaeity eurye Fur any giyen seil. ufeettrse, wettld depend en the suil
preperties heeause the eapillary—pressure head gradient and unsaturated
yalues et‘ hydraulie eunduetiyity guyern the time prugress et' infiltratien.
�
.. . �: : . . ' .. ... . " .. " . . ..: ,. . . . . . . . : .. . . .. . . . . . , .... .. . .. . .. . . '' :",
' . " , "
. .
.
: '
.
: .
"
':
' : .
.
"
..
:
"
': . "
..
'
";:}.��. {.. :,/;.:. :.:;./�,.:..::..:.. /.. .'. ":';,: ,..:.-;:./. '.:::. , : :�:. ; ; ::,.:..::..:.-.- :/:". ,/.:.:. :. ; ";:.:,.->/:.:>. :.
: :
.
.
. .
"
"'
: :
,
.
:
. .
. . .
.
. .
"
� .
. �. ' " . . . .
. 8
.. . . . . . .. . . . ..
: : :
'
'
.
'
· .
' . '
. .. :
. .. .
.:
.. . ;., : : . ' .. .. : . . : .
.
. .
. . .
.
. . . ..
.
.
. . . . .
' '.. . :' . . . . ; ' . . : .. ' ... :' . . ; .'. . : : ..: . '.. . :' . . . ; .'.
.
.
. " . .
. . :. .:.: ... : . . .� . . . . . . . . . . . . ,. .
"
'
.:
. .
: :. : . . : . . . : : : . : . . : : : : . : : . . . . : . :. . : . . : .
, . " , . , .
. .
. .
. .
.
.
: : , " : : : · ·· · ::· : �' : · ·: ' ' f:t· · · · : . . : : . . .: . .. : , : ; : : . : < : . � '. ' : . . : : . <
. . .
. .
. . .
. .
.
. .
. . : . .N h. . . . . . . . . . .,
: .
.
.. . . . . .
. . .
.
.
.
.
' .
.
. :
>.. ' .:.'. · : .·w. en ·· . · ·:. · :.·: .· ." . > . . .. :.. : .. ." .. . >. . ' .:. . : .. : ..
'. " . .
. . .
"
' .. . .· ·· . ·. ·�.,
. " " , . " .
'
. -
: ; : : , : : :' :: ': ... , . : :: ' . : : ': : ', ' . : : :' : : : : ': ... . : :: :
.
.
.
.
.
. "
.
. .
. ':-: : .: :. :. :S. · : .ll/ . ". · .:....:.z· : . . . : : ': ' . .: : :. : : . . .: : .. ..: .: :. : : . .
'
"
. .
.. . ..
.
'.. . .
· '
.
-
:
' . .
.
... .
.
.
. . . . . . . � . ."f . . . . . . . . . .. . . .. . . . . . . . :
.
. . .. . . ..
.
': '. :. . :. . " '.'... . :: ' :. . : . : :'. ' . . ' ::'. :. . :. .. : '. ' . ' :: ' :. :. : ..
'
' .
. . .. .. . . . .. .
. . . . . .. . . . ' .
. .
.
. . . . . .
; :: : . : " . . .. . . . :: ': . .
. .
: . .'. : . : :: ' . ;.:: : ::: . . "" . :: :
. . .
' .
. . . . .
.
.
. . . ' . . . . . '.
.
. .
. .
. . .
.
:
'
'
.
:
.
. :': ::. : : :::.:.: . . . :': : :. >: : ::. : . ...... . �. : : . ."". :. '. :.: ... . :': ::. >: : ::.. : ::. . :': : . >: : .
. .
.
.
.
. .
.
.
.
.
.. . ' . . ""
.
.
. . . .
. . .. .. . . .\Jf. . . . .: . . .
.
..
.
·
>.:.: ." ::." : :' : . >.:.: ." ::: : :' : . : -' : " �, :: ' : : ' : . >.:.:." ::: : :' : . : -. : .: ." : " :
'
"
. . . . . .. . . .. . . . . . .. . .
.
.
. . .
. .
.
. . . .. : .. : .
. .
.
'.: ' .. : . . .. . .. . : : . . , ' .. . : . : . : '
.
. . : : . '.: ' . ';" . .. .: :
. .. :. . .. .. ". . 7
.
. > : .
.
"
Q . .' . . : '. . ".'
.
· . ,
'
. .
.
. ' "
.
,
. " .
. .
.
.
.
'
,
,
,
,
,
,
,
,
,
,
,
Figure 8.8 The vertical profile of capillary·pressure head under conditions
of no flow is a straight line (w = -z). If a measurement of capillary·
pressure head falls on this line, it is reasonable to infer that water flow in
the vertical at that point is negligible. If measured capillary pressure is
smaller (region A) than it would be at equilibrium (for a fixed value ofz),
it is rcasonable to infcr upward movemcnt of water at that point.
Conversely, if measured capillary pressure is greater (region B) than it
would be at equilibrium, downward flow is indicated. The inferences are
exactly correct under steady flow conditions.
282
1.6
1.6
Infiltration
-
E 1.2
.-
E
E 1
-
Infiltration of rain at a steady
.. rate of 1 . 1 5 mm min-1
..
-
c
�
0.6
0
.-
.. 0.6
-
�
-
-
c
.-
-
0.4
0.2
a
o 10 20 30 40 50 60 70 80
Time (minutes)
Figure 8.9 Infiltration into a sandy loam. When water ponds at the surface,
the infiltration ratc (blue curve) starts out aI a very high value because of
the large downward gradients in capillary-pressure head and decreases
steadily toward the limiting value of the saturated hydraulic conductivity
of the soil, a rate that would be achieved when the downward gradient in
hydraulic head was due only to the elevation head. When water is
sprinkled on a soil surface at a relatively slow rate, the infiltration rate will
equal the sprinkling rate initially. Ifthe rainfall rate exceeds the hydraulic
conductivity of the soil and if the rainfall persists for long enough, the
ability of the soil to infiltrate water will drop below the rainfall rate after
hydraulic gradients near the surface become reduced by the wetting
process. The time at which the infiltration rate starts to drop below the
rainfall rate is known as the time to ponding. This occurs to the right of the
infiltration capacity curve because the initial rates are much lower for the
sprinkling case than for the ponding case; approximately the same total
volume of water must infiltrate in the sprinkling case as in the ponding
case before the infiltration rate starts to drop for the sprinkling case.
283
to cause initial ponding of water, that is, the infiltration capacity exceeds
the rate of supply . Initially, all of the water reaching the surface infiltrates.
Providcd thc rainfall rate is largcr than the saturatcd hydraulic conductivity
and that rainfall continues for a long enough time, the infiltration capacity
will decrease continually as the soil becomes wet. The infiltration capacity
ultimately will be reduced below the rainfall rate. Subsequently, the rate of
infiltration will be controlled by the rate at which the soil can transmit
water and "excess water" will accumulate at the surface (Figure 8.9). This
excess water will flow rapidly over the surface to streams and can be an
important process in generating stormflow in streams and rivers, as is
discussed in Chapter 1 0.
284
+
=0
,
,
,
,
,
: :r-:-:--:-:--:-:-� J l � t . . o .
. " . . . .. .
' : :-. " : .. . . ... . :.:... :: .... IjI = IjI'J :
. ". . . . . ' . . . : : :
. .:.> :.� ':.
'-. -
. '. . . ., .
:�. " " '.
' .
0 : : ' "" , : " . . " , .' .. .. . . . :..' : .: : :
.
.
. ..
,
.. .
, ,
.
>
.
.
. .
.
-. . . .
, ,
. .
,
. ..
•
. . . _
.. . . '.", at t,· m e I The so,· 1 . . ... .. . ... ..' ; .. .' : " �.. . . . . " >:
. . . . : ::. . . . ; ..
." : .
.
. z, ·
· ·; s
". : . ... . . . i i n itially d ry .
, : . ;
: . ; . .. .. : . : ..
,
. . " ' .: . .
.. ..
' ::>
:
.>:: . :
. . .. " . .
" . . . . . ..." ' .. . " :. .." " " .:".
. " .
-
:
.
:
..
.
(8.11)
Furthermore, for the type of flow envisioned by Green and Amp! (often
referred to as plug or pistol/flow), the total cumulative amount of water
infiltrated, /, is equal to the product of Lf and /J.O, the difference bctween
the saturated moisture content, Os, and the initial moisture content, 0i"
Then, because the infiltration rate, i, must equal the negative of the time
285
derivative of / (negative flux means downward motion),
60
tlL, _ (-'I1- + L, )
-- K\
. . (S.U)
dt Lf
K I= [ L1
�
:..c
.' L + !/f ln l + 1 .
M " (-",, )
(8.14)
or equivalently
(8.15)
286
calculate a corresponding value for t. For the infiltrometer data, the Green
Ampt equation filS the measured data reasonably well with Ks = 0. 8 mm
min- I , 11'/ = -250 mm, and 110 = 0.1 (Figure 8.1 1).
The Green-Ampt method assumes that the soil column is homogeneous
and the water table is deep in a way that the wetting front does not reach
the water table during a rainfall event. In the presence of a shallow water
table or of a bedrock at a depth Ls' infiltration water may saturate the soil
column . Rainfall onto saturated soil consequently generates runoff, which
is not due to limited soil infiltration capacity but to limited soil storage
capacity, as discussed in Chapter 1 0. The soil column becomes completely
saturated when the cumulative infiltration I is equal to the initial soil
storage capacity 6,(} Lr Thus, the time (since the beginning of the
rainstonn or snowmelt event) when the soil becomes saturated-and
runofT is produced regardless of the soil infiltration capacity-can be
calculated using Equation 8. 1 5 with 1 = 6,(} Ls.
60 ,-------,
50
.!i
�
� 0
oE4
.r: E
- -
Q."
Gl Q) 30
,, -
� �
-
>-
.- -
� -
f: 20
.-
� .-
_
E
�
(.)
10
o ,
o 5 10 15 20 25 30 35
Time from start of infiltration (min)
Figure 8.11 Cumulative infiltration measured for a loamy soil in central
Virginia (blue circles) and calculated using the Green-Ampt equation
(black iiI/e).
287
8.9 Field Measurements in a Soil
Soil surface
TDR probes
288
device for measuring soil moisture.
e 40
,
i;i -o
.- r
0 -- 35
E O:
u -
c v
" 0
30
._
0 -
E S
-
' u
/
g 25 , , , , , , ,
o 1 2 3 4 5 6 7 B 9
0 -------,
o
,
-
'.
::le-O· .1 . t-
e --
o.�-02
�,,� .
-
- -
0 -0' 3 ':
a�
il
-O.4 .f"
0
.t: -
1
-,--
-
2
.,--
3
---,--
-,--,--
--
5 6 7 8
--1
9
4
-
289
content (top) in a soil during a sprinkling experiment. Sprinkling at a rate
of I I mm h..-1 was started at time zero and continued for 4.5 hours.
Subsequent to 4.5 hours, there was essentially no flux at the soil surface
and water drained vertically downward.
The moisture characteristic for the Maine soil shows a hysteresis loop
(Figure 8.1 4). Initially, the soil is relatively dry (Y' is approximately -0.40
m, (} is approximately 26%). As the soil wets up from the sprinkling, Y'
rapidly increases to about -0. 1 6 m. As this change takes place, (} increases
relatively little, from 26% to 27%. The next phase of change is a relatively
slower increase in VJ to about -0. 1 5 m, but a correspondingly large
increase in (j to about 38%. After sprinkling ceases, the moisture content
and capillary-pressure head covary along a different characteristic curve, a
branch above the curve defined by the wetting of the soil. Hysteresis refers
to a phenomenon whereby different paths are followed on the VJ (j plot
-
290
content" will be higher for a given applied capillary-pressure head when
the tube is drajnjng than it will be for the same capillary-pressure head
when the tube is imbibing water. This is exactly the fonn of the hysteresis
loop observed for the Maine soil.
'"
0
'"
-
c
"
" . §'�
c �(;'
-
0 <
>'
u
"
"
�
.<II
-
0
-
E
u
.-
"
-
E ng
" '/'l etti
r
0 ---
-
>
-OAO -0. 5 -
0 10
Capillary-pressure head, \If (m)
.
291
(a) Inbibilion (b) D raining
Figure 8.15 Hysteresis in a capillary tube with a bulb-shaped expansion.
As shown in a, the "moisture content" oflhe tube remains low during
"imbibition" of water, because water cannot entcr the large-diameter
portion of the tube until the capillary-pressure head increases to the value
associated with the large diameter. On the other hand, as shown in h,
during "draining" the "moisture content" remains high until the capillary
pressure head drops to the critical value for the small-diameter tube.
292
to build in this horizon before water "pushed through" into the next lower
horizon, Bs2. Clearly there was an impediment to flow into the Bs2
horizon that caused watcr to "back up" until positive pressures increased
the hydraulic gradient (and hence the flow) into the layer. In the C horizon,
the positive pressure heads indicate that a water table developed with a
saturated depth of 100 mm above the tensiometer. The changes in
hydraulic properties with depth influencc the dynamics of vertical soil
water movement in layered soils. Interestingly, the "backing up" of the
water above the Bs2 horizon is probably because the Bs2 horizon is
coarser than the Bhs horizon. The resistance to flow in unsaturated
materials provided by coarse layers of sediments (described further in the
next section) is an important but somewhat non-imuitive phenomenon.
0.2
0.1
•
E
. •
•
. • •
·
•
•
•
. . .
• •
•
-
• -
�
. .
- •
•
• - - -
. .
. .
'.
•
.
.
0.0
• . .
. . ..
�. . . . . .
. .
. . . . • .
• . • . . . . -
'C
.
� • •
• •
'"
• •
•
• • •
�
• •
',' .
•
.
• •
-0.1
• •
. •
CI.I
.
Oa horizon
. . •
(0.05 m depth)
• •
"
- • •
'"
•
•
• •
'" ·
•
·
-0.2
•
CI.I
•
•
• -
•
Q.
-
•
..
•
•
• .
� ·
•
.
. .
·
•
- -
-
�
•
. -
-0.3
•
•
Q.
•
.-
'"
•
·
•
U
•
-0.4
•
•
·
-0.5
•
. .
o 1 2 3 4 5 6 7 8 9
Time since start of sprinkling (hours)
Figure 8.16 Measured pressure heads in four horizons ofa forest soil in
Maine for a sprinkling experiment. Sprinkling began at time zero at a
constant rate of I I mm hr-1 and stopped after 4.5 hours. The lag in the
pressure response at the various depths is due to the time of travel of
moisture vertically downward in the soil. The pressure builds to slightly
positive values in the Bhs horizon, indicating that moisture flow into the
underlying Bs2 horizon is impeded. A water table builds up in the C
293
horizon because the base material is a glacial till with very low
penneability.
294
evapotranspiration involves the usc of weighing Iysimeters. A weighing
Iysimeter is essentially a large pot, filled with soil, and mounted flush with
the soil surface. A device to record the total weight of the Iysimeter is
placed beneath the Iysimeter. Plants arc allowed to grow in the Iysimeter.
The amount of precipitation that infiltrates through the surface is recorded
carefully as is the amount of water that drains from the base of the
Iysimeter. Changes in the water content within the Iysimeter are
determined by changes in the total weight. All of these measurements on a
Iysimeter can be done with high accuracy. Evapotranspiration then can be
computed directly from Equation 1.7 because all tenns except et are
known.
Rates of removal of water from the unsahlrated zone by
evapotranspiration are controlled by a number of factors. One of these
controlling factors is the wetness of the soil itself (Figure 2.1 1). Typically,
if a vegetated surface is supplied with plenty of water (e.g., a well-watered
lawn), evapotranspiration will be controlled by atmospheric conditions.
That is, evapotranspiration will proceed at the potential rate, which is a
function of solar radiation, wind speed, humidity, and so forth. I f the lawn
is not watered and if there is a prolonged period without rain, soil moisture
will decrease (outflows exceed inflows). Evapotranspiration will proceed
at the potential rate for some time, but ultimately the rate will drop. Why
does the rate drop when the atmosphere is capable of taking up water at the
potential rate? The explanation stems from the soil physical principles that
have been covered in this chapter. As water is pulled from the soil ncar a
plant root, the moisture content in the soil surrounding the root decreases.
Decreases in (} lead to large decreases in K. Therefore, by virtue of Darcy's
law, to maintain a steady flow of water to the plant root, the plant must
exert ever greater suctions (ever more negative capillary-pressure heads)
so that increases in the hydraulic gradient counterbalance the decreases in
K. At some point, the plant cannot sustain this battle with a drying soil and
the transpiration rate falls below the potential rate, as discussed in Chapter
9. The plants regulate the transpiration rate by adjusting the opening of the
stomata on the leaves (sec Section 2.4).
295
through clayey soil than through gravel! The key to understanding how
gravel layers can be capillary barriers to flow of water in the unsaturated
zone is appreciating that gravel has almost all "large" pores. Under
negative capillary-pressure heads, these large pores fill with air and
essentially stop the transport of water. (Water actually does move by vapor
diffusion, but this is an exceedingly slow process.) Clays, on the other
hand, have almost all "small" pores. Thus, under anything but extreme low
(negative) capillary-pressure heads, many of the pores will be filled with
water and will conduct water, albeit at slow rates (relative to saturated
gravel, but at very fast rates relative to vapor diffusion). The situation can
be appreciated by looking at the hydraulic conductivity curves for gravel
versus a clayey material (Figure 8 . 1 7). At saturation, Kgravcl » Kclay, but at
moisture contents not too far below saturation, Kclay » Kgravel' That is,
flows are impeded in the gravel relative to the clay at intennediate to low
values of the moisture content.
The recognition that coarse materials serve as barriers to flow in
unsaturated regimes has several practical applications. In desert regions,
cobbles can be used to mulch agricultural fields. During intense
thunderstonns (the usual method of delivery of water to the surface in
semiarid regions), the gravel layer is wet and readily allows water to
infiltrate. Once the surface gravel dries out in the following fair weather,
however, the evaporative flux of water from the surface is essentially
prevented because the hydraulic conductivity of dry gravel is very low.
The infiltrated water is available for use by plants for transpiration.
Another application is the construction of barriers to flow into waste
disposal trenches. Ifwastes are placed in the vadose zone, a gravel cap is
placed over the trench, and fine material is filled in over the gravel, the
gravel will act as a capillary barrier and deflect the flow of infiltrating
water through the fine material and away from the waste trench (under
unsaturated conditions).
296
K
(log
scale)
•
297
than 150 mm. The site is on a broad alluvial surface some 1 7 m above the
calculated level of the 100-year Ilood in nearby Homer Wash. The alluvial
and basin-fill deposits at the site arc about 600 m thick with the water table
about 225 m below ground surface.
In 1993, after contractors for the state of Califomia had spent more than
two years collecting data to characterize the proposed disposal site and
were ready to apply for licensing, several geologists questioned the safety
of the site for disposal of radioactive wastes. One of the concerns was the
potential for infiltrating waters to dissolve the wastes and recharge the
groundwater beneath the site at rates that would be too fast to allow
enough time for radioactive decay to render the wastes hannless. How can
the claims of (fairly rapid) water movement in the unsaturated zone be
evaluated? One set of measurements that would be useful is the vertical
distribution of capillary-pressure heads. Recall our discussion of the
equilibrium moisture profile above a water table. We concluded that, if
capillmy-pressure head declined with height above dalllm to exactly
balance the increase ill head due to gravity, there would be no movement
of water. That is, if the vertical gradicnt in hydraulic head is zcro, Darcy's
law indicates that Ilow is zero. Conversely, if the gradient in hydraulic
head is in the downward direction, recharge is indicated and if the gradient
is upward, discharge (due to evapotranspiration) is indicated. A committee
of the National Research Council concluded that the data for Ward Valley
were insufficient to estimate the gradient, but that data from similar arid
sites show upward gradients (NRC, 1995c). The committee concluded that
the dry state of the materials at the Ward Valley site did indicate that
vertical flow rates would be very slow. Detailed assessments of potential
waste-disposal sites in the vadose zone require careful measurements and
interpretation in light of the theory outlined in this chapter.
• Pore spaces in the unsaturated zone may be partly fillcd with water and
partly fil led with air. The volumetric moisture content, e, is a measure of
the amount of water held in a soil or rock. {Section 8.1 }
298
dampened. Water flows in this intennediate zone are approximately
steady. Near the water table, a saturated zone develops due to capillary
risc of water in the pores of the soil or rock. This saturated zone above
the water table is known as the capillary fringe. { Section 8. 1 }
• The gradient in hydraulic head drives flow in the unsaturated zone, just
as it does in the saturated zone. In the unsaturated zone the hydraulic
head is defined by II = If + z, where If is the (negative) capillary-pressure
head. {Section 8A}
• Darcy's law for flow in the unsaturated zone has the same fonn as it does
for flow in the saturated zone. For unsaturated flow, hydraulic
conductivity is a function of moisture content, so q = K(8)(dhldl).
-
{ Section 8A}
• Darcy's law can be used to infer the direction of water flow and the
magnitude of the specific discharge. If the gradient in hydraulic head is
positive, flow is downward. If the gradient in hydraulic head is negative,
flow is upward. The specific discharge is the product of K and the
gradient in hydraulic head. {Section 8.5 }
• For the case of zero specific discharge in the vertical in the unsaturated
zone above a static water table, the gradient in elevation head must be
exactly balanced by the gradient in capillary-pressure head . Thus, If = -z
under these conditions. This indicates that the equilibrium profile of
pressure head above a water table is a straight line upward to the left at
45°. The equilibrium profile of moisture content has the same shape as
the moisture characteristic. { Section 8.6}
299
• The Green-Amp! equation can be used to calculate the infiltration-time
curve if several soil parameters can be estimated. These parameters are
the salUrated hydraulic conductivity, the wetting-front pressure head, and
the change in moisture content between saturation and the underlying
soil. {Section 8.8. 1 }
Problem I. Tensiometers are installed at OAm and 0.5m above the water
table in a unifonn sandy soil with the moisture characteristic and hydraulic
conductivity curves given in Figures 8.4 and 8.5, respectively. One set of
tensiometer readings indicates that the capillary-pressure head at the first
of these tensiometers is -0.45 m and at the second is -0.6 m.
300
surface in this sandy loam to a depth of 50 m show a constant moisture
content of 3%. The water table is at a depth of80 m. Over the interval
of the moisture measurements, how do capillary pressure head and
hydraulic head vary?
Problem 3. Assume that a wetting front moves into the sandy loam soil
described in Problem 2. The moisture content at the surface of the soil is
held constant at 27%. The underlying moisture content is constant at 6%.
(fthe saturated hydraulic conductivity for this soil is 3 x 1 0-6 m S- I ,
estimate how long it will take the wetting front to move I m into the soil.
Fetter, C.W. 2000. Applied hydrogeology, 4th ed. Upper Saddle River, NJ:
Prentice Hall. Chapter 6.
301
9 Ecohydrology: Interactions between
Hydrological Processes and the Biota
9 . 1 Introduction
9.2.2 Waterlogging
9.1 Introduction
302
interactions between water and the biota. The study of these interactions
has led to the emergence of ecohydrology, a subdiscipline at the
confluence between ecology and hydrology. Research in this field is
investigating both the controls exerted by hydrologic processes on the
dynamics of terrestrial ecosystems (e.g., carbon sequestration,
biogeochemical cycles, plant and microbial stress), and the effect of the
biota on water flows and stocks. Understanding these hydrologic-biotic
interactions is crucial to the prediction of the effccts of changes in water
availability and climate on ecosystems, as well as the evaluation of land
use change impacts on the water cycle.
Most plants take up water through their roots. Water then moves through
the xylem, capillary tubes within the plant, to the leaves and is lost by the
plant as watcr vapor fluxcs (i.e., transpiration) through the stomata, small
cavities typically located beneath the leaves (Figure 9. 1 ). Plants also
assimilate CO2 from the atmosphere through the stomata during
photosynthesis. Plants can regulate the size of opening of the stomata to
control the rates oftranspiration and carbon assimilation. Thus, the water
and carbon cycles are strongly coupled through stomatal physiology.
303
Transpiration (qlrlUISP) �
-
•
I
/
Figure 9.1 Schematic representation of water flow through the soil-plant
atmosphere system. Plants take up water from the soil through their rools.
This water is transported through the xylem and released as water vapor
fluxes (transpiration) 10 the atmosphere through the stomata.
Classic theories of water transport in soil, roots, and xylem assume that the
flow is laminar. Soil pores and plant vessels are so tiny and the flow
velocities arc so small that the Reynolds numbers arc definitely within the
laminar range (see Chapter 3). Therefore, plant physiologists investigate
water flow in plants using Darcy's law, which assumes the flow to be
laminar (see Chapter 6). According to Darcy's law, the flow rate is
proportional to gradients of hydraulic head or-in h t e jargon of plant
physiologists-gradients of water potential, h [L]. The hydraulic head of
304
soil water (or soil water potential, lis) is the sum of elevation head (zs) and
soil capillary·pressurc hcad (\1'.\.), which in tum dcpcnds on the soil watcr
content through the moisture characteristic curves (Figure 8.4). Water flow
within plants is partly driven by concentration gradients across
semipemleable membranes; known as osmosis, this process entails the
transport of water molecules through semipenneable membranes from
lower to higher solute concentration areas (see Box 9.1). To account for
this phenomenon, the xylem water potential ("x) is expressed as the sum
of elevation head (zx), xylem capillary·pressure head (Wx), and the osmotic
potential (0), which is zero in pure water and negative in the presence of
solutes (0 decreases with increasing solute conccntrations). In a steady
flow the rate of root uptake is equal to the transpiration rate, q,rall!lp [L il]
(see Figure 9. 1 ), and can be expressed as
305
9.2). We recall from Chapter 8 that the relationship between I.fIs and the
waler content is expressed by the moisture characteristic curve (Figure
8.4): the capillary-pressure head IfIs of soil water decreases as the soil
becomes drier. During a dry period, plants might use osmotic
compensation to prevent turgidity loss and hydraulic failure, but below a
critical soil moisture value, 0*, they eventually have to decrease
transpiration by closing their stomata. As a result of stomala closure,
photosynthetic uptake is also reduced because CO2 fluxes occur through
the same stomata. Under these conditions, if the rate of carbon assimilation
is insufficient to meet the metabolic demand, plants need to use their
carbohydrate reserves. Prolonged droughts may lead to the depletion of
these reserves, thereby, causing plant mortality by carbon starvation
(McDowell et aI., 2008). Overall, as soil moisture decreases below the
critical value, 8*, plant transpiration is reduced, eventually becoming zcro
at a soil moisture level, 8.. , termed permanent wilting point (Figure 9. 2).
.
306
1 r
---
l-
w I
c..
- I
"
-
I
0 8
<P
• •
307
Interestingly, even though no semi-permeable membranes exist in
soils, in some clays there is a mechanism for segregating a relatively
ion-rich phasc from a relativcly ion-deficient phase. The surface of
clay particles tends to have a negative charge due to the structure of
the mineral. Because of the attraction of particles having opposite
electrical charges, an increased concentration of cations (positively
charged ions) is found in the space between clay particles. If there is
very dilute water (very low cation concentration) in the pore spaces
exterior to the clay plates, the increased ion concentration between
any two clay plates sustains an osmotic-pressure head gradient that
tends to force water in the direction of greater ion concentration;
that is, into the space between clay plates. In a fashion similar to the
osmometer (see figure), waler flow will continue until the osmotic
pressure is balanced by capillary pressure. This phenomenon
explains the expansion (shrinking) of some clays upon imbibition
(drying).
Salt solution
ii,
Semi-permeabl e
membrane
_3
.
r _______ 1i
Pure water
. . .
A dev1 ee for demonstratmg osmot1 c pressure.
9.2.2 \-\'atel'logging
Most plants need oxygen at their roots for metabolic processes. Because
the oxygen supply to the root zone occurs by diffusion through air-filled
soil pores, oxygen availability becomes restrietcd as soils bccome
saturated. Lack of oxygen makes saturated soils unsuitable for the growth
308
of many terrestrial plant species. Thus water stress in vegetation can be
caused by both a water deficit and by an excess of water (waterlogging).
Plants are sometimc classified on the basis of their tolerance of
waterlogging conditions. While most plants take up water from the
unsaturated zone, phreatophytes place part of their root system below the
water table and take up water from the groundwater. These plants rely on
watcr uptake from thc saturated zone but may still have only a limited
tolerance to waterlogging. Hydrophytes withstand prolonged periods in
which the ground surface is flooded.
The different sensitivity of plant species to flooding typically explains
their distribution along topographic gradients in flood prone areas.
Changes in hydrologic conditions may alter these distributions. For
instance, in the Florida Everglades, shifts in plant community composition
have been observed in response to altered hydrologic conditions. The
Everglades are an extensive wetland with a seasonal 60 km wide sheet
flow delivering freshwater from Lake Okeechobee to Florida Bay over a
distance of about 150 km. This slow shallow flow takes place on a
limestone substrate partly vegetated by herbaceous plants. The landscape
exhibits a number of islands that emerge from the watcr for most of the
year. Because of their higher elevation the islands are more infrequently
flooded than the surrounding marshes and are, therefore, densely covered
by trees and other woody plants with limited flood tolerance. In 1 948 the
U.S. Congress approvcd a project that led to the construction of canals and
levies that partitioned the Everglades into basins used for agriculture,
water storage (Water Conservation Areas, WCAs), and biological
conservation (the Everglades National Park). The WCAs were established
to retain some of this freshwater and make it available for drinking and
agricultural uscs during the dry season. Thc consequent increase in
flooding frequency in portions of the water conservation areas caused a
die-off of trees in the lower elevation parts of the islands and a complete
die-off in low islands.
Most plants keep their stomata closed at night and no transpiration occurs.
When vertical hydraulic head gradients exist along the soil profile, water is
redistributed through the roOI system from shallow 10 deep soil or vice
versa, depending on the direction of decreasing hydraulic head (see
Section 8.5) . Because of their higher hydraulic conductivity with respect to
unsaturated soils, plant roots provide a preferential pathway for the
nocturnal redistribution ofwaler (Caldwell and Richards, 1989). Irthe soil
309
is much drier at the surface than at depth (e.g., during a dry spell), then the
capillary-pressure head at the surface is much lower than in the deep soil;
this gradient in pressure head can overcome the opposite elevation head
gradient and bring deeper water up through the roots to the surface (Figure
9.3). Known as hydraulic lift, this process may facilitate the establishment
and survival of shallow-rooted plants and enhance the uptake of nutrients
stored in the shallow soil. The opposite process can also occur: right after a
rainfall event, if the deeper soil is sufficiently dry, the hydraulic gradient
from the shallow to the deep soil could draw water downward through the
root system (Figure 9.3). This phenomenon of hydraulic descent can be
used by deep-rooted plants to bring soil moisture out of the reach of
shallow rooted plants (Burgess et aI., 1 998). The occurrence of hydraulic
lift has been well documented in several forest ecosystems worldwide,
including the Amazon. Model simulations indicate that hydraulic lift may
enhance root uptake from deeper soil layers, and induce a substantial
increase in dry season transpiration and photosynthesis with important
impacts on the regional and global climate (Lee et aI., 2005).
tt tt
- -
310
9_2A SoilmoistUl-e control 011microbhll acth"it)'
Soil moisture afe f cts microbial processes in multiple ways. Low soil
moisture inhibits microbial activity because the substrate (carbon and
nutrients) used by microbes is transported in the water phase and its
mobility is strongly reduced in dry soils. Moreover, in hyperarid
(extremely dry) conditions microbes are affected by dehydration. Overall
the water stress tolerance of soil micro-organisms is much higher than for
plants; that is, the wilting point is much lower for microbes than for most
types of vegetation. As water content increases above the wilting point,
microbial activity increases because availability of substrate increases. The
activity increases until effects of waterlogging become important after
which increases in moisture lead to decreases in aerobic microbial activity
(Figure 9.4).
Thc sensitivity of microbes to water logging depends on the microbial
metabolism. While anaerobic microorganisms thrive in waterlogged soils
and anoxic conditions, the activity of aerobic communities relies on the
supply of oxygen and requires soil aeration. For example, the
decomposition of soil organic matter is perfonned by aerobic
microorganisms that convcrt littcr and othcr organic rcsiduals into
inorganic compounds while releasing CO, into the atmosphere. Water
logging conditions hinder decomposition, and provide an ideal
environment for the accumulation of organic matter. For instance, peat
deposits are commonly found in wetlands where the rate of organic matter
production exceeds the decomposition rate. It has been estimated that a
large fraction of global carbon stocks is in peat deposits. In boreal regions,
defined by very cold winters and relatively short and cool summers,
peatlands are prone to drainage as an effect of climate wanning. This
phenomenon is expected to increase soil aeration and soil organic matter
decomposition, thereby causing peat loss, decreases in ground surface
elevation (subsidence), and increases in atmospheric CO2 emissions.
3II
/ "\
.
> Substrate Oxygen
-
control control
-
.-
o
III
III
-
.c
.-
e
o
:2
.-
Soil moisture
Figure 9.4 A concepruaJ representation of soil moisture control on aerobic
microbial activity.
Modified after Skopp ct al. (1990).
At the beginning of this section we indicated that the carbon and water
cycles are strongly coupled. At this point it is clear that the coupling acts
both through the process of atmospheric carhon assimilation by plants
(hydrologic controls on stomatal regulation of photosynthetic uptake) and
through soil moisture controls on organic matter decomposition.
3 12
land masses increased by 243% and precipitation by 93%, while runoff
decreased by 24 % with respect to the unvcgctated case. These two
(unrealistic) extreme cases give us an idca of the important role played by
vegetation in the global water cycle.
In Chapter I we explained how the water received at the Earth's surface
as precipitation leaves continental landmasses either as evapotranspiration
or mnoff. Because mosl of the evapotranspiration from terrestrial
ecosystems is contributed by plants, this flux is sometimcs tenncd "green
watcr flow," while surface and groundwater mnoff arc referred to as "blue
water flows" (Figure 9.5 ) (Falkenmark et aI., 2006). To investigate the
impact of the biosphere on the water balance we can consider how
vegetation affects the partitioning ofprecipitalion over land into
evapotranspiration and runoff (i.e., into green and bluc watcr flows).
313
Green water flows
I
, I
iHH Hiii
Transpiration
I I I I I
Evaporation
I I I I I
• •
• • • •
• •
• • •
Surface runoff
• • •
• • •
OJ
). c
• • • •
'"
• •
�
• •
"
• • • ·
• • • • •
'"
-
• • • • •
• Groundwater flow �
0
• • • • • • • •
• • • • • • • • · •
• -
•
• • <- • • • • • • •
'00
• • + • • • • • • •
• • • • • •
• • , . . . . • • • • • • • • • • • •
· , .. . . . . , • • • • • • • • . . . . , , • •
• • • • • • • • • • • • • • • • • • • • • • • •
• • · . . . , . . . . . . . . . • • • • •
• · . . . , • • • " . . . • • •
• • • • • • • • • + • • • • • • + • • •
3 14
Higher LAI, higher roughness, Lower LAI, lower roughness,
deeper roots, lower albedo shallow roots, higher albedo
..
315
are in general more noticeable during the growing season, when vegetation
plays a more active role in the water balance.
•
40
::
•
•
0
c
c
•
30
�
�
-
• •
•
•
•
•
� 20 .. •
�
• •
•
c • • • •
c .. '
• • •
• •
• • •
�
•
1 •
••
I
• • . •
-
'
c • •
•
•
• • •
• •
• •
.
•
�
'"
• • •
• •
0
c
� •
�
u
-
-10 •
c
�
�
u -20
� •
11. •
Deforestation
-30 • Reforestation •
•
-40!;---c;;--;;:--;;;;-
o - 10 --:,:---c; ;---;;:--;;;;---:,:--:e ;---:!.
20 30 40 50 60 70 80 90 100
Percent change in cover
Figure 9.7 Effect of deforestation and reforestation on water yields
(expressed as a percentage of mean annual precipitation). A decrease in
forest cover is generally associated with an increase in runoff, while most
of the reforestation experiments exhibit a decrease in runoff.
Data from Andrcassian (2004).
316
afforestation by bush encroachment or tree plantations is recognized as a
serious threat to water yields and water availability, especially in semiarid
catchments.
317
water input and these inputs may play an important ecological role,
particularly if they occur during the dry scason (e.g., the wcst coast of the
U.S.) or in deserts (c.g., the Namibian or thc Atacama Desert). [t is worth
noticing Ihat occult precipitation would not occur (or would be severely
reduced) in the absence of vegetation canopies (Runyan et al., 2012). To
enhance their water supply, some communities living in "fog deserts" have
built systems of nets and gutters that trap atmospheric moisture imitating
the way plant canopies function in these environments.
In the previous sections we have seen both how plants depend on water
(Section 9.2) and how the rate of major hydrological processes can be
modified by vegetation (Section 9.3). We can now consider the case of
ecosystems in which, by altering hydrological processes, plants are able to
establish conditions that reduce their exposure to water stress caused either
by water scarcity or by waterlogging (Runyan et aI., 2012).
Fog deposition can be crucial to the survival of plant communities in
arid and semiarid ecosystems. This is an example of systems that rely on
mutual interactions between vegetation and hydrological processes. In fact,
the decrease in occult precipitation that would occur in response to the
removal of plant canopies could impede vegetation re-establishment.
Similarly, regional climate models have shown that vegetation removal
from Mongolia, the Sahel region of Africa, or other desert margins would
result in a reduction of precipitation, thereby preventing vegetation re
establishment (Bonan, 2008). Similar impacts of vegetation on
precipitation are often invoked to explain major climate and vegetation
changes in the history of the Earth. For example, fossil records indicate
that up 10 the mid-Holocene (i.e., about 5000 years ago) large parts oflhe
Sahara used to be vegetated. It has been argued that the decline of the
"Green Sahara" was sustained by a decrease in precipitation induced by
the loss of vegetation cover. Although changes in orbital conditions appear
to have played a major role in the shift to the desert state, paleoclimate
models indicate that the decrease in precipitation was likely enhanced by
vegetation losses (Brovkin et aI., 1 998).
Analogous conditions may result from plant-water table interactions.
The typical effect of vegetation on the underlying shallow unconfined
aquifer is to lower the watcr table by taking up water from the overlying
unsaturated zone or through direct depletion by phreatophytic vegetation.
In landscapes with shallow water tables vegetation may play a crucial role
318
in draining the shallow soil, thereby preventing the emergence of water
logging conditions that would be detrimental to the survival of the existing
plant species. Some ccosystcms may strongly depend on these positive
interactions between vegetation and water table dynamics. In some cases,
the removal of native vegetation may result in a dramatic increase in water
table elevation. For example, the remov<ll of eucalyptus trees from the
Murray-Darling basin (Australia) has allowed the water table to fise so
close to the ground surface that direct evaporation of shallow groundwater
is causing salt accumulation at the soil surface, thereby making soils
unsuitable for most plants (Runyan et aI., 2012).
These are just some examples of ecosystems that strongly depend on
mutual interactions between hydrological processes and the biota.
Understanding these interactions is crucial the slUdy of ecosystem
sensitivity to changes in climate and hydrologic conditions.
• Water stress conditions can emerge not only when the ground is dry but
also in saturated soils (waterlogging). Oxygen is needed in the root zone
to sustain plant metabolic processes. Because oxygen is supplied by
diffusion through air-filled soil pores, its availability is restricted in
319
saturated soils. Thus, most plant species do not tolerate waterlogging.
Similarly microbial reactions that require oxygen (e.g., decomposition)
arc restricted in saturated soils. {Sections 9.2.2 and 9.2.4}
320
may be crucial to the survival of some plant communities in water scarce
environments. In these ecosystems vegetation removal can dramatically
alter the water budget to thc point of preventing the re-establishmenl of
the same plant species. {Section 9.4}
d8
Z-=-et,
dl
where Z is the depth of the root zone. I n other words, changes in the water
stored in the control volume (i.e., the root zone) are due to the difference
between water inputs and outputs (see Chapter I) . I n this specific case
there are no rainfall-induced inputs, while the only outflow is due to el.
The irrigation period (i.e., the time between two consecutive applications
of irrigation water) is calculated by integrating the soil water balance
equation between time I = 0 when 0 = O" (i.e., right after irrigation) and
time Ii when 0 = 00 (and it is time to irrigate again), assuming that the
f<
Consider the case of a crop grown on a sandy soil with 61lc = 0.19 and a
crop-specific wilting point, 61 = 0.05. The root zone is 40 cm deep. After
....
32 1
would take for soil moisture to decrease from field capacity to the value (}o
= 2(}", in the absence of any rainrail input.
322
10 Catchment Hydrology: The HiIIslope
Stream Continuum
10.1 Introduction
10.2 Streamflow Hydrographs
10.3 Hydrograph Separation
10.4 Runoff Processes
10.4.1 Direct precipitation onto stream channels
10.4.2 Overland flow
10A.3 Shallow subsurface stonntlow
10.4.4 Groundwater flow
10.4.5 "Old" and "new" water
10.5 Contributing Area and Topographic Controls on
Saturation
10.5.1 Contributing area
10.5.2 Topographic index
10.5.3 Hillslope stability
10.6 Routing Water through a Catchment Using Catchment
Models
10.1 Introduction
323
One of the problems that has occupied hydrologists for many decades is
the identification and quantification of the pathways by which rainfall and
snowmelt movc through catchments and ultimatcly producc runoff. In
Chapters 5, 7, and 8 we dcscribed the individual flow paths over and
beneath the Earth's surface that precipitation might take through a
catchment and provided the physical basis for understanding each. To
describc the catchmcnt hydrological cycle, we have to intcgrate all the
individual processcs of surface and subsurface flow of water. This is a
fonnidable problem to solve in a quantitative fashion. In this chapter we
introduce some of the ideas and tools used in catchment hydrology.
There are many reasons why hydrologists are interested in
understanding flow paths within catchments. One reason is that
understanding the routing of rainfall and snowmelt is required for
undcrstanding how water interacts chemically with rocks, sediments, and
biota within a catchment. For example, hydrologists are interested in
understanding the effects of acid rain on the chemical composition of
streams around the world. The pH of rainfall and snowfall in many parts of
North America and Europe in the latter part of the twentieth century was
typically below 4.5, which is well below the level at which water would be
considered to be acidic « 5.65). "Acid rain" is known to result from the
chemical interaction of water in the atmosphere with gaseous emissions
from power plants, factories, and automobiles. (Although the trend in the
acidity of rain has been positivc ovcr the past decade or so and the extent
of areas with rainfall pH below 4.5 has decreased considerably, concerns
remain about the impact of deposition of acidic compounds on catchments
(Burns et al., 2011 ).) In some catchments, the stream pH shows an
episodic depression (known as acidification) that at first glance might
appear to be caused by rain falling directly on the Sfream surface.
However, for many streams this explanation cannot be correct because the
pH of the stream remains low even after the rainfall has ceased (for
example, Bear Brook in Maine, Figure 1 0. 1 ). How, then, is the
composition of the rain altered when it comes into contact with soils and
rocks? Answers to such questions require knowledge of the flow paths
through and residence times within various portions of the catchment, and
are relevant to a variety of environmental issues. For example, depressions
of pH below about 5 may cause damage to certain types of fish.
Another reason for desiring an understanding of runoff generation in
catchments is to be able to predict the effects that humans will have on the
hydrological cyclc through their land-based activities. It is now known
from experimental studies at research sites throughout the world (for
example, the Hubbard Brook Experimental Forest in New Hampshire) that
324
various silvicultural practices (such as selective cutting, clear-cutting, and
others) can have dramatic effects on stream-water discharge and quality
for many years after the particular managcmcnt activity is performed.
Hydrological changes that have been demonstrated experimentally include
increases in annual water yields, decreases in evapotranspiration, increases
in nitrate concentrations, and increases in stream temperature and sediment
loads.
::-- 0.04 65
• 0
E 0.03 6
8, 0.02
0
"'- 5.5 a
0
"
•
0.01 o 00 0 0 5
•
•
u
C 4.5
re re re
0
� � � , ,
- - - - - - -
8. Ii 8 8 8 8 8. 8
N N N N N
.
E E § E � E § E E §
� ,; z ci
:§
• z ci
f •
z
$ $ $ $ $
:2 �
�
• � �
• � �
• �
325
the flow during precipitation events occurs immediately after the onset of
precipitation, while other portions of the now occur for some time after the
end of precipitation. Thesc hvo components of a streamnow hydrograph
are known as quickflow and baseflow.
The separation of a hydrograph into two components suggests that
water is being routed through two different storage "reservoirs" (Figure
1 0.3). That is, considering the input to a reservoir to be precipitation and
the outflow from the reservoir 10 be either the quickflow or the baseflow,
we observe the typical effects of flow routing through a reservoir: peaks
are delayed and attenuated. From our discussion of flow routing through
surface impoundments (Chapter 5.4), we can infer by analogy that the
baseflow reservoir has characteristics of storage and release that are quite
different from those of the quickflow reservoir. Remember that these
catchment "reservoirs" are conceptual, meant 10 represent portions of the
hydrological system that behave similarly to surface-water reservoirs, by
storing and releasing water.
E
-
g
-
;;.
•
�
�
�
8 � � :. •
8
,
f ! .!
- - - - - - - - - - - - -
,
"
•
• •
,
•
" < "
"
"
•
<
u
,
•
•
�
•
0
0
8 , •
• � •
� 8
- - - - - - - - - - - - -
, ,
"
•
" , "
• • > • • "
• •
z c , <
Figure 10.2 Hyetograph (a) and hydrograph (b) for Coweeta W34, water
year 1983.
326
TIme
•
.'l
E
+
�
•
-
- •
•
� 00
-
'"
L-J....---:Ti
me
� Time
lL '\-'-
'-
�
Time
Figure 10.3 Rainfall routed through two conceptual "reservoirs."
Although depicted here as surface-water reservoirs, in this chapter the term
reservoir refers to portions of the catchment hydrological system with
differing dynamic behavior. For example, the base flow rescrvoir may be
groundwater; because groundwater recharge and discharge arc relatively
slow processes, this reservoir behaves like the one at the right above.
327
10.3 Hydrograph Separation
Questions such as those posed in the previous section have nOl been
answered completely despite many decades of research. One of the
difficulties lies in the detennination of the components of the streamflow
hydrograph. After all, we don't measure two hydrographs (quickflow and
basef1ow); we measure only the total flow. The question then arises: how
can we "separate" the measured hydrograph? The smooth line in Figure
10.2 showing the baseflow was drawn "by eye." The separation seems
reasonable, but has no scientific basis. Hydrograph separation techniques
have been used by hydrologists for many years, but all of the methods are
empirical. These empirical methods are used widely to enable the
calculation of quickflow from catchments. Once the hydrograph has been
separated into quickflow and baseflow, a very simple computation, known
as the unit hydrograph method, can be used to route the portion of the
precipitation that goes through the quickflow "reservoir" to the stream, a
procedure that has application in engineering hydrology where the size of
pipes to carry stormwater is decided on the basis of such routing
calculations.
For many years, a standard practice was to assume that the components
derived from empirical hydrograph separation methods were physically
"real"-that the quickflow arose from the flow of water overland where
precipitation rates exceeded infiltration rates and that the basellow
originated as water that had infiltrated the soil surface, percolated to the
groundwater zone, and discharged into a stream channel over prolonged
periods of time. The "reservoirs" of Figure 10.3 under this view would be
the hillslope surfaces (the quickflow reservoir) and the unconfined aquifer
(the baseflow reservoir). Unfortunately, this simple picture is not valid in
many instances. The surface-water-groundwater-reservoir interpretation of
the empirical results turns out to be counter to observations made by field
hydrologists in forested catchments, for example.
An altemate method for separating hydrographs is based on differences
in the chemical composition of waters in the " reservoirs" thought to store
and release water (e.g., see Sklash and Farvolden, 1979). The assumpl"ion
behind this alternate method for separating hydrographs is that streamflow
during a storm event is a mixture of water that has been precipitated into
the catchment during the particular stonn (i.e., event rainfall or "new"
water) and water that was stored in the catchment prior to the onset of the
storm (i.e., pre-event or "old" water). Provided that the concentration of
some chemical component (i.e., tracer) of these two waters nalUrally
differs (and does not vary during the storm or val)' spatially across the
328
catchment), measurements of the concentration of the tracer in the two
components and in the mixture through time can be used to back-calculate
the contributions from the "old" and " new" reservoirs. For example,
suppose we make measurements of the chloride concentration in the rain
falling on the catchment, in the groundwater within the catchment, and in
streamflow (the mixture) during the course of a stoml. Let Q, be the total
measured streamflow (measured at a standard stream gage). Let Q" and Q"
be the flow contributions to the hydrograph associated with the "new"
water (rain) and the "old" water, respectively. Because we assume that
there are no other flow components, at any time t we know that
Q, = Q,, +Q
•. (10.1)
Note that the streamflow (Q,) and the flow components (Qn and Qol vary
with time.
If the concentration of chloride in precipitation (rain or snowmelt) is
relatively constant over the storm and if the concentration of chloride in
"old" water in the catchment is relatively constant in both time and space,
we can rcprcsent thcse concentrmions as C" and C" (note that we havc
measurements of these concentrations). Because the stream water is a
mixture of the old and the new waters, thc concentration of chloride in the
stream will change over the course of the stonn. We represent this
temporally changing concentration as Ct. Because we assume that the mass
of chloride observed in the stream is derived from only these two
reservoirs, we have:
Q,C, Q,,c,, + Q"C".
= (10.2)
Q =
[
(C'-C.)
" (C�-C")
Q.
] I
(10.3)
329
Hollow catchment, Shenandoah National Park, Virginia. Bazemore (1993)
made hydrological measurements and chemical observations of a 0.082-
km2 sub-catchment at Shaver Hollow. The pre-event groundwater had a
chloride concentration of about 27 mol L-I and the event water that fell
through the forest canopy had a chloride concentration of about 4 mol L- I .
Chloride concentrations in the stream, measured over the storm
hydrograph, were used to separate the hydrograph into "old" and "new"
water (Figure lOA). Note that approximately 90% of the water under the
storm hydrograph was detennined to be "old" (pre-event) water and at the
time of peak discharge more than 80% of the total discharge was
detennined to be old water (Bazemore, 1 993). The use of chemical
composition of throughfall, soil water, and ground water to separate
hydrographs into components requircs a numbcr of assumptions about the
temporal and spatial variability in the various "components." The method
can be criticized because the assumptions may be poor for natural
catchments (e.g., see Harris et aI., 1995). Despite controversy about the
details of chemical hydrograph separations, the frequent observation from
such work that a large fraction of stormflow in forested, temperate-zone
catchments is "old water" is generally agreed to be valid. This approach of
using chemical signatures to identifY contributions to streamflow from
several sources can be useful for a variety of problems and is generally
termed end-member mixing analysis.
� .-------r �
30
�� 25
-' 00 0 0 Chloride
o
; 20 0
-
0 00
a
�•� 15 000
o
"Old" water '-'
discharge
.-
C
10
5
o +-�...,.,..�-,.....""
noon 6:00 p.m. midnight 6:00 a.m. noon 6:00 p.m. midnight
Figure l OA Chemical hydrograph separation for a sub-catchmcnt of
Shaver Hollow, Shenandoah National Park, Virginia. The data cover June
4 and 5, 1 992.
330
Data courtesy of David Bazemore.
The main conclusion from the application of thc chcmical·hydrograph·
separation method to this and other forested, upland catchments located in
humid, temperate regions is that observed streamflow is mostly old water.
Quickflow is not synonymous with "new" water and baseflow is not
synonymous with "old" water. An empirical separation of the hydrograph
into quickflow and baseflow is insufficient to describe actual now paths
within a catchment, which are necessary, for example, to account for
chemical reactions in a catchment. Describing the transport of solutes
within a catchment requires a detailed understanding of the physical
processes that govern the flow of water in catchments.
331
(a)
I I
I
/
-
Groundwater
discharge
Pre c i p i t a t i o n
(c) � � � �
I I
I
-
/
Groundwater
discharge
332
saturated areas (b). Saturation of penneable soil horizons also can produce
shallow lateral flow called subsurface stonnflow (e).
Overland flow is water that flows aeross the ground surface and discharges
into a stream channel. If overland flow is to occur, water must accumulate
at the surface rather than infiltrate into the soil. Overland flow can occur
on a catchment for a number of reasons: (a) the catchment surface may be
nearly impermeable due to the presence of exposed bedrock or paving over
the surface; (b) the instantaneous rate of infiltration through the pervious
surface may be exceeded by the instantaneous rate of rainfall (or
snowmelt) onto the catchment surface, causing ponding of water at the
surface; and (c) the catchment soil upon which the rainfall is precipitated
may be saturated to the soil surface, causing ponding because the
precipitated water cannot infiltrate into an already saturated soil. Overland
flow in catchments is one of the most rapid paths that rainfall can follow to
the stream channel; therefore, overland flow is expected to contribute to
quickflow.
Overland flow dominates the response of many urbanized catchments
with large expanses of directly connected, impervious surfaces. Because
urbanization increases the fraction ofprccipitated watcr that follows the
rapid overland flow path through a catchment, incrcased frequency of
floods is one result of development. Leopold (1 994) reports that after
suburban development of a small catchment in Maryland, flood
frequencies on Watts Branch, the stream draining the catchment, increased
dramatically. Overbank flows in Watts Branch occurred about 1 . 4 times
per year before urbanization and about 6.5 times per year afterwards.
Overland flow of water from a catchment that occurs when the rainfall
or snowmelt rate exceeds the ability of the soil to allow water to infiltrate
is referred to as infLItration-excess overland flow, or Hortonian
overland now, thereby crediting the American hydrologist Robert Horton,
who reported occurrences of this runoff mechanism. Horton ( 1933) argued
that, over the time course of a stonn, the rainfall rate eventually would
exceed the rate at which water could infiltrate into the soils ofa catchment
(as in Figure 8.9), resulting in the accumulation of excess water at the
ground surface and overland flow in sheets. As discussed in Chapter 8,
infiltration into soils depends on hydraulic conductivity of the soil profiIe.
The hydraulic conductivity of the surface soil is conditioned strongly by
vegetation as well as by the texture of the mineral soil. For example,
hydrologists have observed that forest soils typically have a high
333
infi Itration capacity duc to thc presence of above-ground vegetation (and
decomposing vegetation on the surface), which protects the surface from
compaction caused by raindrop impact, and ofbclow-ground vegctation,
which provides roots that keep the soil porous and highly penneable
(Chapter 9.3.2). Infiltration-excess overland flow is not a significant
runoff-gcnerating process in such catchments. For example, up to 150 mm
of water per hour was sprinkled for several hours on an isolated block of
soil in a forested catchment in Maine with essentially no overland flow
(Hornberger et al., 1990). Conversely, when vegetation is absent from the
surface of a catchment, the surface soil tends to develop a "crust" with
relatively low hydraulic conductivity. Infiltration-excess overland flow can
be the dominant runoff-generating mechanism in catchments where the
land surface has been strongly disturbed (e.g., a plowed agricultural field
or construction site where vegetation was removed and the soil surface left
exposed) and in arid and semiarid regions with sparse vegetative cover.
Infiltration-excess overland flow generally does not occur at a constant
rate over entire catchments due to spatial differences in soil properties.
Some areas of the catchment may produce overland flow during virtually
all storms, while other areas may rarely, if ever, do so.
In many forested catchments in the temperate zone, the soil is capable
of allowing essentially all of the incident precipitation to infiltrate. I n such
catchments overland flow develops not because precipitation intensity
exceeds the infiltration rate, but because it falls on temporarily or
permanently saturated areas (wetlands) with no capacity for water to
infiltrate. Flow developing under these conditions is known as saturation
excess overland flow (Figure I O.5b). The areas ofa catchment that are
prone to saturation tend to be near the stream channels, or where
groundwater discharge areas occur. These areas grow in size during a
stonn and shrink during extended dry periods (Dunne et aI., 1975; Beven,
1978). The areas on which saturation-excess overland flow develops,
therefore, change with time over a stonn. Because the expansion of
saturated areas occurs in a similar fashion as the expansion of the channel
network, this mechanism of runoff generation is often referred to as the
variable contributing area concept.
334
downward has thus been observed 10 "pile up," causing localized areas of
saturation in the soil. In these instances, water may move laterally toward a
stream by a process known as shallow subsurface stormOow (Figure
1O.5c). Some of the water in subsurface stormflow moves at a relatively
slow pace through the soil and contributes to the baseflow of streams,
particularly during wetter winter and spring periods. Subsurface stormflow
also may occur along prefcrrcd flow pathways (c.g., soil cracks, old animal
burrows, decayed root channcls, etc.) called macropores (Chapter 8) and
the flow to the stream channel along these pathways may be quite rapid.
Thus, subsurface stormflow can contribute significantly to quickflow. A
portion of the infiltmted water does not become subsurface stormflow; it
percolates downward until it reachcs the watcr table and becomes
groundwater.
335
of a hydrograph) has the chemical composition of water that was resident
in the subsurface prior to the storm event. That is, the bulk of quickflow is
"old" water, indicating that watcr contained in soils and rocks in a
catchment must be able to reach the stream quickly under stonn
conditions. This observation may seem at first glance to be at odds with
the physical processes described above in which "old" groundw<lter is
viewed as contributing primarily to the (slow) baseflow portion of the
hydrograph. There arc a number of physical processes that can be called
upon to explain the observation that a large fraction of quickflow is "old"
water. Consideration of the physics of soil water suggests that processes
that cause "new" water delivered to the surface of a catchment to displace
"old" water and force it into thc stream are common. For example, even if
only a small amount of infiltrated water reaches the capillm)' fringe
quickly (e.g., through flow in macropores, especially in the riparian, or
near-stream, are<l), it can change the negative capillary-pressure head in
the capillary fringe to a positive pressure head. Such rapid increases in
positive pressure head can force " old" groundwater into the stream rapidly.
Another mechanism that contributes to the delivery of "old" water to a
stream under storm conditions is the entrainment of "old" soil water by
overland flow. Overland flow in riparian areas occurs in patches, with
surface water flowing into the soil, mixing with shallow soil water, and
then reemerging on the surface. The resulting overland flow contains a
portion of the "old" soil water with which the "new" water has mixed. The
conceptual explanations offered above appear to be consistent with
observations, but by themselves do not provide quantitative estimates for
the various flows. Quantification requires consideration of inflow, outflow,
and storage ofwatcr within a catchmcnt.
336
of the blocks in a catchment could provide the starting point for routing
water through the catchment.
337
10.5.2 Topographic iude\:
TI = In(altanlJ). (10.5)
<a>
Figure 10.6 Local slope and contributing area the water balance
for a catchment "block." The inflow rate is proportional to the contributing
area A, which depends on how long the hillslope is as well as whether it is
convergent, divergent, or planar (a). The local slope controls the outflow
338
from the blocks (b). lfinflow is smaller than outflow (lIpper left in b), the
water table declines. Conversely, if inflow is greater than outflow ({olver
right ill b), Ihe water table will rise and surface saturation may occur.
Specific contributing area, a = Ale, can be defined for each point within a
catchment if the topography ofa catchment is known. Although it would
be possible to estimate a from a high-resolution topographic map, most
studies of this sort use digital elevation data that can be used in
conjunction with gcographic infonnation system (GIS) software to
detennine the specific contributing area for each point in the catchment. It
is important to note that the results of an analysis such as this are highly
dependent on the quality and resolution of the digital elevation data.
Accurate identification of the channel network, in particular, depends on
using high-resolution elevation data. The current state of the art technique
for generating high-resolution topographic data is light detection and
ranging ([idar). In fact, the critical need for good topographic data to
define channels and channel networks has led to calls for lidar maps 10 be
produced to allow accurate mapping of flood plains (NRC, 2007).
339
C I > c2 Convergent
< c2
CI
Divergent
C1 = c
2
Planar
Figure 10.7 Influence of topography on contributing area.
340
Topographic index
In("/tan�) 5 6 7 8 9 10 11 12 13 14 15
(b) 0.14
•
•
n:I 0.12
"
•
0
0
� 0.1
�
u
o 0.08
-
t;-
o
II) 0.06
,
�
-!:: 0.04
•
•
>
III 0.02
�
•
a:
O 2:---4- 10
6 8 12 14 16
index
Topographic
Figure 10.8 Topographic indices for a catchment in Shenandoah National
Park. The spatial pattern (a) indicates a likelihood of saturation in the
central valley of the catchment. Thc distribution of values (b) is uscd in
TOPMODEL calculations as described in Section 1 0.6.
The hillslopes of ca\chments are not static feafUres of the landscape. Over
time, hillslopes change slowly due to erosion by runoff and rapidly due to
341
landslides (also known as debris flows). Landsliding on steep hillslopes
typically occurs in localized areas where the soils are saturated and the
water pressurc in the porc spaces is high. Surface runoff most often occurs
when soils become saturated so that precipitation can no longer infiltrate
(saturation-excess overland flow). If the surface runoff is deep enough
and/or the slope steep enough, the flow can dislodge and carry soil
particles from the hillslope to the channel, resulting in erosion of the
hillslope.
Dietrich ct al. ( 1 992) combincd simple cxpressions describing the
thresholds of saturation-excess overland flow, landsliding, and hillslope
erosion with detailed digital elevation data and careful field observations
to predict localions within a catchment where each of these processes
dominates. Assuming a constant transmissivity T of the surface soil layer,
the saturated subsurface soil discharge across a contour line of Icngth e is
given by Darcy's law as
The water-table slope is assumed to be equal to the surface slope. The total
amount of watcr reaching the length of contour (e) over a specified period
of time is AqlOloi> where qlOlal = R (the recharge rate, [L liD and A is the
upslopc contributing arca. In othcr words, q'QIaI is the volume of watcr pcr
unit surface area (or depth) that is moving through the hillslope per unit
time. The dilTerence between total runoff past a contour interval (qlOl(ll) and
saturated subsurface discharge (qs"bsllrjilce) is saturation-excess overland
flow. Thus, overland flow occurs when:
0'
A T
> 1M P (10.8)
c q.u
-
Erosion by overland flow will only occur in the parts of the catchment
where the overland flow is deep enough (large specific contributing area
Alc) or the slope is steep enough (large tan fJ) for the flow to dislodge the
soil grains. Dietrich et al. ( 1 992) proposed the following expression for the
erosion threshold:
A T a
-> lan/3+ .
Atan/3F
( 10.9)
c qM,w q.,•
•
342
where a [L2 II] characterizes the resistance of the soil to erosion.
Cohesionless material on a sloping surface becomes unstable, leading to
shallow landsliding, when the slope of the surface exceeds a critical value
dependent on the soil and water properties and the degree of saturation
described by:
343
The question thai we address in this section is how precipitation can be
rouled through a catchment to calculate streamflow. The most
straightforward way to usc the thcory developcd in prcvious chapters to
solve the catchment routing problem is to link together equations for
overland flow (e.g., Manning's equation), for flow in the unsaturated zone
(e.g., Richards' equation), and for flow in the water-table aquifer using an
equation for groundwatcr flow. The earlicst application of such a modcl
was presented by Freeze (1 9 7 1, 1972a, I 972b), who examined thc runoff
responses of a hypothetical hillslope to precipitation inputs. More recently
with advances in computational power, this approach has been used
successfully to simulate flows through catchments in great detail. For
examplc, the Pcnn Slate Integrated Hydrologic Model (PIHM, Qu and
Duffy, 2007; www.pihm.psu.edu) has bcen applicd at sevcral catchmcnts
to study coupled hydrological-biogeochemical processes.
, o' �
,
E
�
�
ti '
..
..
�
'"
c
::1 ,
.�
.c
'"
.�
c
-
0 -
c::
... ..,
... In
-
-
..,
.�
... c::
.�
..
Co �
(/) No surface runoff
,
1 0 -1 ,
Slope, Ian P
344
Figure 10.9 Regions of saturation overland flow, erosion, and landsliding.
345
Evapotranspiration Precipitation
t t
I Interception reservoir I
Throughfall
""\
I' '\
,
-
-
-
,
--, "
"- ,
\ " ,--� �) ,-' , ,
"
Soil
-
'
- "
' " .
reservo" , " \
-
" '
�
,
"
,,
, " ,
,
'
,
, ,
" , ,
���,
,
" '
, '
,
. .
Figure 10.10 Schemallc diagram of the TOPMODEL concept.
346
drain the water downslope is limited (where slopes flatten at the base of
hollows). Conservation of mass can be applied to the segment depicted in
Figure 1 0. 1 1 to detennine the fluxes.
A
. ..
. .
Asat
/
(10.1 1 )
where q/olll/ is total streamflow. It has dimensions of[L il] (discharge [LJ
i l] divided by area [L2]); all of the flow quantities in TOPMODEL have
these units. The surface flow contribution is qllwrillll</ and q"ubs"r/<.ee is the
subsurface flow contribution (Figure 1 0. 1 1 ).
Surface flow is generated when precipitation falls on a saturated area
and from return flow, so:
(10.12)
347
where Asa/A is the fraction of the hillslope area that is saturated (Figure
1 0. 1 1 ), P [L [1] is the throughfall or snowmelt rate, and (jrellirn [L [1 ] is
the return flow.
We calculate (j$llbsur!l1u [L [ I] from total subsurface discharge Q"Ub.51"il1ce
= Tc tan f3 [L3 [11 (Equation 1 0.6), where T is the transmissivity of the
soil [L2 [lJ, c [L] is the contour width (length perpendicular to the flow
direction), and tan f3 is the slope. The transmissivity is equal to the soil
depth multiplicd by thc soil hydraulic conductivity. Note that the slope of
the water table is assumed to be the same as that of the land surface.
We assume that the saturated hydraulic conductivity of the soil
decreases with soil depth exponentially, a situation often observed:
(10.13)
(10.14)
The term e-jD is generally much smaller than the term e-ft, so Equation
1 0. 1 4 can be simplified:
T= fe-f
K
t. (10.15)
Combining Equations 10.6 and 10. 1 5 gives the following equation for
subsurface flow:
(10.16)
348
Q,....�,
.. = ; -J�fclan/l
K
t? (10.17)
(10.18)
(10.19)
where R [L II] is the recharge rate and A is the area of the hills10pe slice
-the section of hillslope that drains past the section of contour (c) in
question. The "great leap" of TOPMODEL is to assume steady-state
conditions. Then, Qsubswface = QR' or
•
RA 0:= r.....
,e--;;crnnfJ. (10.20)
s O:= -lIItn ( ) ( )
r.
R
....,
-mln a
Ian
fJ
. (10.21)
where (I = Ale, the specific contributing area. The second tenn on the right
of Equation 1 0.2 1 describes the way in which topography controls the
propensity for every point in the catchment to reach sahlration (i.e., the
propensity of each point to generate saturation-excess overland flow
during stonns). Ifs is less than or equal to zero, the soil is saturated. From
Equation 1 0.2 1 we see that this occurs most easily for points within the
catchment where the topographic index (TI = In(aitan {f)) is large.
Until this point in the discussion, we have been referring to an
individual catchment hills10pe, or hi11slope segment, defined by a pair of
streamlines and extending from the stream to the catchment divide (Figure
1 0. 1 1 ). However, we could also consider any point in the catchment and
349
calculate the upslope contributing area and the local slope. In this way, we
can compute the distribution of topographic indices for the entire
catchment. In practice, the computations arc often done for "blocks"
delineated on the basis of DEM (Digital Elevation Model) or surveying
data. The topographic index, and therefore the contributions of surface and
subsurface flow to streamflow, can be calculated for each block. The
saturation deficit can also be calculated for each block, using Equalion
10.21 . Furthennore, these quantities will be identical for two blocks with
the same topographic index, as long as R and Tm,u: are spatially constant.
To solve for the catchment-avcrage saturation deficit (x), we can
integrate Equation 10.21 over the catchment and divide by the area. Here
we assume that R and Tmax are constant over the catchment:
(10.22)
where /.. is the mean In(a/tan fi) for the catchment. Combining Equations
10.21 and 10.22 gives:
(10.23)
This equation states that the saturation deficit at any point in a catchment
is equal to the average saturation deficit for the catchment plus a soil
parameter, III , times the diffcrence between the average topographic indcx
and the local topographic index.
Now we have a way to calculate Am/A-compute s at any point and
check to see if it is less than or equal to zero. We can estimate 11/ from soil
characteristics, calculate /.. and In(altan P) from a topographic map, and can
keep track of .'II by water-balance accounting
(p, interception, el,
subsurface flow, and overland flow). Ifs < 0, the soil is completely
saturated and any rain on the surface will become overland flow. The rate
of flow produced by this mechanism is detem1ined lIsing the throughfall
intensity and fractional catchment area that is saturated (e.g., Equation
10. 1 2). Return flow occurs where s < 0, and the rate of return flow is equal
to IsIAs,,/A .
Next we develop an expression for the mean subsurface discharge, q
subsurface' Integrating Equation 10. 1 8 over the catchment area and dividing
the result by the catchment area yields:
( 1 0.24)
350
Thus, in TOPMODEL, subsurface flow is controlled by soil characteristics
CTma:< and m), topography CA.), and the average saturation deficit of the
catchment.
This is not all of TOPMODEL, but it is a summary of the main
conceptual points. Added to the formulations above are the other standard
components of the watcr budgct-evapotranspiration, snowmelt, channel
routing (e.g., a simple reservoir routing method). All of the water-balance
accounting parts of the model are simple applications of the conservation
of mass. A fuller description of TOPMODEL is available in Wolock
( 1993 ).
351
Simulated
Observed
!.,
-
-
.,
."
E
E
'"
-
�
•
'"
.,
�
.r=
<.>
'"
.-
.g (ij )5. � § � a. U
O E; �
� �
I::
� < �
::J
� � • � Z 0
Figure 10.12 Observed discharge ( 1984) in the Snake River, Colorado,
compared with TOPMODEL calculations.
Data courtesy of Ken Bencala and Diane McKnight.
352
through a catchment and discussed some of the complexities in the
processes that give rise to observed streamflow dynamics. The advances
that have been made in calchment modeling have been substantial, but the
state of the science is still inadequate to allow acceptable routing for some
of the most serious problems (Hornberger and Boyer, 1995). One example
is predicting the flow paths of natural solutes and contaminants in
catchments, an example of which is the acidification of streams during
stonns observed in the northeastern United States (Figure 1 0. 1 ) and
elsewhere in the world.
Models such as TOPMODEL are necessary for calculation of
acidification and of the movement of other chemicals through a catchment.
A mass-balance model is required to do an accounting of chemicals, in
addition to an accounting of waler. Robson et al. ( 1 992) combined
TOPMODEL with a simple chemical mixing model 10 describe the
episodic acidification observed in a stream in Wales. This is an example of
the approaches that are being used to study a variety of pollution problems
in catchments. The transport of chemicals through a catchment is
controlled to a great extent by interactions between the water and the soil
and rocks in the catchment. Detennination of the flows through the soil
and groundwater reservoirs in the catchment is critical for realistic
predictions of the fate of pollutants.
TOPMODEL calculates not only the stream hydrograph but
infonnation that is useful for linking hydrological calculations to
hydrochemical models (Cosby el al., 1987). For example, TOPMODEL
calculates subsurface flow, overland flow, and saturation deficit (depth to
water table), quantities that can be used to detennine how waters from
different parts of a catchment mix to produce the chemical composition
observed in streams. As a final illustration, we look once again at the
Snakc River in Colorado.
In many lakc and strcam systems, the concentration and composition of
dissolved organic material (DOM) is a critical water quality characteristic.
One example of a process controlled by DOM is the formation of
trihalomethanes, compounds that are known carcinogens, in a drinking
water supply as a result of interactions between chlorine (used to treat
drinking watcr in public supplies) and components of the DOM during
water treatment. The DOM also can have indirect effects on water quality
by influencing internal processes of aquatic ecosystems, such as
photosynthesis and heterotrophic activity. Therefore, one question that
hydrologists and ecologists arc intercstcd in answering is how hydrological
processes transport dissolved organic carbon (DOC) to streams. An
example is the Snake River catchment, where DOC builds up in near-
353
surface soils beneath the snowpack due to microbiological activity (Boyer
et aI., 1997) and is released during snowmelt in the spring. Quantifying the
"flushing" of DOC from ncar�surface soils requires knowledge of the
extent of saturation of the soils and the rates of water movement through
them. TOPMODEL can provide a computation of these hydrological
items. For example, based on the TOPMODEL simulation shown in Figure
1 0. 1 2, the concentrations of DOC in the stream can be calculated (Figure
1 0 . 1 3). The calculated values early in the melt period do not rise to the
measured levels because the snowmelt was not calculated in TOPMODEL
correctly al early times (Figure 1 0. 1 2), but the general timing and
magnitude of calculated DOC levels are consistent with measurements.
These results suggest that such calculations could be useful to water
managers who might want to avoid taking water into the water�supply
system during times when DOC concentrations in the stream are high.
Forecasts oft1ushing of DOC would enable managers to plan ahead to
store treated water so they could avoid undue production of
trihalomethanes in the drinking water supply.
5 -_
+ + Simulated
+
�
° + 0 Observed
4-
�
-
en
E 3.5 -: +
� °
c: 3
0
-
..
0
. -
� 2.5: 0 ° + 0
c:
-
., + +
u 2 -: +
c: + +
0 0
u 1 .5 00
0 ++
6 00 qlQ
� fjl
0 1 -_ 0
C
0.5 -:
0 ,. . , • • •
. ' , " , .
>- c: '" c. > "
. ' '
::> "
'" ::> ::> w 0 w
- -
...,
::; ..., <i (J) 0 z 0
354
Figure 10.13 Measured and simulated dissolved organic carbon (DOC)
concentrations ( 1984) in the Snake River, Colorado.
Data courtesy of Ken 8encala and Diane McKnight.
• Runoff from a catchment can follow four different flow paths: direct
precipitation onto stream channels, overland flow, shallow subsurface
stormflow, and groundwater flow. {Section lOA}
355
• Shallow subsurface stormflow may occur when permeable surficial soils
become saturated. Water may then flow to the stream through these soils.
{Section 1 0.4.3 }
• The ratio of contributing area per unit contour length to the local slope of
the topography, tenned the topographic index, provides a useful measure
ofthc likelihood of saturation in a scction ofa hillslopc. {Section 1 0.S.2}
356
,
Segment 1 Segment 1
Calculate the topographic index for each segment and indicate which
segment is more likely to produce saturation-excess overland flow.
Problem 4. The soil of a given catchment has a porosity (l/J) of 0.4 and
hydraulic conductivity as described in Problem 2. The catchment has an
average value of the topographic index of 3.5. Under conditions where the
average saturation deficit of the catchment is 100 mm, calculate the
subsurface flow to the stream, qs"bsllrP�c(" If the throughfall rate at this time
is 4 x 10-3 mm S-I (14 mm hr-1), then what fraction of the catchment
would have to be saturated if saturation-excess overland flow were to be
equal 10 one-half of the subsurface flow?
Dunne, T., and L.B. Leopold. 1978. Water in environmental planlling. San
Francisco: W. H. Freeman, Chapter 9, pp. 255-278.
Wolock, D.M. 1993. Simulatillg the variable-source-area concept of
streamflolV gelleralion lVith the Watershed Model TOPMODEL. U.S.
Geological Survey Water-Resources Investigations Report 93-4124.
357
11 Water, Climate, Energy, and Food
358
the early 2 1 st century. The report notes:
Several years ago, Richard Smalley of Rice University posed the question
to a variety of audiences, "What will be the top ten problems facing
humanity over the next fifty years?" The answers consistently showed
energy, wafer, and/ood as top items (http://cnst.ricc.edu) . Without a doubt,
these resources are of critical importance and will only become more so in
the future as global popUlation increases toward 9 billion and as the effects
of anthropogenic climate change become manifest. As we seek 10 manage
these resources to meet demands, we are confronted by the realizat"ion that
the development and use of each has impacts on and is impacted by Ihe
development and use of the others (Perrone and Hornberger, 2013).
359
and then the steam is condensed by cooling. The cooling process requires
water, most often fresh water, and lots of it. According to the U.S.
Geological Survey (Barber, 2009) water withdrawals in thc United States
for thennoelectric power generation in 2005 represented 49% of total
water withdrawals and exceeded the sum ofwilhdrawals for irrigation
(31 %), public water supplies ( 1 1%), and industry (4%). Much of the water
withdrawn for cooling is returned to streams, albeit at a higher
temperature, but some of the water is consumed by evaporation. In 2008,
U.S. power plants withdrew between about 225 and 650 million cubic
meters and consumed from I I to 23 million cubic meters of water (Averyt
et aI., 2 0 11) .
Thermoelectric plants are not the only form of power generation
contributing to the withdrawal and consumption of large volumes of water.
For instance, solar power generation requires substantial amounts of water
to elean the solar panels from the dust deposited on their surfaces.
Needless to say, solar panels are often deployed in predominantly clear sky
regions with prevailing cloudless conditions and consequently arid
climatc. Thus, the supply of the water rcquired to clean solar panels and
sustain solar energy production can be problematic. Although hydropower
often is considered to be fairly non-consumptive in terms of water, an
analysis of large hydroelectric dams worldwide indicates that the blue
water consumption from them amounts to about 10% of the blue water
consumed by crops (Mckonncn and Hoekstra, 2012). Also, in some
instances the release of water from dams to run the turbines can preclude
ils diversion into irrigation canals causing a conflict between water for
crops and water for energy. Consider the case of Sri Lanka, where about
40% of the electricity is generated through hydropower. In times of plenty,
there may be enough water to supply all uses, but under drought, the
competition for water between agriculture and electricity generation can
create problems, and decisions have to be made about how to apportion the
releases. The issues involved in making decisions are complex and there is
no "right answer" (e.g., see Molle et aI., 2008).
Conversely, water distribution, treatment, and end-use require quite a
significant amount of energy. A 2005 study for California indicated that
"water-related energy use consumes 1 9 percent of the state's electricity, 30
percent ofils natural gas, and 88 billion gallons [330 billion liters] of
diesel fuel every year-and this demand is growing" (Klein, 2005). This is
a tremendous amount of energy and one conclusion of the study was that
the state could most readily meet its short-tenn (5 years) energy
conservation goals by conserving waler.
The water-energy linkage is further complicated when we consider the
360
third of Smalley's top problems. The link between water and food is
obvious. In rain-fed agriculture, crops are part of the green water portion
of thc watcr cyclc whcrcas irrigation convcrts substantial portions of blue
water into green water (see Chapter 9) . Therc also are links of food to
energy, perhaps most notably with respect to the increasing demand for
biofuels which leads to displacement of land and water resources away
from growing food. Globally withdrawal of fresh water from strcams and
aquifcrs is dominatcd by irrigated agriculture, accounting for some 70% of
the total (Rosegrant et al., 2009). Considering population growth, the
escalating demand for biofuels, and shifts toward diets with more meat
protein, which requires more water than a diet dominated by grains (see
Section 1 1.2), the demand for water for irrigation is likely to grow
globally.
361
is consumed, while the rest is eventually returned to water bodies. Because
transpiration sustains the productivity of natural and agro ecosystems, it is
often referred to as a productive consumptive usc of water, while
evaporation is defined as an unproductive fonn of water consumption.
Thus, water consumption in irrigated crops is partly contributed by uptake
and transpiration by plants (productive consumption) and partly lost in
evaporation (unproductive consumption); the relative importance of these
productive and unproductive forms of consumption determines the
efficiency of an irrigation technique.
The estimation of the blue water footprint of crops or other
commodities is based on consumption and not withdrawal rates (Table
1 1 . 1 ). The calculation of the overall water footprint of these commodities
should account for both blue and green water consumption. Such a
calculation requires an analysis of all the water consumed in the course of
the production cycle of thaI good (a life cycle analysis). The results of this
analysis are available for all the major agricultural and industrial
commodities. Some examples are summarized in Table 1 1 . 1 .
Animal products have in general a much bigger water footprint than
vegetables. The production of the same amount of food calories requires
on average 8 limes more waler in the case of meat (4 mJIIOOO kcal) than
for plant products (0.5 mJII 000 kcal; Falkenmark and Rockstrom, 2004).
Thus, the water footprint of food varies across societies (600-1800
m3/person/y), depending on the prevalent diet, particularly on the reliance
on meat. Falkenmark and Rockstr6m (2004) estimated that the water
footprint ofa balanced diet (20% of kcal intake from meat) is about 1300
m3/personly. Typical patterns of economic development exhibit an
increase in meat consumption as nations become wealthier. This shift in
diet and the escalating demographic growth are both responsible for the
ongoing increase in human demand for water resources. Thus, mankind's
appropriation of water resources poses some important ethical questions
on how far humanity should go in subtracting water (and land) from
natural ecosystems and depleting groundwater stores (i.e., aquifers), and
how water should be shared among different societies.
While climate controls the supply of freshwater resources (e.g., see Figure
1 . 1 ), population size and the type of diet determine the demand (Figure
I I . 1 ). In regions where the demand exceeds the supply, societies might
either have to sustain smallcr consumption rates or import food products
from other, more water-rich regions (Figure 1 1 . 1 ). Exporting and
362
importing food can be thought of as an indirect trade in water; this water is
referred to as virtual water (Allan, 1 998).
Virtual water trade allows nations to virtually share and exchange water
resources. Often known as the globalization of water (Hoekstra and
Chapagain, 2008) this phenomenon can be visualized by mapping the
global network of virtual waler trade (Figure 1 1 .2). In 20 I 0 the total
volume of virtual water traded in food commodities (2.8 x lOll m) y-I)
accounted for about 24% of the average global freshwater resources used
for food production ( 1 . 1 8 x 1013 mJ y- I ; Carret al., 2012). Virtual water
trade is crucial to the food security of many countries. Although the trade
in virtual water can be viewed as benevolent in that it can reduce
malnourishment in countries where local food supplies become limited by
water availability, there is some concem that in the long run, the ability of
countries to cope with drought might be compromised by an over-reliance
on virtual water (D'Odorico et aI., 2010).
363
Producing Country
Glob�1
Product
Aver�ge
Austr�lI� Br��il Chln� USA
I
Water deficit Water sumlus
r '---Import
-l
Consumption < demand Virtual Export
I
of food water trade of food
364
..
Figure 11.2 The global network ofvinual water trade for the year 2010,
based on data from FAOSTAT. Only the top links, which transfer 50% of
the total virtual water flux, are shown. Net exporters are shown in dark
blue; net importers in light blue.
Courtesy of Joel A. Carr.
• Increase the lise o/blue waterfor irrigation. Crop yields can be greatly
enhanced by switching from rain fed to irrigated agricuhure. However,
most of the suitable freshwater resources are already overexploited and
overcommitted. Many water courses are left with only minimal flows,
and some of them do not even make it to the ocean anymore. Therefore
there are only limited opportunities 10 increase water withdrawals for
irrigation.
365
• "More crop per drop." Food production can be enhanced by using
genetically modified crops with higher water use efficiency and "climate
sman" agriculture.
It is clear from Section 1 1 .2 and elsewhere in this text that humans are
directly affecting the water cycle in dramatic ways through agriculture,
energy production, groundwatcr extraction, dcforcstation (and
reforestation), and urbanization, to mention just a few. Humans are also
affecting the water cycle more indirectly by increasing atmospheric levels
of carbon dioxide (C02), which alters air temperature and transpiration. It
has been suggested that these alterations may cause an acceleration of the
water cycle (see Box 1 1 . 1 ) and global redistribution of water, leading to
more extreme weather events that may increase the number of floods in
some regions while making other regions are more drought-prone. Because
the water cycle is tightly linked to biogeochemical cycles (e.g., carbon and
nitrogen) and vegetation distributions, changes in the water cycle due to
climate change are likely to have far-reaching effects on ecosystems and
landscapes.
366
surface air temperature of 1.8°C to 3.4°C by 2100 (Meehl et aI., 2007).
These increases will not be uniformly distributed across the globe, with
somc areas, such as ponions ofthc southwest U.S., cxpericncing a largc
increase in the number of days with a heat index > 100°F (37.8°C) by the
end of the century (NRC, 20 12b).
Changes in air temperature will affect precipitation, evaporation and
transpiration. Higher temperatures will increase evaporation rates of
surface water (from oceans, lakes, or soil). In addition, wanner air can
hold more watcr (sec Figure 2 . 1 ). An increase in cvaporation will decrease
soil moisture, which could reduce transpiration, but changes in air
temperature and vapor pressure will also affect transpiration rates. Because
some of these changes can be offsetting, sorting out how precipitation and
evapotranspiration respond to increascs in air temperature is challenging.
Hydrologists and other environmental scientists are combining theory,
models and measurements (in the lab and field) to make the best forecasts
of future change in precipitation, evapotranspiration and runoff given our
current understanding (see Box 1 1 .1 ) .
p = el = K'e...A I - ele"",),
367
temperature is shown in Figure 2 . 1 : because e.al is an increasing
function of T, climate warming is expected to increase global
precipitation, an effect known as the intensification of the global
water cycle. The dependency between eSIJI and T suggests that global
precipitation should increase by 6.8% per each degree Celsius of
global wamling. Climate change models and observations, however,
indicate that the increase in precipitation associated with IOC of
wanning is more in the 2-3% range, presumably owing to other
factors affecting the global conductance parameter K' (e.g., Katul et
aI., 2012). This analysis is based on global mean values of
temperature and precipitation. Climate change research has shown
that changes in both temperature and precipitation are not expected
to be unifonn around the globe (see Figure 1 1 .3).
-,
Figure 11.3 Annual mean change in precipitation (in inches of water per
year) for 2081 -2100 relative to 1950-2000 calculated bascd on the
Geophysical Fluid Dynamics Laboratory (GFDL) CM2.1 model in which
carbon dioxide was increased from 370 to 7 1 7 ppm. Positive values
correspond to wetter conditions.
Redrawn from an image developed by NOAA GFDL (www.gfdl.noaa.gov).
Climate models have been used to predict the change in amounts and
global distribution of precipitation as the air warms. Figure 1 1 .3 shows the
distribution of predicted changes in annual mean precipitation in 208 1-
2 1 00 relative 10 1950-2000 calculated based on the Geophysical Fluid
Dynamics Laboratory (GFDL) CM2.1 model in which carbon dioxide was
increased from 370 to 7 1 7 ppm. Whereas high latitudes are likcly to
368
experience more precipitation (as rain or snow), low latitudes, except in
the immediate vicinity of the equator, tend to become drier.
Not surprisingly, these patterns are generally mirrored in predicted
changes in runoff in thc 2 1 st century (Figure 1 1 .4; Milly et aI., 2005).
Incrcases in evaporation caused by increasing temperatures coupled with
decreases in precipitation lead to decreases in runoff in many parts of the
world. In other parts, increases in precipitation are leading to increases in
runolT. For example, reconstruction of monthly discharge from the world's
largest rivers indicates that global continental runolT increased during the
last century (Labat et ai., 2004), most likely due to increases in
precipitation in the watersheds feeding these rivers. Regardless of whether
precipitation is increasing or decreasing, observations suggest an increase
in the intensity of the largest rainfall events.
369
40
30
20
10
2
-2
As the Danish physicist Niels Bohr once said, "Prediction is very difficult,
especially about the future." Nevertheless, it is clear that we face serious
issues with regard to what are called social-ecological systems in the
coming decades, and that many of these issues will be framed by the topics
above--ensliring water and food security in the face of a changing climate.
Whether predictions (or forecasts or sccnarios) about thc futurc arc borne
out will depend on actions taken by individuals and collectives of
individuals, including nation states. We believe that students of
environmental science should have a broad understanding of the
fundamentals of physical hydrology so they can help shape the future in
ways that arc infonllCd by scicncc. Thc NRC (21 02a) rcport notcd at thc
beginning of this chapter concludes that: "Compelling challenges and
opportunities lie ahead in understanding, quantifying, and predicting water
cycle dynamics, the interaction of water and life, and how to build a path
to the sustained provision of clean watcr for pcople and ecosystems." The
hydrological underpinnings for mccting the ehallcnges and taking
advantage of the opportunities fonn the substance of this book.
1 1 .5 Suggested Readings
370
National Research Council (NRC). 2012a. Challenges and opportunities in
the hydrologic sciences. Washington, DC: The National Academies
Press.
National Research Council (NRC). 2012b. Climate change: Evidence.
impacts. alld choices. Washington, DC: The National Academies Press.
371
APPENDIX I
Units, Dimensions, and Conversions
Hydrological Quantities
Unit Conversions
Hydrological Quantities
372
total amount (volume) of water delivered to the catchment during a single
cvent:
373
Quantity Dimension Unit 51 Symbol Formula
Sase units
Oerlved units
Area l
(l ] square meter m'
374
Prefix 51 symbol Multiplication factor
Needless to say, a wrong answer will result ifthe units on one side of an
equation are different from those on the other side.
The most common system of units employed today is the SI (System
illternational d'Unites; Table A 1 . 1 ). Other widely used systems of units
include the Ellglish !.ystem (foot, pound) and the cgs !.ystem (centimeters,
grams, seconds). The SI system includes a sequence of standard prefixes to
indicate magnitude (Table AI .2). For example, a kilogram is equal to 1 000
grams, and a nanogram is equal to 10-9 grams.
375
x 105 m2 = 2684.5 mJ. This answer is correct mathematically, but it is
expressed with greater relative precision than is justified by the measured
values. There arc too many significant figures in the answer. This requires
some explanation. First, the measured quantities have a certain absolute
precisioll. For the measured precipitation depth, the absolute precision is I
mm. It is likely that the rain gage could only measure precipitation depth
to the nearest 111111. The absolute precision for the area (206,500 m2) is 100
m2 . This is apparent because we wrote the value initially using scientific
notation. For a number written as 206,500, it is not clear whether the
absolute precision is 100 m, 1 0 m, or I m. The relative precision is best
thought of in terms of sigllificalllfigures. In a given quantity, a significant
figure is any given digit, except for zeros to the left of the first nonzero
digit, which serves only to fix the position of the decimal point. Some
examples are given in Table A I .3 .
30 l' 3 X l01
.01 1 l x lO-.1
0.01 1 1 x lO-.1
0.010 2 1.0 X 10 -l
376
area (2.065 x 105 ml) has four. Therefore, our answer should have only
two significant figures, 2.7 x 103 m3.
When adding or subtracting numbers, the rule is that the answer should
be expressed using the same number of significant figures as the number
with the/ewest decimal places. Finally, numbers should not be rounded to
the appropriate precision in a calculation with several steps unti! the vely
end.
Unit Conversions
Note that the answer is expressed in scientific notation, using the same
number of significant figures (2) as the least relatively precise number,
which in this case is the hydraulic conductivity value. [t should also be
obvious that the multiplication has resulted in the original units (ft, day)
being " canceled."
Tables A 1.4 through A 1. 8 provide equivalent values for many
quantities used frequently in hydrology. To use the tables, look down the
left column to find the unit you are interested in using to express a
quantity. In the example above, we wanted to convert feet to meters.
Looking down the first column of Table A 1 .4 you will find a row
beginning with "meter." Scanning across this row, you will find that 1
meter is equivalent 10 3.281 feet, which is the value used to constmct the
ratio in the example above.
377
'"1.'''''' '
"..; , �..
, , ',0]' 11' •.DO],., 0.00'000 '.D�,a-< D.Io>,. . ,a-<
- 1.'09,000 �� ". ,- ,- ,
aln Tables A 1.4 through A I ,8, values are shown to four significant figures, and the
S I expression, in base or derived units, is in italics.
- - - - - �,- .- -
"0' , 0._' "'.>x ,a-< ",.. "a-< ..." " ,... "', ' x ,a"" '''1 ' , ,a""
•• -, - � .- ,-' ,� -� --
KO' , D.O" '" .-� '1I.1xllt'" 2l.... .a-< 10, ,,, " "" 11.H. W""
aa.·_ 1'-1h.0' " llh ,," l.IS h .(1' <."h.D' li" "" ,
- .......
� , ••m O.Q(l<'" 0.00>>1' " .C9 x ,a-<
--- 15,1500 ,� - �, ,
378
1....-
379
APPENDIX 2
Properties of Water
General Properties
Water Density and Viscosity
General Properties
380
float on top of cooler water at temperatures above 3.98°C, while below
3.98°C, cooler water will float on top of wanner water. As a result, ice will
lend 10 form and remain at the surface of a water body. When it melts, the
waler becomes denser again and sinks below the surface. This process
leads to spring mixing of surface and deeper water in many mid- to high
latitude lakes.
The viscosity of water (p, often referred to as the dynamic viscosity)
also varies with temperature (Figure A2. 1 ; Table A2. 1). In fact, the
variation in water viscosity is much greater than thai of density over the
temperature range most commonly encountered in hydrology. For
example, from 1 0 to 30°C, waler density decreases from 999.73 kg m-3 to
997.07 kg m-3, a decrease of 0.27%. Over the same temperature range, the
viscosity of water decreases from 1 . 307 x 10-3 Pa . s (at 10°C) to 0.7975 x
1 0-3 Pa ' s (at 30°C), a decrease of39%. For this reason, hydrologists are
often coneemed with the temperature effects on water viscosity when
making their calculations. [n the Sl system (see Appendix I ), the units of
�__�""':::::::: ==��=:�
:
viscosity are Pa ' s.
1000 ____
_ - 1 .9 x l 0-a
3
- 1 .8xl0-
\ 1.7 xl0-3
999
3
- 1 .6xl0-
-
7
t!'
-
. 1 . 5 x l 0-
3
..
E 998
'" - 1 . 4 x l 0-
3
e:.
""
�
·
- 3
,..
1 .3x l 0-
- 1 .2xl0- 8
- '"
.- 3
'"
c
997
:£
_
.. 1 . 1 x l 0-3
C
_
- 1 .Ox l 0-3
996 : 3
O.9xl0-
3
- O.8x l 0-
995 O.7xl0-3
o 5 10 15 20 25 30
Temperature (oC)
Figure A2.1 The density (P) and viscosity (P) of water as a function of
temperature, from 0 to 30°C.
Table A2.1 The density (p) and viscosity (�l) ofwllter as a function of
temperature, from 0 10 50°C
381
Temperature (-e) Density (kg m-J) Viscosity (Pa·s)
3.98 1000
25 997.07 8.904 x 1�
382
APPENDIX 3
Basic Statistics in Hydrology
383
1 979 995
1 980 780
1981 1004
1982 930
1983 1 192
1 984 950
• •
In other words, the cdr for a chosen value, a (1 100 mm, in the example
above), is the probability thai the outcome ofa random process, x
(precipitation next year), will be less thall or eqllal lo the chosen value.
Notice that the limits of integration are negative infinity and Ihe value a;
the integration is adding up all of the probabilities associated with all of
the values less than and including a. Conversely, the compleme1/f(fl)'
Cl/IIII/Ialille distriblltiolljilllclioll, often written as G(a), is defined as:
384
Equation A3.5 indicates that the total probability is equal to I (annual
precipitation must have sOllie value).
The distribution of a sample of values of a random variable can exhibit
a varicty of forms and can often be describcd by relatively simple
probability density functions. Examples include the normal distribution,
the log-nomlal distribution, and the exponential distribution.
The normal distribution can be described with two parameters: the
mean (pJ and the standard deviation (oJ. Given a sample of data (e.g.,
several decades of annual precipitation), the mean and standard deviation
are approximately x (for /Ix) and Sx (for ox):
LX,
"
,
x=
,
(A3.6)
.
"
L(x, -",)'
"
, ,
" (A3.7J
.
/1-1
. •
.. . =
(A3.')
385
is I (Figure A3 .2). Values of the cumulative distribution funclion (Table
A3.2) or complementary cumulative distribution function of the standard
normal distribution can be used in calculations. Thc z-value represeOls the
nonnalized outcome for which the probability is desired ( 1 100 mm, in the
earlier example). This value is calculated according to:
(AJ.IO)
a-x
For the example data in Table A3 . 1 , with a = I 100 mm, x = 979 mm, s" =
I 0 I mm, the calculated z-value using equation (A3.1 0) is 1 .20. (Note that
the z-value is dimensionless.) The edf for this z-value (Table A3.2) is
0.8849 (i.e., F(z) = 0.8849), which indicates that the probability that
precipitation in any year will be less than or equal to 1 1 00 mm is about 0.9
(Figure A3.3).
386
(a) (b)
"N'
t.;!.
, F(x)=F(z)
�
�
"
--< --<
�
" �
"
Jlx o -1 0 1
0.4
0.35
0.3
0.25
N-
o:::; 0.2
0.15
0.1
0.05
-4 -1 1 4
387
For l. = 1 .26
F(z)
-4 -3 -2 -1 o 1 2 3 4
z
Figure A3.3 The standard normal distribution showing the portions of the
distribution given by the cumulative distribution function, F(z), and
complementary cumulative probability, G(z).
T
1 1
= =10
",,,,,,, excecdance probability 0.1 yr�l
_
yr. (A3.1 1 )
388
might be used to predict possible behavior. Some of the concepts
introduced, such as probability, probability distribution, exceedance
probability, and return period, arc used throughout the book.
389
• 0.00 0.01 0.02 0.0] 0.04 0.05 0.06 0.07 0.08 •."
,., ' ''''' 0.5040 0.5080 a.SIlO 0.51(;0 0.5199 0.5239 O.S179 0.5319 0.5359
,.. 0.5398 0.5438 0.5478 0.5517 0.5557 0.5596 0.5636 0.5675 0.5714 0.5753
,.. 0.5793 0.5832 0.5871 0.5910 0.5948 0.59117 0.6026 0.6064 0.610] 0.6141
,., 0.6179 0.6217 0.6255 0.6291 0.6331 0.6]68 0.&406 0.6443 0.6480 0.6517
, .• 0.6554 Q,WU 0.6628 0.6664 0.67(1) 0.6736 0.6772 0.6808 0.6844 0.6879
OS 0.6915 0.6950 0.6985 0.7019 0.7054 0.7088 0.7123 0.7157 0.7190 0.1224
,.• 0.7257 0.7291 0.7324 0.7357 0.7389 0.7422 0.7454 0.7486 0.7517 0.7549
,., 0.7580 0.76U 0.7642 0.7673 0.7704 0.7734 0.7764 0.7794 0.7823 0.7852
,.• 0.7881 0.7910 0.7939 0.7967 0.7995 0.8023 0.805\ 0.8078 0.8106 0.8133
,.• 0.8159 0.8186 0.8212 0.8238 0.8264 0.8289 0.83\5 0,8340 0.8365 0.8389
.., 0.8.413 0.8438 0.8461 0.8485 0.8508 0.8531 0.8554 0.8577 0.8599 0.8621
U 0.8641 0.8665 0.8686 0.8708 0.8729 0.8749 0.8770 0.8790 0.8810 0.8830
.., 0.8849 0.8869 0.8888 0.8907 0.8925 0.8!144 0.8962 0.8980 0.8997 0.9015
U 0.9032 0.9049 ' '''' 0.9082 0.9099 0.9115 0.9131 0.9147 0.9162 0.9177
L. 0.9192 0.9207 0.9222 0.9236 0.9251 0.9265 0.9279 0.9292 0.9306 0.9319
L' 0.93l2 0.9345 0.9357 0.9370 0.9382 0.9394 0.9406 0.9418 0.9429 0.9441
••• 0.9452 0.!I463 0.9474 0.9484 0.9495 0.9505 0.9515 0.9525 0.9535 0.9545
U 0.9554 0.9*, 0.9573 0.9582 0.9591 0.9599 0.9608 0.9616 0.9625 0.9633
L' 0.9641 0.9649 0.9656 0.9664 0.9671 0.9678 0.9686 0.9693 0.9699 0.9706
L' 0.9713 0.9719 0.9726 0.9732 0.9738 0.9744 0.97SO 0.9756 0.9761 0.9767
,., 0.9772 0.9778 0.9783 0.9788 0.9793 0.9798 0.9803 0.9808 0.9812 0.9817
U 0.9821 0.9826 0.9830 0.9834 0.9838 0.9842 0.9846 0.9850 0.9854 0.9857
>.> 0.9861 0.98&4 0.98&8 0.9871 0.9875 0.9878 0.9881 0.9884 0.9a87 0.9890
U 0.9893 0.9896 0.9898 0.9901 ,."" 0.9906 •."'" 0!�911 0.99H 0.9916
, .• 0.9918 0.9920 0.9922 0.9925 0.9921 0.9929 0.9911 0.9932 0.9934 0.9936
,.. 0.9938 0.9940 0.9941 0.9943 0.9945 0.9946 0.9948 0.9949 0.9951 0.9952
U 0.9953 0.9955 0.9956 0.9957 0.9959 0.99&0 0.9961 0.9962 0.9963 0.9964
,., 0.99&5 0.9966 0.9967 0.99&8 0.9969 0.9970 0.9971 0.9972 0.9973 0.9974
U 0.9974 0.9975 0.9976 0.9977 0.9977 0.9978 0.9979 0.9979 0.9980 0.9981
U 0.9981 0.9982 0.'�982 0.9983 0.9984 0.9984 0.9985 0.9985 0.9986 0.9986
Note: To use the table, scan the left column to locate the z-value to the first
decimal place, and then scan across the row to find F(z) for the second decimal
place. Note that for z < 0, use F(z) I F(IzI).= -
390
Answers to Example Problems
Chapter 1
Chapter 2
Chapter 3
Chapter 4
Chapter 5
Chapter 6
Chapter 7
Chapter 8
Chapter 9
Chapter 10
Chapter I
From Table AlA, 1.0 inch is equivalent to 25.4 mm. The conversion
is as follows:
. _).
,- 4 mm
Death Valley: 1 .6 m x = 4 l mm
1 in
25.4 mm
nun
.
= 1 . 1 7 x l O,
1
Mt. Waialeale: 460 m x
in
391
equivalent flow (in m3 S- I) of 18.2 ft3 S-l ? (You might want to review
Appendix I on units, dimensions, and conversions.)
From Table A 1 .7, 1.0 rtl S-I is equivalent to 0.02832 m3 S-I. The
conversion is as follow:
0.02832 m) S-I
? ft3 s-I x
1 8._ J = 5 . 1 5 x 10-1 m3 s- 1
I ft S-I
106 m1
Volume = Depth x A rea = 1.0 III X \03 km1 x = 109 m).
I km1
B. In gallons?
264.2 gal
Volume = I 09 m) x = 3 x 10" gal.
I mJ
Problem 1.4. The polar ice caps (area = 1.6 x 107 km2) are estimated to
contain a total equivalent volume of2.4 x 107 km) of liquid water. The
average annual precipitation over the ice caps is estimated to be 5 inches
per year. Estimate the residence time of water in the polar ice caps,
assuming their volume remains constant in time.
39.37 in I kml
V 2 4 x l 016 m3
Residence time = Tr = - = . 3 = 1 .O x I 04 yrs.
J 2.032 x 1 0 12 In yr- I
392
=
Problem 1.5. In an average year, a small (area 3.0 km2) agricultural
catchment receives 950 mm of precipitation. The catchment is drained by a
stream, and a continuous record of stream discharge is available. The total
amount of surface-water runoff for the year, detennined from the stream
discharge record, is 1 . 1 x 106 mJ.
A. What is the volume of water (in mJ) evapotranspired for the year
(assume no change in water stored in the catchment)?
r, = 1 . 1 x I 06 mJ•
p = 950 mlll X
1m
x 3 .0 km2 x
1 06 m2 ) = 2.85 x I 06 m3.
1 000 mm I km2
B. What is the depth of water (in mm) evapotranspired for the year (again,
assuming no change in water stored in the catchment)?
�= .,.
.:.
1 -:.:x I ::.
0..
1 .:..:. ' m
:..
.:.: ' --:-
=- � = 1 . 1 x 106 m 3 . .
= 0 39
P
_ __
____
1 m 106 m2 2.85x 1 06 m3
950 rmn x x3.0 km2 x
I km-
7
1 000 mm
Chapter 2
Problem 2.1. Two tipping bucket rain gages arc used to collect the
following rainfall data:
393
Cumulative precipitation (mm) Cumulative precipitation (mm)
Station III Station 112
Time
6:00 p.m. 19 17
8:00 p.m. 19 17
10:00 p.m. 19 17
12:00 midnight 19 17
2:00 a.m. 19 17
4:00a.m. 19 17
A. Calculate the mean daily rainfall intensity for each station (mm hr-I).
B. Calculate the maximum 2-hour rainfall intensity for each station (mm
hr-1 ),
By inspection, tbe greatest 2-hr rainfall occurred between 10:00 a.m.
and noon:
C. Calculate the maximum 6-hour rainfall intensity for each station (mm
h,.-I).
The greatest 6-hr rainfall occurred between 8:00 a.m. and 2:00 p.m.:
D. Using the arithmetic average method and knowing that the drainage
basin area is 176 mi2, calculate the total volume of rainfall (m3)
394
delivered to the basin during the event.
p = ( l 9 + 17)12.0 = 1 8 mill.
, (
Cylindervolume=areaxdepth= I. 1 4 m- x IO.O rnx
.
.
2.54 em
I In
x
1 m
100 em
) = 0.290 m,
B . Initially, the pan contains 10.0 U.S. gallons of water. Calculate the
depth of water in the pan (mm).
0.003785 on'
J
VoI lime = 1 0 .0 gaI x = 3.79x 1 0-2 m .
I gal
395
D. After 24 hours in an open field (no precipitation), the pan is checked
and the volume of water left in the pan is detennined to be 9.25 gallons.
Calculate the average evaporation rate (mm hr-1 ) from the pan.
el=O.lOmm hr-1
E. The pan is emptied and refilled with 10.0 gallons of water and left in an
open field for another 24 hours. During this period, rain fell for a 3-hour
period at a constant intensity of 2.5 mm hr- 1 ; after 24 hours, the volume
of water in the pan was 1 1 .50 gallons. Calculate the average
evaporation rate (mm hr-1) from the pan during this period.
et = p - (dV/dt)/(pan area).
The total precipitation for the 24-hour period is (3 hr)(2.5 rum hr-I)
= 7.5 rum. Over the 24-hour period, this is equal to an average
precipitation rate of:
dl 24 hr I gal
396
empty as a result of evaporation.
dl = --;-:
.,-'-
----:
dV
(el X pan area )
0.003785 1n3
dV = - I I .SO gal x = -4 .3Sx 10-2 In ) .
I gal
et P.. �,.
E
I ,
=
PM)'"
el or E, =
I hr
ct = 1 .04x I O""" m hr-1 x = 2.89 x I O-s ml s-l.
3600 s
E, = (2.89x 10-' m ,-')(1 000 kg m-')(2.45x 1 0' J kg- ' )= 70.8 W m-'.
Problem 2.4. A small (area = 300 ha) catchment in Iowa absorbs a mean
RIO = 330 W m-2 during the month of June. In this problem, you will apply
the energy balance approach to estimate evapotranspiration from the
catchment during the month of June.
A. Write a complete energy balance equation (i.e., including all terms) for
the catchment for the month of June.
dQ
= = R. - G - H - E,.
dl
B. Neglecting conduction to thc ground (G) and thc changc in energy
stored (dQ/dl = 0), simplify your energy balance equation for the
catchment so that it can be solved for the latent heat flux, E/. Also,
replace the term H (the sensible heat flux) with B x E/ (where B is the
397
Bowen ratio).
R,, - H - E, =O.
E -
R"
,
_
B+ I
C. Using a mean Bowen ratio of 0.20 for the catchment, calculate the mean
daily nux of latent heat to the atmosphere (W m-2) and the mean
evapotranspiration rate (mm day-I) from the catchment. Use a water
density, Pw = 1000.0 kg m-3.
1 .20
EI ' 75 W m -,
-
el :::; = 1 . 1 2 x lO-7 m s-l•
P.)" ( \ 000 kg m lX2.4Sx 106 J kg I )
1 000 mm 86400 s
e/ ::: 1 . 1 2 x 1 0-7 m S-I x x = 9.68 mill day- I =9.7 mill day-I .
1 m I day
Chapter 3
398
B. Would the pressure change significantly if the water temperature was
22°C instead?
d = 30.6m.
D. What depth (m) of mercury, with a unit wcight of 133 kN m-J, would
be required to produce a pressure ofJOO kPa?
Problem 3.2. A plate is pulled over a horizontal layer of water that is 10.0
mm deep (Figure 3. 1 ). The temperature of the water is 20°C. If the plate
exerts a shcar stress of 0.0 I N m-2 on the upper surface of the water, what
is thc spccd (m S- I) ofthc platc?
From Equation 3.1 lIplale = (FIA)(d4/). FIA is the shear stress, equal
,
in this case to 0.01 N m-2 or 0.01 kg m-1 s-2. The viscosity at 20°C is
1.00 x 1 0-3 Pa . s (Table 3.1) or 1.00 x 10-3 kg m-I S-I . Therefore,
399
is again laminar up to RI'/ille =
2000. [A parameter called the hydraulic
radius (Chapler 4.5) can be used 10 find values of R corresponding 10 Ihe
laminar -Iurbulent transition for different flow cross-sections. This is
explored further in an example problem in Chapter 4.]
Rearranging Rl'lal pllplate dip to solve for IIplate gives lip/lite Rl'lal"
= =
"
dT aT aT
- -:;- + II- (0.2°C hr- I ) + (I OOO In hr- 1 XI x 1 0....oC m- I ) O.3°C hr- I.
df 01 a.\"
= = =
Problem 3.5. A tank like the one pictured in Figure 3.6 is filled to a
constant level ofO.70m. The center of the outflow opening near the bottom
is 0.1 Om above the bottom of the tank. What is the velocity (m S-I) of flow
exiting from the outflow opening?
d=
The depth between the water surface in the tank and the outflow
opening is: =
Zl - Z2 0.70 m - 0.10 m 0.60 m. From Equation =
400
This is the velocity of flow exiting the tank.
- Ul-Vj •
2g
pg 2g 2g
"'
_
2 (p,-p,) 2(7.5x 10 ' I'a)
U- -
1 15p 1.5x 104 kg m-J
20-mm diameter hose. The viscosity of the water is 1.0 x 10- 3 Pa . s, and
the density of the water is 1000.0 kg m -3•
401
Because R > 4000, the flow is turbulent.
B. What is the friction factor and the head loss per unit length for this flow
(both are dimensionless)?
C. What is the change in pressure (Pa) over a 10-m length of the hose?
(1 .0111) 1
u=_ p 02 = -(-2 0x I Ol Pa m -1)
d = O.063 m s- l .
dx3211 32(LOxIO'Pa·,)
Chapter 4
402
bottom) with no head loss or change in the width of the channel.
A. Calculate the specific discharge (m2 S-I) and specific energy (m) at the
upstream station.
1
- 0 .)-00m' S-1 .
Q 1.50m} S
q.., - -
_ _ - _
w 3.00m
B. Calculate the specific discharge (m! S-I) and specific energy (m) at the
downstream station.
£2 = EI + 0.100 m = 1.61 m.
0.333 m S-l
r U 0.086.
hh J(9.81m s-2X1.50 m)
= = =
403
(0.5 m' s·' ) '
- = 1.606 m -
" ll) 2(9.81 m s·' X2.0 111)'
= 1 . 6028 m.
Problem 4.2. A discharge of2.0 m3 S-l is carried in a canal with the cross
section shown in Figure 4. 1 5. The canal is 1400 m long and drops 0.50 m
in elevation over that distance. Manning's II for the channel is estimated to
be 0.020. What is the value of IV (m) for this canal?
k
Q=llwh= - RJPSV211'h= - RNJS1l2�
k \\,2
fI II 2.5
where the hydraulic radius is given by:
Therefore,
213
1 ml13s-1
'"
m' -
'"
-
0.5 m ",'
,
- ,
s 0.020 4.5 1400 t11 2 .5
404
w
1V/2.S
.
Figure 4.15 Canal cross sectIOn for Problem 4.2.
A ,,(012)' D
R II
�
P ,,0 4
� �
R in terms of R".
405
C. Set the expression found in 3 8 equal to 2000, the critical Reynolds
number for pipes. Rearrange this to find the equivalent critical value for
RRH = pURfi,l.
JI JI
- -
4
•
If RII;::;: II, then the critical Reynolds number for the channel flow
will be 500 as in 4.3C.
Problem 4.4. Use Equation 4.32 to estimate the depth and mean velocity
of a flow in a channel with a slope S = 0.003, width IV = 1 5 m, discharge Q
1
= 1.0 m3 s - and Manning's /I = 0.075. Does the assumption that RH;::; II
Chapter 5
Problem 5.1. Irthe flood used in the reservoir example delivered the same
volume of water in a shorter amount of time (shorter duration with higher
peak dischargc), as given by the inflow hydrograph in the table below,
how would the outflow hydrograph change? Complete the table below,
assuming the initial conditions and other reservoir parameters remain the
same.
406
Time t. 2Vp 2Vp:1
I. I.+I.�,
(days)
Step n -0. +° + 0••, t••,
'" At .1
A B C 0 ,
2 0 25
. 4.0 22.0 346.0 368.0 5.31 0.50
6 1 25
. 3D 4.5 359.3 363.8 4.30 1.50
The completed table for this larger inflow shows that the peak
outflow is a smaller fraction of the peak inflow (O!'cak/l!'cak = 0.56)
compared to the example in Table 5.3 and Figure 5. 1 0 (O!'eak/1peak =
0.63).
Problem 5.2. The Muskingum routing coefficients for a stream reach are
detem1ined to be: Co = 0.26, C1 = 0.55, C2 = 0.19. For the inflow
hydrograph given in the table below, complete the calculation of the
predicted outflow hydrograph.
0000 10 10
0600 50 20
1200 130 6S
2400 70 100
A. Fit a line through the data and determine the return period of an 8000
m 3 S-1 flood.
407
period T
....'urn
= 110.20 = 5 yr.
408
o
0.05
� 0.10
.g 0.20
.g 0.30
0.40
�
c.
..
u
c:
0.70
..
"C
a: 0.80
l:l
w
o
o
1
'
Discharge, Q (m s-')
0.01
0.02
0.05
0.10
,.,
!::
.
-
.
.c
-
..
.g 0.30
0.40
�
c.
..
u
c:
..
�
�
"
w 0.90
0.95
0.99
10' lcr 103 104 10'
Discharge, Q (ml 5-1)
409
Figure PS.I Probability plot for Eel River, California, peak annual
discharge data, 1950--1 989 (a), and 1917-1 996 (b).
Chapter 6
Problem 6.1. You arc charged with designing a very simple filtration
system for a community water supply, using cylindrical sand columns (K =
5.0m day- I). The filter needs to be 3.0m long to adequately trap
particulates in the water, and since the system will be driven by gravity,
the pressure heads at the top and bottom of the (vertically oriented) filter
wi11 be zero.
A. What diameter filter is required to treat 4.0 x 103 gallons of water per
day? Is this value feasible (anything larger than about I m is not
feasible)?
Because the pressure heads at the top and bottom of the filter are
zero, the hydraulic gradient must be equal to I. Therefore:
I rn3
Q=4.0x I03gai day-Ix =15.1 mJday-'.
264.2 gal
,
dh
=-KA-=
o
Q 5.0mday-lxJr -
dz 2
,
or
410
112
4 15.1 mJday-l
D= 2 .0 m.
!C 5.0
=
m day-l
B. Consider each of the alternatives and how you might modify your
design:
A. Which way is water flowing in the column? [s water flowing from high
to low hydraulic head? From high to low pressure?
41 1
q = Q/A = (500.0 mm3 min-I)/[lI'(25.0 mm)2] = 0.25 mm min-I
K=2.50mm min-I.
k= KJ!.... =4.2 x 10-' rn 5-1 x [1.0 x 10-3/(1000.0 kg rn-J x 9.81 m 5-2)] =4.3 x 10-12 m l.
pg
A. Calculate the elevation and hydraulic heads (rclativc to mean sea level),
pressure head, and fluid pressure in the two piezometers, and fill in the
table below.
Piezometer n Piezometer #2
412
Is the flow of water upward or downward?
Problem 6.4. Consider the now net for a drainage problem shown in
Figure 6. 1 5. Drains such as pipes and culverts placed in a wet field may be
uscd to removc groundwater by ercating a "sink" or area of low hydraulic
head. In the figure, a cross section through such a field is shown. The
hydraulic conductivity, K, of the surficial material is 1 .0 x 10- 5 m S-I. The
thick black lines represent impenneable boundaries; a constant head is
assigned to the top and lower left side. The cross seclion is 20 m long by
1 0 m deep. The gray lines are equipotentials and the blue lines are
streamlines.
. . ..
. .
. . .
. .
. . - . .. . . - ...
.
..
. .
.. . : . . . .
. '. . ' . .
. . .
'. '
.
.
,
. . '
' . . .
·
. . .
' . . .
.
.. ' ..
. . .- .
'
: ,
:.
.
...
. . ' . . '
' '
,
. '. ' "
.. . . . .
.
.
. ' ' .
,
. . '. ' " ' ' . '. .
. . . ..
·
.
'
,". " ' '
' ' .
.'
. : .
. . : . . . :
' . . . .. . . . '
,
. . . . . .
. . . " . _ . .
,, . . . . . , . .. . . . .. . . . .
.
.
: -
' ' "
"
. . "
' .
. .
. .. . . : ... . .
' . .
.
'. ' .
, . .
. . " ' .' . .'
; -. :-:- :,-, : : :
: . . . :
.
. ' . . . ' '
.
. .
. .
-- . . . - . .
'
. " ' .
.. : . - .. .:.
. . " : , ,
.. . . .. . : ..
' . ' "
. .
. ..
-
.
:. "
: - . :
:
,. . : .. . .. ... : .: : .. . :
. . . . -.
. . ' ' '.
.
. ..
. . ' . ' .
. . . . .
.. .
. . . .
. . . . . . .
. .. . ... .. . ... .
. . . .
.
.. . .
.
·
. . . .
.",
. .
•
. ' .
. . .
.
'
'. . .
B. Calculate the discharge through each streamtube, and the total discharge
or ratc at which thc field is draincd (m 3 day-I pcr m width of material).
413
Qs = Kb dh = 10- 5 m S-I x I m x 2 m = 2 x 10- 5 m 3 S-I x 86400 S day-I
= 1.7 m3 day-I.
Chapter 7
Problem 7.1. Oetennine the natural basin yield (m3 yr-I per meter basin
width) for the following cases.
C. For the basin in Figure 7.5a with L = 200 m and K= 100 m yel. You
might want to begin by calculating the discharge through each local
flow system (dashed box).
Problem 7.2. Answer the following questions for the basin in Figure 7.25.
414
e
-
x (m)
B. What fraction of the total basin yield passes through the lower unit?
Problem 7.3. Consider the flow net shown in Figure 7.26. The sides of the
region are groundwater divides, the top boundary is the water table, and
the bottom is an impermeable boundary. Streamlines are blue and
equipotentials are gray.
415
D. Indicate at least one area within the flow net where flow is relatively
fast, and one area where flow is relatively slow.
The open portions of the wells intersect the 12.5 m (well A) and 30
m (\\'ell B) equipotentials.
"
30
E
-
N ".
20
A
'0
s
0 +- ��
'� 5 �
__ .5 2��
71 �� o �
n . 5�25"-
� 27��
5. ��
� ____ *32�
.5"--.
____,;35"--,__ ___
o 10 20 30 40 50 60 70 80 90 100
x (m)
Figure PS.2 Answers to Problem 7.3, parts A through D.
A. What is the change in water-table level during this time period (m)?
Does this indicate an increase or a decrease in water stored within the
aquifer?
B. If the specific yield of the aquifer is 0.25, and the aquifer has an area of
600 km2, what is the change in water stored over the same time interval
(m))?
416
Again, assuming that the change observed in the one well is
representative of the aquifer as a whole, the change in water stored
is found by multiplying the specific yield by the area and the
change in water level:
Problem 7.5 Determine the recession constant, c (day- I), for the two
groundwater "reservoirs" contributing to Pescadero Creek during the
spring and summer of 20 I 0 (Figure 7.23).
c = "I"
:.",
Q Til Q",.
- -'.",
.! o_
I
417
1.0
0.8
0.6
�
-
in
0.42.
E 0.4
0.14- 1-
- _ I
0.115
- L _ I-
0.1
May Jun Jul Aug Sep
Month in year 2010
Figure PS.3 Groundwater recession analysis for Pescadero Creek,
California, showing points used in analysis of recession constants
(Problem 7.5).
Chapter 8
Problem 8.1. Tensiometers are installed at OAm and O.Sm above the water
table in a uniform sandy soil with the moisture characteristic and hydraulic
conductivity curves given in Figures 8.4 and 8.5, respectively. One set of
tensiometer readings indicates that the capillary-pressure head al lhc !irs!
of these tcnsiometers is 0 45 In and at the second is -0.6 m.
-
.
418
A. What is the direction of water movement between the two
tensiometers?
419
30
'"
-
-25
'"
- "
r::
"
-
r:: 20
0
"
"
�
:J 15
-
...
0
-
E
." 10
-
-
"
E
:J 5
0
-
>
o
102 10 ' -1 10 1 10 2 104 10 ....
Capillary-pressure head, 'V (m)
_ _ - - _ - _ -
420
20 20 ,------r--,
0-1------
� 20 � 10 0 o 1 2 3 4 5 6
Capillary-pressure Volumetric moisture
head, III (m) content, e (%)
Figure PS.S Distribution ofcapillary-pressure head (let
f ) and volumetric
moisture content (right) for Problem 8.2, part B.
421
70 .....,. --, 70 , ,-
...,
-, -
__ _ _ _ _ _ _
____
30-,
-L __ ...
, _ ---,
, __ �
-7 -<l -5 20 30 40 50 60 70
Capillary-pressure Hydraulic head (m)
head, \11 1m)
Figure PS.6 Distribution of capillary-pressure head (left) and hydraulic
head (right) for Problem 8.2, part C.
Problem 8.3. Assume that a wetting front moves into the sandy loam soil
described in Problem 8.2. The moisture content at the surface of the soil is
held constant at 27%. The underlying moisture content is constant al 6%.
If the saturaled hydraulic conductivity for this soil is 3 x 10--6 m 5-1 ,
estimate how long it will take the wetting front to move 1 m into the soil.
Chapter 9
422
root zone are by evapotranspiration. Assuming that no rainfall occurs, the
temporal dynamics of depth-average soil moisture 0 in the root zone can
be expressed by the soil wmer balancc equation.
dO
Z-=-el'
dl
where Z is the depth of the root zone. In other words, changes in the water
stored in the control volume (i.e., the root zone) are due to the difference
betwcen water inputs and outputs (sec Chapter I). In this specific case
there are no rainfall-induced inputs, while the only outflow is due to el.
The irrigation period (i.e., the time between two consecutive applications
of irrigation water) is calculated by integrating the soil water balance
equation between time' = 0 when 0 = Ofc (i.c., right after irrigation) and
time 'i when 0 = 00 (and it is time to irrigate again), assuming that the
evapotranspiration rate linearly decreases from the potential rate PET at 0
= O/<: to zero at 0 = Ow (i.e., et = PET(O-O)..)J(Oc
j -O ».
...
Consider the ease of a crop grown on a sandy soil with Ofc = 0.19 and a
crop-specific wilting point, Ow = 0.05. The root zone is 40 cm deep. After
cach application of irrigation water, soil moisture is equal to Ofe. Assuming
that PET = 6 mm d- 1, calculate the irrigation pcriod, Ii' i.e., how long it
would take for soil moisture to decrease from field capacity to the value 00
= 20", in the absence of any rainfall input.
Using the above equation with PET = 6 rum d-1, Of<. = 0.19, OK' =
0.05,00 = 20 , and Z = 0.40
... 01, we find: I;::::; 10 days.
423
increased from 00 = 20", to
0fc' The volume of water (per unit area)
required to make this cbange is Z{(Jjc - 00) = 36 mm. Because of
evaporation losses the irrigation method has an 80% efficiency (i.e.,
only 80% of the irrigation water infiltrates). Thus the amount of
water that needs to be spent in irrigation is 36/0.80 = 45 mm every
10 days.
Chapter 10
424
If K decreases exponentially with depth, the data should define a
straight line on a plot of K on a logarithmic scale and z 011 a linear
scale. The relationship can be inspected visually and the
parameters of Equation 10.13 (Ko and}) can be found as the slope
and intercept of a line drawn through the data (Figure PS.7).
10-'-.-----
,.,
-
.� 10-3-_ o
>
o
.-
tJ
::I
't:1
r::
o Slope of line if) -2.2 m-1
o =
.!1 10-4.
10-5-��� �������� �
o 0.5 1 1.5
Soil depth (m)
Figure PS.7 Data and fitted line for Problem 10,2.
Segment 1 Segment 2
Calculate the topographic index for each segment and indicate which
segment is more likely to produce satumtion-excess overland flow.
425
TI = In(a/tan fl).
Problem 10.4. The soil ofa given catchment has a porosity (t/J) of 0.4 and
hydraulic conductivity as described in Problem 10.2. The catchment has an
average value of the topographic index of 3.5. Under conditions where the
average saturation deficit of the catchment is 100 mm, calculate the
subsurface flow to the stream, qSllhs!lrJace.lfthe throughfall rate at this lime
is 4 x 10-3 mm S-I ( 1 4 mm hr-1), what fraction of the catchment would
have to be saturated if saturation-excess overland flow were to be equal to
one-half of the subsurface flow?
TIHu.� = K.lf= 1.85 x 10-3 m s-1/2.2 m-1 = 8.4 x 10-4 m2 S-l. Thc soil
parameter, m, is ,plf= 0.4/2.2 m-I = 0.2m. Thus, CJsubsurJac£= 1.5 x lO
I l
sm S- = I.S x 10-2 mOl S- .
[\'cn if the entire catchment were
saturated (in which case -; would be zero, not 100 mOl) the
contribution of overland flow cannot be half of the subsurface flow.
426
Glossary
alternate depths the two physically allowable water depths for flow in a
channel for a given specific energy and a given specific discharge.
{Section 4.2}
anisohydric plants plants that lend to maintain the stomata open and
sustain higher photosynthetic uptakes even when soil moisture is low.
{Section 9.2. 1 }
427
body to process waste discharges such that they are harmless. {Section
l.2}
average linear velocity v=q/¢ [L II], the average velocity of fluid within
the pores of a porous medium, equal to the specific discharge divided
by the porosity. {Section 6.3.1 and Section 7.6}
basin aspect ratio the ratio of basin length (the direction parallel to flow)
to basin depth or thickness (above a low-pemleability unit). { Section
7.3.1}
body forces forces that act unifonnly on each fluid element; examples
inelude gravitational forces and electromagnetic forces. { Section 3.3}
Bowen ratio B [dimensionless], the ratio of the sensible heal flux (H) to
the lafell1 Ilea/jlux (E,). { Section 2.4.2}
capillary forces forces that are exerted on soil water due to the strong
attraction of water by soil minerals; see capillmy-pressllre head.
{Section 8.2.1}
capillary fringe the zone immediately above the water table in which the
pores are filled with water but the water is under pressure less than
atmospheric. {Section 8.11
428
cavitation fonnation of vapor bubbles in (liquid) water under low
(negative) pressure (i.e., strong suction). { Section 9.2.1}
conservation of mass dMldt= I' - 0', the law that states that for any
particular compartment (usually referred to as a control volume), the
time rate of change of mass stored within the compartment is equal to
the difference between the inflow ratc and the outflow rate. { Section
1.4}
contributing area the area of catchment upslope from a given block that
contributes inflow to that block. { Section 10.5.1}
429
critical flow flow that occurs at the minimum value of specific energy
allowed for a given specific discharge. The Froude number equals I
for critical flow. { Section 4.2.2}
elevation head z [L], a component of the total head that may bc thought of
as the potential energy per unit fluid weight. {Section 3.S.2}
430
cvaporation the physical process involving a phase change from liquid to
vapor by which water is returned to the atmosphere. { Sections 2.1 and
2.4)
field capacity the relatively constant moisture content that a sandy soil
tends to attain following drainage. { Section 8.1 }
free surface the upper boundary of an open channel flow, between the
watcr and thc atmosphcre. { Scction 4.2}
431
friction slope the slope of the imaginary line that is a distance U2/2g
above the water surface in an open channel. {Section 4.4}
head loss ilL [L}, the loss in head due to friction within the fluid and
between the fluid and the side walls of a pipe or channel. {Section 3.6}
hollow-stem auger auger for drilling wells that has a hollow core that
allows placement ofa casing and construction of the well inside the
augers, which are then removed, letting the casing remain in place.
{Section 6.4. 1 }
432
or ill1rinsic permeability; adjectival form is homogeneous. {Sections
3.2 and 6.5.1}
hydraulic head h =(p/pg)+z [L), the mechanical energy per unit fluid
weight, used in the study of flow through porous media. { Section 6.2}
hydraulic lift nocturnal transfer of water from the deep (and wetter) to the
shallow (and drier) soil through the root system. {Section 9.2.3}
433
pressure and depth in a fluid at rest. For a homogeneous fluid, the
hydrostatic equation is p = pgd. { Section 3.4}
infiltration the movement of rain or melting snow into the soil at the
Earth's surface. { Section 8.8}
infiltration capacity the maximum rate at which water can infiltrate into a
soil. { Section 8.8}
434
intrinsic permeability k = K(jI/pg) [L2], the ability of a porous medium to
transmit fluid, independent of the fluid properties. { Section 6. 3. 1 }
isohydric plants plants that tend to close the stomata and reduce
photosynthesis as the soil becomes drier. { Section 9.2 . 1}
latent heat [M L2 ,2], the portion of the inrernal ellergy ofa substance
that cannot be "sensed" (i.e., is not proportional to absolute
temperature); latent heat is the inlemal energy that is released or
absorbed during a phase change at constant temperature. { Section
2.4.2}
latent heat flux £, [M ,3], the rate per unit area at which latent heal is
transferred to the atmosphere. { Section 2.4.2}
latent heat of vaporization ).� [L2 ,2], the amount of energy per unit
mass absorbed during a phase change from liquid to vapor at constant
temperature. For evaporation of water at O°C, ...1.v = 2.5 x 106 J kg-I .
{Section 2.4.2}
4 35
leaky aquifer an aquifer that is not perfectly confined but that has leakage
across the surrounding confining layers. { Section 7.4.2}
local acceleration the change of velocity with time at a fixed location, aliI
at [L " l. (Section 3.5.1 }
natural basin yield the average rate of discharge From a basin under
natural or undisturbed conditions (i.e., in the absence of anthropogenic
groundwater withdrawals or changes in climate or vegetation).
{Section 7.2}
436
normal stress a force per unit area oriented perpendicular to a surface of a
fluid or solid object. Pressure is a nonnal stress. {Section 3.3}
permanent wilting point the driest soil moisture conditions that a plant
can withstand without wilting. { Section 9.2. 1 }
permeameter a device used to measure the flow rate through and the
hydraulic conductivity of a porous medium. {Section 6.3.1 }
phreatophytes deep rooted plants that draw part of their water from
beneath the water table. { Section 9.2.2}
437
thereby hydraulic head) at a point in the subsurface. {Section 6.4}
pressure p [M L-] [2], the force per unit area acting perpendicular to a
surface, or normal stress. { Section 3.3}
pressure head p/pg [L], a component of the total head that may be thought
of as the "flow work" or the work due to pressure per unit fluid weight.
{Section 3.5.2}
438
pumping test a technique to estimate aquifer transmissivity and storativity
by pumping water from onc well and obscrving induced changcs in
watcr level ovcr time in anothcr wcll. {Scction 7.S.2}
rain shadow low rainfall area on the leeward side ofa mountain range.
{ Section 2.2}
return flow the process by which groundwater reemerges from the soil at
a saturated area and flows downslope as overland flow. { Section
IOAA}
return period Trel,,,,, [T], a measurc of how often (on averagc) an cvent
(precipitation, flood, etc.) will occur that is greater than some chosen
valuc; the invcrsc of the exceedancc probability. { Section 2.2.3}
439
high Reynolds numbers (>4000 for pipe flow) are turbulent. Flows of
different fluids with the same Reynolds number will be similar.
{ Section 3.7}
runoff ratio the ratio of average annual surface runoff to average annual
precipitation for a given land area V/"j5) { Section 1.4. 1 }
440
as when a perched water table forms above a layer of tile soil with low
permeability; some of the flow may occur along preferred pathways
known as macropores. { Section IOA.3}
shear stress t [M L-I 'I2J, a langential force per unit area acting on the
surface ofa solid or fluid. { Sections 3.2 and 3.3}
soil horizon a soil layer defined on the basis of physical and chemical
properties and on the history of its fonnation. { Section 8.9}
soil moisture water that is held in soils and rocks under pressures less than
atmospheric; water in the unsaturated zone. { Section 8. 1 }
soil water potential lis [L], the hydraulic head of soil water. {Section
9.2.1 }
solar energy energy deriving from radiation from the sun. It is the driver
of the hydrological cycle. { Section 1 .3 }
specific discharge (open channel flow) the discharge per unit width of
channel in a rectangular open channel, q", = Qlw = Uil [L2 ' ].
11
{ Section 4.2}
specific energ.y [L], the energy per unit weight of water in a channel with
respect to the channel bottom as datum. The total energy H [L] is the
sum of tile bottom elevation and the specific energy. { Section 4.2 }
specific energy diagram a graph of specific energy versus water depth for
441
a given specific discharge. The diagram shows the physically
allowable water depths for a given specific discharge and a given
specific cncrgy. { Scction 4.2}
steady now a flow that is constant in time at each location in the flow. [n
steady flow, the local acceleration is zero. { Section 3.5.1 }
stress a force per unit area (S[ units: N m-2 or Pal. { Section 3 .3 }
442
sublimation the physical process by which water in the solid phase
changes to water vapor and is directly returned to the atmosphere.
{ Section 2.4.2}
surface forces forces that act through direct contact on specific surfaces of
a fluid or solid body. { Section 3.3}
surface runoff waler from rainfall or snowmelt thaI runs over the surface
of the Earth in sheets, rivulets, streams, and rivers. { Section 1 .3 }
total head H [L], the sum of the pressure, elevation, and velocity heads in
a frictionless fluid. { Section 3.5.2}
total stress (Ir [M L-I [2], the weight (a force) of material overlying a
plane of unit cross-sectional area in the subsurface. {Section 7.4.2 }
443
transpiration the physical process by which water changes phase from
liquid to vapor, is released through the stomata ofa plant, and returns
to thc atmosphcre. { Section 2.1 }
uniform flow a flow that does not change from place to place along the
flow path; the convective acceleration is zero. { Section 3.5.1}
unit weight y[M L-2 T- 2], the gravitational force per unit volume, pg,
acting on a fluid or solid (SI units: N m-3). { Section 3.2}
unsaturated zone the zone in a soil or rock between the Earth's surface
and the waleI' table; pores in the unsaturated zone are partly filled with
water and partly filled with air. { Section 8.1 }
variable contributing area concept the idea that catchment areas where
saturation-excess overlandjlolV develops expand and contract with
time over a stonn. {Section 10.4.2}
velocity head [L], a component of the total head that may be thought of as
the kinetic energy per unit fluid weight. {Section 3.5.2}
444
watcr budgct a calculation of the inflows, outflows, and change in storage
for a particular control volume (such as a lake or a catchment) over a
particular timc pcriod. { Section 1 .4}
waterlogging saturation ofthc shallow soil rcsulting from the rising of the
water lable to the ground surface. {Section 9.2.2 }
water vapor density [M L-3], mass of water vapor per unit volume.
{ Scction 2.4.2}
weir an artificial obstruction such as a stcp or dam ovcr which all thc
water in a channel must flow and that can be used 10 measure stream
discharge. { Section 4.3 }
well casing a pipe, typically steel or PYC, that serves to line a well.
{ Section 6.4. 1 }
well screen a section at the end of a well casing that allows water to enter
the wcll; construction may be holes drilled or slots cut in a pipe or may
be a wire mesh material. { Section 6.4. 1 }
445
References
446
Brown A.E., L. Zhang, T.A. McMahon, A.W. Western, and R.A. Vertessy.
2005. A review of paired catchment studies for determining changes in
water yield resulting from alterations in vegetation. Journal oj
Hydrology 3 10:28-6 1 .
Brutsaert, W. 1982. Evaporation into the atmosphere. Dordrecht, the
Netherlands: D. Reidel.
Brutsaert, W. 2005. Hydrology: An introduction. New York: Cambridge
University Press.
Burgess, S.S.O., M.A. Adams, N.C. Turner, and e.K. Ong. 1998. The
redistribution of soil water by tree root systems. Oec% gia 1 15:306--
311.
Bums, D.A., J.A. Lynch, B.J. Cosby, M.E. Fenn, and J.S. Baron, and US
EPA Clcan Air Markets Div. 20 1 1 . National Acid Precipitation
Assesslllelll Program repor/ IO Congress 201 I: An il/legrated
assessment. Washington, DC. National Science and Technology
Council.
Caldwell, M.M., and J.H. Richards. 1989. Hydraulic lift water effiux from
upper roots improves cffectiveness of water uptake by deep roots.
Oecolagia 79: 1-5.
Campbell, G.S., and J.M. Norman. 1998. All illlroeil/ction to ellvironmelllal
biophysics. New York: Springer-Verlag.
Carr, J.A., P. D'Odorico, F. Laio, and L. Ridolfi. 2012. On the temporal
variability of the virtual water network. Geophysical Research Lerrers.
doi: 39, L06404.
Chanson, H. 2004. Environmental hydralllics ofopen channelflolVs.
Oxford: Elsevier-Butterworth-Heinemann.
Childs, E.C. 1969. The physical basis ofsoil waleI' phenomella. New York:
John Wiley & Sons.
Christenson, S.c., and 10M. Cozzarelli. 2003. The Norman Landfill
Environmental Research Site: What happens to the waste in landfills?
U.S. Geological Survey Fact Sheet 040-03. http://pubs.usgs.gov/fs/fs-
040-03/.
Clapp, R.B. 1977. Infiltration in relation to rainfall intensity: An
investigation of an empirical equation using simulated data. M.S.
Thesis, Department of Environmental Sciences, University of Virginia.
Cooper, H.H., and C.E. Jacob. 1946. A generalized graphical method for
evaluating fonnation constants and summarizing well field history.
Transactiolls oJthe American Geophysical Ullioll 27:526-534.
Cosby, 8.J., G.M. Hornberger, D.M. Wolock, and P.F. Ryan. 1 987.
Calibration and coupling of conceptual rainfall-runoff/chemical flux
447
models for long-term simulation of catchment response to acidic
deposition. In Systems analysis in water quality management. edited by
M.B. Back, pp. 1 5 1-160. Oxford: Pcrgamon Prcss.
Court, A. 1960. Reliability of hourly precipitation data. JOllmal of
Geophysical Research 65:4017-4024.
Darcy, H. 1 856. LesJon/ailles pllbliqlles de la ville de Dijoll. Paris: Victor
Dalmont.
DePaul, V.T., D.E. Rice, and O.S. Zapecza. 2008. Water-level changes in
aquifers of the Atlantic Coastal Plain, predevelopment to 2000: U.S.
Gcological Survey Scientific Investigations Report 2007-5247.
DeWalle, D., and A. Rango. 2008. Principles ojsnow hydrology. New
York: Cambridge University Press.
Dietrich, W.E., c.J. Wilson, D.R. Montgomery, J. McKean, and R. Bauer.
1992. Erosion thresholds and land surface morphology. Geology
20,675-679.
D'Odorico, P., F. Laio, and L Ridolfi. 2010. Does globalization of water
reduce societal resilience 10 drought? Geophysical Research Letters
37,L i3403.
Domenico, P. A., and F.W. Schwartz. 1990. Physical and chemical
hydrogeology. New York: John Wilcy & Sons.
Dunnc, T, and R.D. Black. 1970. Partial area contributions to stonn runoff
in a small New England watershed. Waler Resources Research 6: 1 296-
131 I .
Dunne, T, and LB. Leopold. 1978. Water ill ellvironmental planning. San
Francisco: W. H. Freeman.
Dunne, T, T.R. Moore, and C.H. Taylor. 1975. Recognition and prediction
of nmofT-producing zones in humid regions. Hydrological Sciences
Bulletin 20:305-327.
Eaglcson, P.S. 1970. DYllamic hydrology. Ncw York: McGraw-Hill.
Eltahir A.B.B., and R.L. Bras. 1996. Precipitation recycling. Reviews oj
Geophysics 34:367-378.
Falkcnmark, M., and J. Rockstrom. 2006. The new blue and grcen water
paradigm: Breaking new ground for water resources planning and
management. Journal oj Water Resources Planning and Managemellt
I EI29-132.
Falkenmark, M., A. SemIen, A. Jagerskog, J. Lundqvisl, M. Matz, and H.
Tropp. 2007. 011 the verge ofa lIew water scarcity: A callJor good
governallce alld human ingenllity. SIWI Policy Brief.
Falkenmark, M., and J. Rockstrom. 2004. Balancillg walerJor humalls alld
nature. London: Earthscan.
448
Food and Agriculture Organization of the United Nations (FAO). 2010.
Global forest resource assessment 20 I 0: Key findings.
hllp:llwww.fao.orglforcstry/fra/fra20 I Olen!.
Fedoroff, N.V., D.S. Battisti, R.N. Beachy, P.J.M. Cooper, D.A. Fischhoff,
C.N. Hodges, V.c. Knauf, D. Lobell, B.J. Mazur, D. Molden, M.P.
Reynolds, P.c. Ronald, M.W. Rosegrant, P.A. Sanchez, A. Vonshak, J.
K. Zhu. 20 I O. Radically rethinking agriculture for the 21st century.
Sciellce 327:833-834.
Feller C.W. 2000. Applied hydrogeology, 4th cd. Upper Saddle River, NJ:
Prentice Hall.
Foley J.A., N. Ramankutty, K.A. Brauman, E.S. Cassidy, J.S. Gerber, M.
Johnston, N.D. Mueller, C. O'Connell, O.K. Ray, P.c. West, C. Balzer,
E.M. Bennett, S.R. Carpenter, J. Hill, C. Monfreda, S. Polasky, J.
RockstTom, J. Sheehan, S. Siebert, O. Tilman, and D.P.M. Zaks. 201 1 .
Solutions for a cultivated planet. Nature 478:337-342.
Fraedrich, K, A. Kleidon, and F. Lunkeit. 1999. A green planet versus a
desert world: Estimating the effect of vegetalion extremes on the
atmosphere. Journal of Climate 12: 3 1 56-3 163.
Freeze, R.A. 1971. Three-dimensional, transient, saturated-unsaturated
flow in a groundwater basin. Water Resources Research 7:929-941 .
Freeze, R.A. 1972a. Role of subsurface flow in generating surface mnoff:
1 . Baseflow contributions to channel flow. Wafer Resources Research
8:609-623.
Freeze, R.A. 1972b. Role of subsurface flow in generating surface runoff:
2. Upstream source areas. Water Resources Research 8: 1 272-1283.
Freeze, R.A., and J.A. Cherry. 1979. Groundwater. New York: Prentice
Hall.
Freeze, R. A., and P. A. Witherspoon. 1966. Theoretical analysis of
regional groundwater flow: 1 . Analytical and numerical solutions to the
mathematical model. Water Resources Research 2:641--656.
Freeze, R. A., and P. A. Witherspoon. 1967. Theoretical analysis of
regional groundwater flow: 2. Effect of water-table configuration and
subsurface penneability variation. Wafer Resources Research 3:623-
634.
Freeze, R. A., and P. A. Witherspoon. 1968. Theoretical analysis of
regional groundwater flow: 3. Quantitative interpretations. Water
Resources Research 4:581-590.
Gedney, N., P.M. Cox, R.A. Betts, O. Boucher, C. Huntinglford, and P.A.
Stott. 2006. Detection of a direct carbon dioxide effect in continental
river runoff records. Nature 439:835-838.
449
Giordano, M. 2009. Global groundwater? Issues and solutions. Annual
Review ofEnvironment and Resources 34: 153-178.
Gleiek, P.H. (Editor). 1993. Water in crisis: A guide to the world 'sfresh
water resources. New York: Oxford University Press.
Green, W.H., and G.A. Ampt. 191 1. Studies on soil physics: I . The flow
of air and water through soiIs. Journal ofAgric/lllllral Science 4: 1-24.
Guymon, G.L. 1994. Unsaturated zone hydrology. Englewood Cliffs, NJ:
PTR Prentice Hall.
Haan, c.T. 2002. Statistical methods ill hydrology, New York: J. Wiley &
Sons.
Hamilton, P.A., and R.J. Shedlock. 1992. Arefertilizers andpesticides in
the groundwater? A case study ofthe Delmarva Peninsula: Delm'v'are.
Marylalld. and Virginia. U.S. Gco[ogica[ Survey Circular [080.
Harr, R.D., 1977. Water flux in a soil and subsoil on a steep forested slope.
Journal of Hydrology 33:37-58.
Harris, D.M., J.J. McDonnell, and A. Rodhe. 1995. Hydrograph separation
using continuous open system isotope mixing. Water ResOllrces
Research 3 1 : 1 57-1 7 1 .
Heidmann, L.J., M.G. Harrington, and R.M. King. 1990. Comparisoll of
moisture retentioll cun'es in representative basaltic and sedimentmy
soils ill Arizona prepared by two methods. Research Note RM·500,
USDA Forest Service, Rocky Mountain Forest and Range Experiment
Station.
Helm, D.C. 1982. Conceptual aspects of subsidence due to fluid
withdrawal. In Recent trends in hydrogeology, edited by T. N.
Narasimhan. Geological Society of America Special Paper 189, pp.
103-139.
Helton, J.c., D. R. Anderson, M.G Marietta, and R. Rechard. 1997.
Perfonnance assessment for the waste isolation pilot plant: From
regulation to calculation for 40 CFR 1 9 1 . 1 3 . Operations Research
45,1 57-177.
Hoekstra, A.Y. and A.K. Chapagain. 2008. Globalization ofn;ater:
Sharing the planet'sfreshwater resources. Malden, MA: Wiley·
Blackwell.
Hornberger, G.M., K.E. Bencala, and D.M. McKnight. 1994. Hydrological
controls on dissolved organic carbon during snowmelt in the Snake
River near Montezuma, Colorado. Biogeochemistly 25: 147-165.
Hornberger, G.M., KJ. Beven, and P.F. Gennann. 1990. Inferences about
solute transport in macroporous forest soils from time series models.
Geoderma 46:249-262.
450
Hamberger, G.M., and E.W. Boyer. 1995. Recent advances in watershed
modeling. In u.s. Natiollal Report to IUGG. 1991-1995. Contributions
/11 Hydrology. pp 949-957. Washington: American Geophysical Union.
Horton, R.E. 1933. The role of infiltration in the hydrologic cyclc.
Transactiolls ojthe Amer/can Geophysical Vllioll 14:446-460.
Hsieh, P.A. 200 1 . Topodrive and Particleflow: Two computer models for
simulation and visualization of groundwater flow and transport of fluid
particles in two dimensions. U.S. Geological Survey Open File Report
0 1 -286.
Hubbert, M.K. 1940. The theory of ground-water motion. Journal oj
Geology 48:785-944.
Jonkman, S.N. 2005. Global perspectives on loss of human life caused by
floods. Nat/lral Hazards 34: 1 5 1-175.
Jury, W.A., and H. Vaux. 2005. The role of science in solving the world's
emerging water problems. Proceedings oJthe National Academy oj
Sciences (PNAS) \02: 1571 5-15720.
Jury, W.A., and H.J. Vaux, Jr. 2007. The emerging global water crisis:
Managing scarcity and conflict between water users. Advances in
Agronomy 95: 1-76.
Katul, G.G., R. Oren, S. Manzoni, C. Higgins, and M.B. Parlangc. 2012.
Evapotranspiration: A process driving mass transport and energy
exchange in the soil-plant-atmosphere-climate system. Reviews 0/
Geophysics 50:RG3002. doi: I 0.1 029/20 1 1 RG000366.
Klein, G. 2005. Califomia's Water-Energy Relationship. California
Energy Commission, Final Staff Report, CEC-700-2005-0 I I-SF.
Kohn, M.S. 2012. Bathymetry of Totten Reservoir, Montezuma County,
Colorado, 20 I I : U.S. Gcological Survcy Scientific Invcstigations Map
3203. hup://pubs.usgs.gov/sim/3203/.
Konikow, L. F. 201 1 . Contribution of global groundwater depletion since
1900 to sea-level rise. Geophysical Research Lef/ers 38:L 17401.
Konikow, L.F., T.E. Reilly, P.M. Barlow, and c.1. Voss. 2006.
Groundwater modeling. In The handbook o/groundwater engineering,
2nd ed., edited by J. W. Del1eur, Chapter 23. Boca Raton: CRC Press.
L'vovich, M.I. 1979. World water resources and tileirjilfllre. translated by
R.L. Naec. Washington: American Geophysical Union.
Labat, D., Y. Godderis, J.L. Probst, and J.L. Guyo!. 2004. Evidence for
global runoff increase related to climate warming. Adval/ces ill Waler
Resources 27:63 1-642.
Leakey, A.D.B, E.A. Ainsworth, C.J. Bamacchi, A. Rogers, S.P. Long,
451
and D.R. Ort. 2009. Elevated CO2 effects on plant carbon, nitrogen, and
water relations: Six important lessons from FACE. JOIIl"l1al of
Experimental Botany 60:2859-2876.
Lee, J.-E., R.S. Oliveira, T.E. Dawson, and I. Fung. 2005. Root
functioning modifies seasonal climate. Proceedings ofthe National
Academy ofSciences 102: 1 7576-1 7581.
Leopold, L.B. 1994. A view ofthe river. Cambridge, MA: Harvard
University Press.
Leopold, L.B., and M.G. Wolman. 1957. River channel patterns: Braided.
meandering, and straight. U.S, Geological Survey Professional Paper
282-B, pp. 39-85.
Leopold, L.B., M.G. Wolman, and J.P. Miller. 1964. Fluvial processes in
geomOlp!lOlogy. San Francisco: W. H. Freeman.
Luckey, R.R., and Becker, M.F. 1999. Hydrogeology, water use, and
simulation of flow in the High Plains aquifer in northwestern
Oklahoma, southeastern Colorado, southwestern Kansas, northeastern
New Mexico, and northwestern Texas: U.S. Geological Survey Waler
Resources Investigations Report 99-4104.
Maidmenl, D.R. 1993. Hydrology. In Handbook ofHydrology, edited by
D. R. Maidment, pp. 1 . 1-1.15. New York: McGraw-Hill.
Margat, J., S. Foster, and A Droubi. 2006. Concept and importance of
non-renewable resources. In NOI/-renewable Groundwater Resources: A
Guidebook 01/ Socially-SuslailJable Managemclllfor Water-Policy
Makers. edited by S. Foster, and D. P. Loucks, IHP-VI Ser.
Groundwater, vol. 10, pp. 1 3-24. Paris: UNESCO.
McDowell, N., W.T. Pockman, C.D. Allen, D.O. Breshears, N. Cobb, T.
Kolb, J. Klaut, J. Sperry, A. West, D.G. Williams, and E.A. Yepez.
2008. Mechanisms of plant survival and mortality during drought: Why
do some plants survive while others succumb to drought? New
Phytofogist 1 78:719-739.
McGuire, V.L. 201 1 . Water-level changes in the High Plains aquifer,
predevelopment to 2009, 2007-{)8, and 2008-{)9, and change in water
in storage, predevelopment to 2009: U.S. Geological Survey Scientific
Investigations Report 20 I \-5089, hup://pubs.usgs.gov/sir/20 \ 1/5089/.
Meehl, G.A., T.F, Stocker, W.O. Collins, P. Friedlingstein, AT Gaye,
J.M. Gregory, A Kitoh, R, Knutti, J.M. Murphy, A Noda, S,C.B.
Raper, I.G, Watterson, AJ, Weaver, and Z.-c. Zhao, 2007, Global
climate projections. In Climate Cha/Jge 2007: The Physical Science
Basis. Contribution of Working Group I fo the Fourth Assessment
Report ofthe Illfergo�'erl/lnen1a1 Panel 01/ Climate Challge. edited by S.
452
Solomon, D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M.
Tignor and H.L. Miller. Cambridge, UK: Cambridge University Press.
Meinzer, O.E. 1923. The occurrence 0/ground lVater in the United States.
U.S. Geological Survey Water-Supply Paper 489.
Mekonnen, M.M., and AV. Hoekstra. 2010. The green, blue and grey
water footprint of crops and derived crop products. Value of Water
Research Report Series No. 47, Delft, the Netherlands: UNESCO-lHE.
hup://www.waterfootprint.orglReportslReport47-WaterFootprintCrops
Voll .pdf.
Mekonnen, M.M., and AY. Hoekstra. 2012. The Blue Water footprint of
electricity from Hydropower. Hydrology and Earth System Sciel/ces
1 6 ; 1 79-1 87.
Miah, M.M., and K. R. Rushton. 1 997. Exploitation of alluvial aquifers
having an overlying zone of low permeability: Examples from
Bangladesh. Hydrological Sciel/ces JOllrnaI 42:67-79.
Milly, P.C.D., K.A Dunne, and A.V. Vecchia. 2005. Global pattern of
trends in streamflow and water availability in a changing climate.
Natllre 43S:347-350.
Moile, F, P. Jayakody, R. Ariyaratne, and H.S. Somatilake. 200S.
Irrigation versus hydropower: Sectoral conflicts in southern Sri Lanka.
Water Policy 1 0 (Supplement I ) : 37-50. doi: 1 O.2166/wp.200S.051 .
Moran, M.S., and R.D. Jackson. 1991. Assessing the spalial distribution of
evapotranspiration using remotely sensed inputs. Journal of
Environmental Quality 20:725-737.
National Research COllncil (NRC). 1 995a. Flood risk management and the
American River Basill. Washington: The National Academies Press.
National Research Council (NRC). 1 995b. Mexico City's lVater supply:
Improving the olltlookfor sllstainability. Washington: The National
Academies Press.
National Research Council (NRC). 1 995c. Ward Valley: An e.xaminatioll
o/seven issues in earth sciences and ecology. Washington: The
National Academies Press.
National Research Council (NRC). 2000. Seeing illto the Earth:
Noninvasive characterization a/the shallolV slIbsllljacefor
envirOl/mental alld el/gineering applications. Washington: The National
Academies Press.
National Research Council (NRC). 2007. Elevation daraforfloodplain
mapping. Washington: The National Academies Press.
National Research Council (NRC). 2012a. Challenges and opportunities in
the hydrologic sciences. Washington: The National Academies Press.
453
National Research Council (NRC). 2012b. Climate change: Evidence.
impacts. alld choices. Washington: The National Academics Press.
National Research Council (NRC). 2012c. Water reuse: Potential/or
expanding the natioll 's water supply through reuse a/municipal
wastewater. Washington: The National Academies Press.
Ortega-Guerrero, A., J.A. Cherry, and D.L. Rudolph. 1993. Large-scale
aquitard consolidation near Mexico City. Ground Waler 3 I :708-71 8 .
Perrone, D., and G.M. Hornberger. 2014. Water, food, and energy
security: Scrambling for resources or solutions? WIREs Water 1 :49--68.
doi: 10. I 002/wat2. I 004.
Qu, Y., and C.J. Duffy. 2007. A semidiscretc finite volume formulation for
mUltiprocess watershed simulation. Water Resources Research
43:W0841 9. doi: l 0.1 029/2006WR005752.
Robson, A., K.J. Beven, and C. Neal. 1992. Towards identifying sources
of subsurface flow: A comparison of components identified by a
physically based runoff model and those determined by chemical
mixing techniques. Hydrological Processes 6: 199-2 14.
Rockstrom, J., M. Lannerstad, and M. Falkcnmark. 2007. Assessing the
water challenge of a new green revolution in developing countries.
Proceedings a/the Nationul Academy a/Sciences (PNAS) 104:6253-
6260.
Rosegrant, M.W., C. Ringler, and T. Zhu. 2009. Water for agriculture:
Maintaining food security under growing scarcity. Anllllal Review 0/
Ellvirollmell1 alld Resources 34:205-222.
Runyan, C.W., P. D'Odorico, and D. Lawrence. 2012. Physical and
biological feedbacks on deforestation. Reviews a/Geophysics
50:RG4006.
Scanlon, B. R., I. Jolly, M. Sophoc1eous, and L. Zhang. 2007. Global
impacts of conversions from natural to agricultural ecosystems on water
resources: Quantity versus quality. Waler Resources Research
43:W03437.
Scanlon, B.R., K.E. Keese, A.L. Flint, L.E. Flint, c.B. Gaye, W.M.
Edmunds, and I. Simmers. 2006. Global synthesis of groundwater
recharge in semiarid and arid regions. Hydrological Processes
20:3335-3370.
Scholl, M.A .. and S.c. Christenson. 1998. Spatial variation in hydraulic
conductivity determined by slug tests in the Canadian River alluvium
near the N0n11an Landfil1, Norman, Oklahoma. Water Resources
Investigations Report 97-4292.
Shah, T., D. Moldcn, R. Sakthivadivel, and D. Seckler. 2000. The global
454
groundwater simatiol/: Overview ofopportunities and challenges.
Colombo, Sri Lanka: International Water Management Institute.
Shiklomanov, I.A. 1999. World water resources and their use: A joint
SHlIUNESCO product.
http://webworld.unesco.orglwalerJihpJdb/shiklomanov/index.shlml.
Siebert, S., J. Burke, J.M. Faures, K. Frenken, J. Hoogeveen, P. 0611, and
F.T. Portmann. 2010. Groundwater use for irrigation: A global
inventory. Hydrology alld Earth System Sciences 14: 1 863-1880.
Sklash, M.G., and R.N. Farvolden. 1979. The role of groundwater in storm
runoff. Journal ofHydrology 43:45-65.
Skopp, J., M.D. Jawson, and J.W. Doran. 1990. Steady-state aerobic
microbial activity as a function of soil water content. Soil Sciellce
Society ofAmerica J01/rnal 54: 1 6 1 9-1625.
Smith, J.A., M.L. Baeck, M. Steiner, and A.J. Miller. 1 996. Catastrophic
rainfall from an upslope thunderstorm in the central Appalachians: The
Rapidan storm of June 27, 1995. Water Resources Research 32:3099-
3 1 13.
Smith, J.A., OJ. Seo, M.L. Baeck, and M.D. Hudlow. 1996. An
intercomparison study ofNEXRAD precipitation estimates. Water
Resources Research 32:2035-2045.
Smith, T.R. 1 974. A derivation of the hydraulic geometry of steady-state
channels from conservation principles and sediment transport laws.
Journal of Geology 82:98-104.
Soeder, D.J., J.P. Raffensperger, and M.R. Nardi. 2007. Effects of
withdrawals on ground-water levels in southern Maryland and the
adjacent Eastern Shore, 1980-2005. U.S. Geological Survey Scientific
Investigations Report 2007-5249.
Stednick, J.D. 1996. Monitoring the effects of timber harvest on annual
water yield. Journal of Hydrology 176:79-95.
Tilman D., C. Balzer, J. Hill, and B.L. Befort. 201 1 . Global food demand
and the sustainable intensification of agriculture. Proceedings ofthe
National Academy ofSciences (PNAS) 108:20260-20264.
T6th, J. 1962. A theory of groundwater motion in small drainage basins in
central Alberta, Canada. Journal of Geophysical Research 67:4375-
4387.
T6th, J. 1963. A theoretical analysis of groundwater flow in small drainage
basins. Jo/(rnal of Geophysical Research 68:4795-4812.
Wada, Y., L.P.H. van Beek, C.M. van Kempen, J.W.T.M. Reckman, S.
Vasak, and M.F.P. Bierkens. 2010. Global depletion of groundwater
resources. Geophysical Research Letters 37:L20402.
455
Winograd, 1.1. 1981. Radioactive waste disposal in thick unsaturated
zones. Science 2 1 2 : 1457-1464.
Winter, T.e. 1983. The interaction of lakes with variably saturated porous
media. WaleI' Resuurces Research 19: 1203-1 218.
Wolman, M.G. 1955. The natural channel oj Brandywine Creek
Pennsylvania. U.S. Geological Survey Professional Paper 27 1 .
Wolock, D.M. 1993. Simulating the variable-source-area concept oj
streamflow generation with the watershed model TOPMODEL. U.S.
Geological Survey Water-Resources Investigations Report 93-4124.
456
Index
457
Average linear velocity, 1 54, 199
458
tensiometer based measurement of, 222, 222
vertical flow and, 220-2 2 1 , 222-223
Capillary tube, 147-148, 148
Carbon assimilation, 241 , 244
Carbon dioxide, atmospheric, 241 , 248, 290-291, 291-294, 293
Catchment, 1 3-14, 15
Coweeta, North Carolina, 259
Hubbard Brook, New Hampshire, 258
Catchment divide. See Divide
Catchment hydrology, 257-282
cgs system, 299
Channel bottom roughness. See Roughness: channel bed
Channel flow, 90, 92, 103, 104, 108 , 11 3
Chezy equation, 106
Kannan-Prandtl equation, 108, 109
Manning equation, 106, 1 1 0, 1 1 2
Chczy equation, 106
Climate change, 291-294, 293, 294
Colorado River, 2 1, 140
Combination method, 50
Condensation, 22-23, 252
Cone of depression, 189, 1 91-1 93, 192, 194, 201
Confined aquifer. See Aquifer: confined
Conservation of mass, 9-12, 75-76, 76
flow nets and, 167, 176. See also Continuity equation
Consumptive water use, 2 1 0, 286, 288, 291
Contamination, of water supplies, 5, 146, 163, 198-200, 2 1 1, 281
Continuity equation: for steady flow, 75-76, 76, 93-94
for unsteady channel flow, 125
for vertical groundwater flow, 2 1 9
Continuum assumption, 6 1
Contributing area, 268-270, 269, 270
variable contributing area, 266. See also Specific contributing area
Control volume, 9, 1 1 , 13, 42, 45, 47, 48, 125, 176, 2 1 9
Convective acceleration. See Acceleration: convective
Cooper-Jacob method, 197-198, 199
Coweeta, North Carolina, 259
Critical flow, 94, 97, 98, 99-100, 101
459
Cumulative distribution function, 308
460
Dissolved organic material (DOM), 281 , 282
Divide, 14-1 5, 15
Drainage basin. See Catchment
Drawdown, 189, 1 9 1 , 198, 199
Drought, 243-245, 285, 287, 290, 291
Ecohydrology, 240-256
Effective stress, 1 89- 1 93, 191, 20 I
Elevation head. See Head: elevation
Energy: balance, 41 , 45-49, 47
production, 286--287
water-energy nexus, 286--287
English system, 299
Equipotential, 1 6 1- 1 62, 162, 164, 165, 166, 1 67, 177
Erosion, 272-273, 274
Eutrophication, S
Evaporation, 7-8, 8, 9, 21
increasing air temperature and, 291
water consumption and, 286, 287-288. See also Evapotranspiration
Evapotranspiration, 7-9, 20-2 1 , 40
actual, 41 , 52, 52
in a catchment, 1 5-16, 1 5
climate change, 293-294
energy balance and, 45-49
in global water budget, 9, 10, 1 1-1 3 , 293
green and blue water flows and, 249, 249
latent head flux and, 48
plant water relations and, 242
potential, 41 , 52-53, 52
soil moisture and, 245, 245
unsaturated soil, 234, 235
vegetation and, 248-250
water balance and, 41 , 42
water cycle and, 7-9, 8, 9
water table fluctuation and, 187, 187
water vapor diffusion and, 42-43
Evapotranspiration, methods of estimation: Bowen, 50
combination, 50
461
energy balance, 45--49, 47
mass transfer, 42--45, 43
Penman, 50
water balance, 4 1 --42
Event water. See Hydrograph separation: "old" water
Everglades, Florida, 245-246
Exceedance probability, 30-3 3, 35-36, 1 3 7, 139, 308, 3 1 1 , 311
Extreme events, 34-36, 35, 1 3 6, 140, 291
462
friction factor, 79-80, 80, 83, 84, 105
friction slope, 103, 103, 1 04, 106
Frictionless flow, 69, 72, 76
specific energy, 92-94
Froude number, 100
critical depth, 94, 101 , 101
critical flow, 94, 99, 100, 101, subcritical flow, 94, 97, 98, 99, 100
supercritical flow, 94, 97, 99, 100
463
Heterogeneity, 1 67, 168
effect on groundwater flow systems, 183-1 85, 185
Heterogeneous. See Heterogeneity
High Plains aquifer, 1 74, 206
Hillslope: contributing area, 268, 269
convergent, 268, 270
divergent, 268, 270
runoff, 263-267, 264, 276
stability, 272-273, 274
subsurface flow, 1 63, /63, 177-\78, 272, 276-277
topographic index, 268-271 , 27/
water balance, 278-279, 276
Hollow-stem auger, 1 57, /58
Homogeneity: homogeneous fluid, 6 1 , 73
homogeneous porous media, 1 67, 177, 1 80, 228
Homogeneous. See Homogeneity
Hortonian overland flow. See Overland flow: infiltration-excess
Humidity. See Water vapor: density
Hydraulic conductivity, 149, 151, /53
basin yield and, I 76, 1 79- 180
capillary barriers and, 235, 236, 236
Darcy's law and, 149, 1 5 1- 153, 151
flow nets and, 164-1 67
infiltration and, 226-227
soil moisture and, 2 1 8, 218
soil properties and, 152, 153
transmissivity and, 1 64-165
unsaturated flow and, 2 1 6-2 1 8, 218
Hydraulic geometry, 1 1 3-1 14, JJ3
Hydraulic gradient: Darcy's law and, 150-1 52, 150, 15I
depression cone and, 189, 195
regional groundwater flow and, 1 8 1-1 82
Hydraulic head, 84
Cooper Jacob method and, 1 97-198, 199
Darcy's law and, 150
flow nets and, 1 6 1
groundwater flow and, 148, 149
laminar flow and, 84
464
Poiseuille's law and, 84
water table and, 1 56, 156, 188
Hydraulic radius, 105-106, 105
Hydraulic redistribution, 246-247, 246
hydraulic descent, 247
hydraulic lift, 246
Hydrograph, 1 19-122, 120, 122, 123, 259-260
discharge, 120, 120, 258, 259
rating curve, 120, 122
stage, 120, 120
TOPMODEL simulation of, 279-280, 280. See also Floods
Hydrograph separation, 261-263, 260
baseflow, 12 1 , 259-260, 259, 260, 261-263
chemical, 261-263, 263
"new" water, 261-263, 267
"old" watcr, 261-263, 267, 263
quickflow, 259-260, 260, 261-263
Hydrological cycle, 7-9, 8, 9, 10, 291-294
global hydrological cycle, J 0, 1 2, 13, 292. See also Water cyete
Hydrology, 1
Hydrostatic equation, 63-66, 74, 93
Hyetograph, 25, 25, 26, 122, 259
Hystercsis, 23 1-232, 231, 232
465
Internal energy, 45--46
Intrinsic permeability, 152-1 53, 153, 155, 1 67-168
Irrigation, 1 74, 2 1 0, 249-250, 290-291
period, 255
withdrawals for, 286-288
Isohyet, 27, 28, 29
isohyetal method, 27, 28
Isotropic. See Isotropy
Isotropy, 167
466
in a pipe, 83-84
Mexico City, 174, 1 84, 201-202, 202, 205
Moisture characteristic, 2 1 5, 216, 217, 221 , 221
field measurements, 230-231 , 231
Moisture content. See Soil moisture; Volumetric moisture content
Moisture profile, equilibrium, 221 , 221
Murray·Dariing basin, 253
Muskingum method, 1 34-136, 135
Penman's method, 50
Pcnneability. See Hydraulic conductivity; Intrinsic penneability
Pcnncameter, 152
467
Pescadero Creek, California, 202-203
Piezometer, 1 56, J 56. See also Well
Piezometric surface. See Potentiometric surface
Plants and soil water, 241, 242-246, 243, 246
anisohydric plants, 243, 243
hydrophytes, 245
isohydric plants, 243, 243
phreatophytes, 245
Poiseuille's law, 84, 1 5 1 , 1 55
implications regarding permeability, 1 5 1-1 52, 2 1 6
simi larity with Darcy's law, 148-149, 1 5 1 , 155
Pollution. See Contamination, of water supplies; Radioactive waste
Porosity, 154
Porous medium, 146-149, 147, 148, 1 5 1-1 54
Potenlial evapotranspiration, 41, 52-53, 52, 245
Potentiometric surface, 156, 157, 158, 161-162, 189-193, 190, 192
Power generation, 286-287
Precipitation, 7-9
catchment water budget and, 14-16, 14
climate change, 291-293, 293
direct, 263-265, 264
extreme events, 34-36, 35
hyelograph, 25, 25, 26
intensity, 30
isohyet, 28, 27, 29
occult, 23, 252-254
physics of, 22-24
recycling, 8-9, 23, 252
satellite data, 28
statistics of, 30-39, 31, 33, 34, 37
vegetation effect, 252, 253
water budget and, I 1-1 3
water cycle and, 7-9, 9, 10
weighted average, 27-28
Precipitation gage, 24
Pressure, 62, 63-66, 65, 69, 70
absolute, 64
atmospheric, 193
468
gage, 64, 156
hydrostatic, 63--65, 65, 66
transducer, 157
Pressure gage, 222, 222
Pressure gradient, 70, 79
Pressure head. See Head: pressure
Pressure plate, 2 1 5, 2 I 6
Probability, 307-3 1 2
cumulative distribution function, 308, 3 I I, 3 I 2
density function, 34, 34 308
exceedance, 30-33, 137, 308, 3 1 1, 3 I I
Gaussian distribution, 30
Gumbel distribution, 36
normal distribution, 30, 137, 308-3 12, 309, 310, 31 1, 312
random variable, 30, 307-308
Probability and floods, 136--140, 139
Probability and precipitation, 29-39, 33, 34, 35
Productive consumption usc. See Consumptive water use
Psychrometric constant, 44, 50
Pump test, 197-1 98, 199
469
Regional groundwater flow, 1 80-185, 182, 183
regional flow system, 182-183, 183
Reservoir: concept for catchment water routing, 259, 260
concept in TOPMODEL, 274-275, 275
flood control, 1 19, 126, 127, 1 3 2-133, 132
Folsom Reservoir, 1 1 9, 1 3 3
routing, 126-1 3 3
Totten Reservoir, 127-128. See also Dams
Residence time, 1 2, 13, 258
Return flow, 264, 267
Return period, 3 I , 35-36, 35, 137
Reynolds experiment, 80, 81, 82
Reynolds number, 80, 82, 106, 154
Richards' equation, 2 1 9
Rivers: American River, California, 1 1 8-1 1 9, 1 3 2-133
Brandywine Creek, Pennsylvania, I I I , 1 1 2-1 1 3
Colorado River, 2 1
Pescadero Creek, California, 202-203
Powell River, Virginia, 137
Snake River, Colorado, 279-280, 281-282
Root: hydraulic redistribution, 246-247, 246
water uptake, 241 , 241, 242-244
zone, 2 1 2, 212, 241, 245, 249, 250, 250
Roughness: channel bed, 84-85, 105-107, 107, 108, I I I, 1 1 1
element, 84, 85
height, 45, 46
pipe, 80, 84, 85
surface, 43, 249, 250
Routing, catchment, 273-280, 275
Routing, flood, 1 19, 125-136
Muskingum method, 134-136, 135
reservoir, 126-1 3 3
river, 134-136, 135
Runoff, 7, 8, 1 0
catchment water budget, 1 5-16, 15
global water budget, 1 1- 1 3, 14
vegetation effect, 250-2 5 1 , 251. See also Surface mnoff; Groundwater
mnofT
470
Runoff processes, 263-267, 264
Runoffratio, 1 3, 14
Sahara, 253
San Joaquin Basin, 20 I
Saturated zone, 145, 1 55, 21 2
Saturation deficit, 277-279
Saturation-excess overland flow. See Overland flow: saturation-excess
Saturation value, of moisture conlent, 213
Saturation vapor pressure. See Vapor pressure: saturation
Sensible heat, 46
flux, 47, 48
Shallow subsurface slonnflow, 264, 266
Shear stress, 59, 62
Shenandoah National Park, 262-263, 271
Significant figures, 299-300, 300
Slug test, 194-1 97, 195, 197
Snake River, Colorado. See Rivers: Snake River
Snow, 8, 26
snowfall rates, 26, 28
snowmelt runoff, 8
snowpack, 7, 8
Soil: erosion, 272-273, 274
horizon, 233-234, 233
water potential, 242. See also Hydraulic conductivity; Lntrinsic
penneability; Porosity
Soil moisture, 210, 212, 2 1 9, 223
climate change and, 292, 293-294
equilibrium profile, 221, 221
measurement, 229-230, 229
saturation value of, 2 1 3
water redistribution by plants, 245-248, 246. See also Volumetric
moisture content
Solar energy, 7
Solar radiation, 4 1
Specific contributing area, 269-270, 272-273, 274, 278
Specific discharge, of groundwater, 149-150, 1 5 1-153, 151, 159, 1 64-- 1 65
Specific discharge, in open channels, 94, 96, 107
471
Specific energy, 92-100, 103
alternate depths, 94, 95, 98
diagram, 94, 95, 96-97, 97
Specific heat capacity, 46
Specific yield, 1 88-189, 188
Spreading basing, 173
Sri Lanka, 287
Stage, 119-120, 120, 12/
Standard deviation, 31, 33, 308-309, 309
Steady flow, 68
flow nets, 175-178, 176
of groundwater, 159, 166
in pipes or channels, 71, 73, 75
soil moisture, 219, 223
Steady state, 1 1
Stemflow, 39, 40
Stomata, 40, 45, 241, 243, 243, 294
Storage of water in a confined aquifer, 189-193
Storativity, 189-193, 190
Streamflow 90, 91, 106, 110, 1 /0, 112-114
baseflow, 12 1 , 145, 1 87, 202-205, 259-260, 260
rating curve, 120, 122
roughness coefficient, 105-107, 107, 108, I I I , I I I
stage, 119-120, 120, 12J See also Channel flow
.
Rivers
Streamline, 73, 161, 162, 164-165, 165, 1 76-178
Streamtube, 164-167, 176-180
Stress, 59, 62
effective, 1 89-193, 191, 201
normal, 62
shear, 59, 62
total, 189, 191
Subcritical flow, 94, 97, 99, 100
Sublimation, 47
Subsurface flow, 266, 272-273, 276-279, 276. See also Groundwater flow
Suction, 215
Supercritical flow, 94, 97, 99, 100
Surface discharge. See Surface runoff
472
Surface force. See Force, surface
Surface mnoff, 7, 8, 1 1-1 3 , 14, 1 5-16, 15, 264, 265-266
climate change, 291-293, 294
Hortonian overland flow, 265-266
infiltration-excess overland flow, 265-266
saturation-excess overland flow, 264, 265-266, 268, 270, 272-273, 274,
275
in TOPMODEL, 276, 276, 278-279
vegetation effect, 250-251 , 251
Surface tension, 2 1 3-2 14, 214
System International d'Unites (SI), 298, 299, 299
473
Richards' equation, 2 1 9. See also Soil moisture
Unsteady flow, 1 25, 21 9. See also Steady flow
474
Water management, 2-6
Water quality, 4-6
Water resources, 1 , 2--6, 3, 4, 173-174, 205, 286-291
Watershed. See Catchment
Water stress: and human consumption, 3, 289
and microbial activity, 247-248, 247
and plants, 242, 245, 253
Water table, 156, 163
depth to, 1 86-189, /86, /87
groundwater flow and, 1 77-1 78, / 78, 1 80-183
hydraulic head and, /79
Mexico City, 174
and specific yield, 1 88, /88
Water-table aquifer. See Aquifer: unconfined
Water-table topography, 1 81-1 83, 182, /83, /84
Water vapor, 1 , 7-9, 1 1 , 22
change of phase, 47-48
density, 22, 42, 292
fluxes, 42-43, 43
Water year, 25
Weighing Iysimeter, 234
Weighted average, 27
Weir, 1 00-102, 101, 126-127
Weir equation, 102, 125, 126-127
Well: casing, 157, 158
construction, I 57, 158
hydrograph, 185-193, 186, 187, 193
tests for aquifer properties, 193-198, 195
water-level measurements, 1 57-1 58
Wetlands, 245, 266
Wetting front, 226-228, 226
Wilting point, 242, 243, 243, 247
475
Significant groundwater withdrawals or infrastructural developments like large-scale irrigation can alter natural flow patterns, leading to lower water tables, reduced surface water availability, and land subsidence. Such changes can disrupt ecosystems, reduce water quality by concentrating contaminants, and cause conflict over dwindling water resources. Long-term withdrawals may also lead to irreversible aquifer depletion, necessitating sustainable management practices to balance human needs with environmental protection .
Capillary pressure and soil moisture characteristics significantly influence water infiltration and the movement of wetting fronts in soils. As water infiltrates a soil, capillary pressure determines how water is retained within soil pores, affecting the rate at which water moves downward. Small soil pores maintain water at higher negative pressures compared to larger pores, influencing the moisture content and thus capillary pressure head . Moisture characteristics describe the relationship between volumetric moisture content and capillary pressure; lower capillary pressures correspond to higher moisture contents . The movement of a wetting front is dictated by these interactions as water fills smaller pores first due to capillary forces until reaching saturation . As capillary pressure increases with infiltration, water moves downward faster, altering the wetting front's progression . Macropores, large compared to the surrounding matrix, can quickly channel water past the wetting front once saturated, contributing to rapid infiltration scenarios . Overall, these factors combined determine the dynamics of water movement through soil profiles.
Manning's equation is a practical and widely-used empirical equation to predict water velocity in open channels, relying on parameters like hydraulic radius, slope, and a roughness coefficient . It is straightforward and suitable for situations where detailed data and resources for complex modeling are not available. Using Manning's equation allows engineers to estimate flow velocity with reasonable accuracy, especially when combined with direct measurements for calibration . However, it is less accurate for highly variable or complex flows where significant factors such as turbulence or channel irregularities are involved . More complex models like the Karman-Prandtl equation offer enhanced precision by capturing turbulence effects, but they require detailed input data and computations . Consequently, the choice between using Manning’s equation and more sophisticated models depends on the balance between required accuracy and available resources .
Understanding the dynamics of subsurface aquifers is crucial for water resource management because aquifers store and transmit groundwater, which is a key resource for agriculture, industry, and drinking water. Permeable units like aquifers enhance groundwater flow by acting as conduits, rapidly transmitting water and increasing the natural basin yield. This dynamic affects recharge and discharge patterns and can produce features like springs, especially in heterogeneous basins .
Evapotranspiration and precipitation are critical components of the catchment annual water budget, which is expressed as precipitation (P) minus evapotranspiration (E) and runoff (R), i.e., P = E + R. This budget is crucial for water resource management as it helps determine the availability of water for agricultural, industrial, and domestic use. An understanding of these water fluxes is essential for managing water sustainably, particularly in arid regions .
Fluid dynamics principles like the Bernoulli equation inform our understanding of water flow in natural channels and pipes by expressing the conservation of energy along a streamline. The Bernoulli equation relates velocity head, elevation head, and pressure head, showing that their sum is constant for steady, frictionless, incompressible flow . This principle helps in calculating the velocity and pressure variations along channels and pipes, which are crucial in hydrology for understanding flow dynamics . Formulated under assumptions such as no friction, steady flow, and incompressibility, the equation is often applied by neglecting friction in initial analyses before modifying it to consider real-world factors like friction and changes in cross-sectional area . Its application in scenarios such as water exiting a spigot or flowing through a garden hose illustrates how it predicts changes in velocity and pressure based on changes in height or constriction . The equation's adaptation for channel flow helps hydrologists measure and predict flow behaviors in open channels, crucial for flood management and water resource planning ."}
Specific yield is the measure of the volume of water that drains from an unconfined aquifer per unit area per unit decrease in hydraulic head. It represents the water that is drained by gravity and is typically expressed as a dimensionless ratio, ranging from 0.01 to 0.30. This is because the water is held by forces at the particle surfaces, which is why specific yield is less than the total porosity of the aquifer . In contrast, storativity (or storage coefficient) in confined aquifers describes the volume of water released from storage per unit area per unit declivity in head, due to the expansion of the water and compression of the aquifer material, rather than gravity drainage . Storativity values are typically much lower, often ranging from 0.00005 to 0.005, reflecting the different mechanism of water release in confined aquifers . In essence, specific yield is a measure of water released from an unconfined aquifer by drainage, whereas storativity refers to water released from a confined aquifer through changes in pressure and compression of the aquifer material .
Altering the balance of precipitation and evapotranspiration in a catchment due to land use or climate changes can significantly impact water flow and availability. Deforestation reduces evapotranspiration, leading to increased runoff and potential for flooding, as less water is retained in the soil and more flows directly to streams and rivers . Vegetation plays a crucial role in evapotranspiration and water balance; replacing forests with cropland generally results in decreased evapotranspiration and increased surface runoff . Changes in climate, such as increased temperatures, can enhance evaporation rates, potentially decreasing runoff in areas where precipitation does not increase sufficiently to compensate for increased water loss . Thus, management of water resources and agricultural practices need to carefully consider these hydrological changes to maintain a sustainable water balance ."}
The hydrological cycle functions as a global-scale, endless recirculatory process that links water in the atmosphere, on the continents, and in the oceans. It is conceptualized in terms of reservoirs, such as oceans, atmosphere, glaciers, and groundwater, and involves the volumetric flows of water between these reservoirs. Solar energy drives the formation and transport of water vapor, while gravitational and capillary forces drive precipitation and the flow of water at or below the Earth's surface .
Seasonal changes in precipitation and evapotranspiration critically affect water resource availability in the Colorado River Basin. Precipitation is the primary source of water for the basin, while evapotranspiration represents a significant mode of water loss. In arid regions like the Colorado River Basin, a large portion of water is returned to the atmosphere through evapotranspiration, which is driven by the properties of the land surface and atmospheric conditions . During periods of high precipitation, the availability of water increases as inflows exceed losses through evapotranspiration and runoff, allowing more water to be stored and utilized. However, during drier periods, increased evapotranspiration can significantly reduce water availability as soils lose moisture and inflows decrease, leading to critical water shortages . The historical miscalculation of the Colorado River's water budget illustrates the impact of seasonal extremes on the basin's water resources; water allocations were based on unusually high flow rates during a wet period, which led to overestimation and subsequent shortages as drier conditions prevailed . These dynamics underscore the importance of understanding both precipitation and evapotranspiration to effectively manage and allocate water resources in the basin .