Sei sulla pagina 1di 415

STEVEN A.

MARGULIS

Introduction to
Hydrology

including a MATLAB-based Modular Distributed


Watershed Educational Toolbox (MOD-WET)
2014A EDITION

2014 Steven A. Margulis. All rights reserved.

Cover photo credit: NASA Goddard Space Flight Center Image by Reto Stckli (land surface, shallow water, clouds). Enhancements by Robert Simmon (ocean
color, compositing, 3D globes, animation). Data and technical support: MODIS Land Group; MODIS Science Data Support Team; MODIS Atmosphere Group;
MODIS Ocean Group Additional data: USGS EROS Data Center (topography); USGS Terrestrial Remote Sensing Flagstaff Field Center (Antarctica); Defense Meteorological Satellite Program (city lights).

Throughout the text I have attempted to provide credit to the author/creator/reference from which any graphics
were obtained. Most of the graphics that are not original creations for this text were taken from the publiclyavailable COMET (MetEd) program. Others were taken from the public-domain and/or from credited sources. I
apologize for any omissions and would greatly appreciate being informed of them for correction in future versions.

Preface
This open-access e-textbook is the first iteration in an
experiment designed to provide a novel tool for learning the
basic concepts in hydrologic science. I hope it is a tool that
will be useful to the hydrologic community at large. It is
expected that the book and accompanying codes will be a
dynamic entity that will continue to evolve going forward. The
book is optimized for use as an electronic book on an iPad (or
other iBooks platform) to take full advantage of multimedia,
search capability, web links, etc. However, for those without
access to those platforms we also oer a PDF version (which
will not have the full functionality of the iBooks version).
Note: To be able to play the embedded multimedia
links in the PDF you will need to view the document
using the freely available Adobe Reader program
(www.adobe.com).
The material in the book is primarily derived from notes and
other material used in the undergraduate Civil &
Environmental Engineering (CEE) 150: Introduction to
Hydrology course oered in the Department of Civil and
Environmental Engineering at UCLA. CEE 150 is a 10-week
course designed to provide a survey of the hydrologic cycle.
The format of the course and topical coverage owes much of
its genesis to a similar course taught at MIT by Dara
Entekhabi. I have taught the course at UCLA for over 10
years and it has undergone a significant evolution over those
years.

The structure of the course is based on 4 lecture-hours per


week with a 2-hour discussion section for review and example
problems. This e-textbook is an attempt at organizing all of
the material in one format for students in the class (as well as
any others who may find the topic and material useful). It
would be relatively easily adaptable to either quarter-based or
semester-based courses. Some material beyond that which is
covered in the course is provided in the book for completeness.
It could also be useful as a reference text for other
undergraduate or graduate hydrology courses.
The book is organized into ten main sections: 1. The
Hydrologic Cycle, 2. Atmospheric Composition and
Thermodynamics, 3. Radiation Processes, 4. Large-scale
Atmospheric Circulation, 5. Precipitation Processes, 6. Snow
Processes, 7. Unsaturated Flow and Infiltration, 8.
Evaporation Processes, 9. Groundwater Flow and Well
Hydraulics, 10. Runo and Streamflow. The chapters in the
book are organized based on this topic list. A final chapter is
used to develop and present a distributed watershed model
developed from concepts covered throughout the previous
chapters of the book. The model is meant to show how
modular hydrologic concepts can be built-up to form a fully
functional watershed model. Each chapter comes with a set of
Learning Objectives that explicitly lay out the key things you
should know by the end of that chapter. Sample conceptual
questions and sample problems are also provided to assist in
the learning of the material.
It is my strong belief that the most important aspect of the
course is building students knowledge base through realistic
ii

problem solving. Weekly homework assignments provide a


significant amount of hands-on problem solving, with a
particular emphasis on realistic problem solving using
numerical modeling and applications. For numerical problems
I use MATLAB (http://www.mathworks.com/products/matlab)
as a framework since it is taught to all CEE undergraduates
earlier in their curriculum and is a relatively accessible and
user-friendly numerical tool. If not freely available in a
student computer lab, the student version of MATLAB provides
a useful alternative.
A novel aspect of the book is a companion set of MATLAB
functions. The functions are presented with each chapter as
modular units for doing associated numerical calculations and
modeling. The functions are ultimately pieced together to
form a simple distributed watershed model for educational
usage. Together, the set of functions and model are described
as the Modular Distributed Watershed Educational Toolbox
(MOD-WET). Apart from specific applications using
individual codes, MOD-WET aims to illustrate for students
how modular codes are useful because they can be reused in
many applications (not just in this course, but other related
courses) and built-up to form more complicated analysis and
modeling frameworks. The code along with the book are
available for download at: http://aqua.seas.ucla.edu/
margulis_intro_to_hydro_textbook.html.
The book benefits greatly from figures and animated videos,
many of which were created or inspired by others. In
particular the textbooks of Bras (1990), Mays (2005),
Marshall and Plumb (2007), and Dingman (2008) are

referenced extensively throughout the text. Additionally, the


resources posted at the MetEd site (www.meted.ucar.edu)
provide a fantastic set of modules for learning many of the
concepts covered here (as well as many other related to
meteorology and associated fields). The MetEd site is
maintained by the COMET program (www.comet.ucar.edu)
with the primary goal of assisting education through advanced
learning materials. Many of the figures and movies used in the
text are those provided by MetEd. I am greatly appreciative
of this resource and would strongly encourage students to
explore the MetEd learning modules in more detail on their
own.
I greatly acknowledge the teaching assistants over the years
(in particular Michael Durand, Bart Forman, Keith
Musselman, Manuela Girotto, Laurie Huning, Mahdi Navari,
Liz Baldo, and Gonzalo Cortes) who have significantly
contributed to the development of the material into its current
form. In particular, Laurie and Manuela have greatly helped
in the development of the book and the integrated watershed
model. Additionally Ben Wong was very helpful in updating
and standardizing many of the MOD-WET codes.
Any errors (typographical or otherwise) or omissions in the
text are attributable to me. Bringing them to my attention
would be greatly appreciated and allow for continued
improvement of the text. I would be happy to add other
relevant problems to the book, watershed simulations, and/or
functions to MOD-WET based on user contributions.
Steve Margulis (margulis@seas.ucla.edu); July 2014
iii

Table of Contents
Preface ..........................................................i
Chapter 1: The Hydrologic Cycle .................7
Section 1: Learning Objectives ..........................................8
Section 2: Motivation ........................................................9
Section 3: The Hydrologic Cycle ......................................12
Section 4: Unique Properties of Water .............................14
Section 5: Mass balance, fluxes, and units .......................18
Section 6: Global Hydrologic Cycle and Average

Mass Balance.........................................................22
Section 7: Watershed Mass Balance .................................26
Section 8: MOD-WET Codes ...........................................32
Section 9: Conceptual Questions ......................................33
Section 10: Sample Problems ...........................................34

Chapter 2: Atmospheric Thermodynamics ..38


Section 1: Learning Objectives ....................................... 39
Section 2: Atmospheric Composition .............................. 40
Section 3: Atmospheric States .........................................42
Section 4: Metrics for Water Vapor Concentration

in Air.....................................................................45
Section 5: Vertical Profiles of Atmospheric States ...........50
Section 6: MOD-WET Codes ...........................................55
Section 7: Conceptual Questions ......................................56
Section 8: Sample Problems .............................................57

Chapter 3: Radiation Processes ..................60


Section 1: Learning Objectives ........................................61
Section 2: Basics of Radiation .........................................62

Section 3: Shortwave vs. Longwave Radiation ..................65


Section 4: Radiative Properties of Media .........................67
Section 5: Modeling Top-of-Atmosphere Shortwave

Fluxes ...................................................................70
Section 6: Modeling Surface Shortwave Fluxes ................78
Section 7: Modeling Longwave Fluxes at the Surface .......89
Section 8: Net Radiation at the Surface ...........................93
Section 9: MOD-WET Codes ...........................................95
Section 10: Conceptual Questions ....................................97
Section 11: Sample Problems ...........................................98

Chapter 4: Atmospheric Circulation ..........103


Section 1: Learning Objectives ......................................104
Section 2: Global Distribution of TOA Net Radiation ...105
Section 3: Atmospheric Motions Driven by

Latitudinal Energy Imbalance .............................108
Section 4: Summary of Key Characteristics of

Circulation ..........................................................114
Section 5: Fundamental Equations of Atmospheric

Motions ...............................................................120
Section 6: Conceptual Questions ....................................126
Section 7: Sample Problems ...........................................127

Chapter 5: Precipitation Processes ............129


Section
Section
Section
Section
Section
Section
Section
Section

1:
2:
3:
4:
5:
6:
7:
8:

Learning Objectives ......................................130


Thermodynamics of Cloud Formation ............131
Cloud Microphysics .......................................136
Meteorology of Precipitation .........................143
Precipitation Climatology and Extremes .......153
Precipitation Measurement ............................157
MOD-WET Codes .........................................169
Conceptual Questions ....................................170
iv

Section 9: Sample Problems ...........................................171

Chapter 6: Snow Processes ........................176


Section 1: Learning Objectives ......................................177
Section 2: Surface Energy and Mass Balance .................178
Section 3: Snowpack Characteristics ..............................181
Section 4: Snowpack Accumulation and

Metamorphism ....................................................186
Section 5: Snowmelt ......................................................194
Section 6: Impact of Vegetation .....................................198
Section 7: Snow Climatology .........................................201
Section 8: Snow Measurement ........................................204
Section 9: MOD-WET Codes .........................................210
Section 10: Conceptual Questions ..................................211
Section 11: Sample Problems .........................................212

Chapter 7: Unsaturated Flow and



Infiltration.........................................214
Section 1: Learning Objectives ......................................215
Section 2: Unsaturated Zone Characteristics ..................216
Section 3: Flow in Unsaturated Porous Media ...............222
Section 4: Modeling Unsaturated Zone Flow

Dynamics ............................................................229
Section 5: Infiltration ....................................................232
Section 6: Infiltration Capacity Models ..........................236
Section 7: Modeling Actual Infiltration with the Time
compression Approximation .................................239
Section 8: MOD-WET Codes .........................................245
Section 9: Conceptual Questions ....................................246
Section 10: Sample Problems .........................................247

Chapter 8: Evaporation .............................250

Section
Section
Section
Section
Section
Section
Section
Section

1:
2:
3:
4:
5:
6:
7:
8:

Learning Objectives ......................................251


Basics of Evapotranspiration .........................252
Mass-transfer Model for Evaporation ............254
Transpiration ................................................262
Additional ET Models ...................................269
MOD-WET Codes .........................................275
Conceptual Questions ....................................276
Sample Problems ...........................................278

Chapter 9: Groundwater Flow ...................281


Section 1: Learning Objectives ......................................282
Section 2: Basic Groundwater Characteristics ................283
Section 3: Development of Groundwater Flow

Equation .............................................................288
Section 4: Groundwater Flow to Pumping

Wells ...................................................................294
Section 5: Superposition of Groundwater

Solutions .............................................................301
Section 6: Conceptual Questions ....................................311
Section 7: Sample Problems ...........................................312

Chapter 10: Runoff and Streamflow ..........315


Section 1: Learning Objectives ......................................316
Section 2: Basic Runoff and Streamflow Definitions .......317
Section 3: Runoff Generation Mechanisms .....................320
Section 4: Streamflow Hydrographs ...............................327
Section 5: Rainfall-Runoff Modeling:

Unit Hydrograph .................................................331
Section 6: Rainfall-Runoff Modeling:

Physically-based Models ......................................345
Section 7: Streamflow Routing: Unsteady Flow ..............351
Section 8: Streamflow Routing: Hydrologic Routing .......359
v

Section 9: Measurement of Streamflow ...........................361


Section 10: Conceptual Questions ..................................364
Section 11: Sample Problems .........................................365

Chapter 11: A Simple Watershed


Model ................................................368

Section 1: Learning Objectives ......................................369


Section 2: Development of a Distributed Watershed Model

Using MOD-WET ...............................................370
Section 3: Numerical Implementation of the MOD-WET

Model ..................................................................378
Section 4: Example MOD-WET Model Applications ......384
Section 5: Sensitivity of Hydrograph Response to

Watershed Characteristics ...................................398
Section 6: MOD-WET Codes .........................................402
Section 7: Conceptual Questions ....................................403
Section 8: Sample Problems ...........................................404

Chapter 12: MATLAB Basics ......................406


Chapter 13: References .................................410

vi

Chapter 1

The
Hydrologic
Cycle

S ECTION 1

Learning Objectives

10. Construct a watershed from topographic data for use in


hydrologic analysis

By the time you finish this chapter you should be able to:
1. Provide a basic definition of hydrology
2. List some of the key motivations for studying hydrology
3. Describe the primary reservoirs in the global hydrologic
cycle and their relative sizes
4. Describe the primary water fluxes connecting the
reservoirs in the global hydrologic cycle
5. Describe and dierentiate the unique properties of water
relevant to hydrology
6. Convert between mass, volume, and energy fluxes and flux
densities
7. Write down and apply the mass balance equation for a
particular control volume
8. Identify the average rate of the global hydrologic cycle and
the corresponding average residence times of water in
dierent global hydrological reservoirs
9. Define a watershed and explain its relevance to hydrology

S ECTION 2

Motivation

Before beginning our discussion of hydrology we should
define what it is and why its study is important. Simply put,
hydrology is the study of water. In particular, hydrologic
science is a branch of geoscience that aims to diagnose and
predict: 1) the spatial and temporal characteristics of water in
its various storage reservoirs (terrestrial, atmospheric,
oceanic) and 2) the corresponding fluxes of water between
these reservoirs.

There are many motivations for studying hydrology. The
fundamental motivation is that water is required for life,
making it of paramount societal importance. While the overall
amount of water is fixed, population is increasing and
therefore having an understanding of water and its storage
and movement is key to our survival and will likely dictate
how and where we will live in the future.

There is also a significant scientific motivation for
studying water because it plays a key role in the Earth system
as a whole, including weather and climate processes,
landscape evolution, and biogeochemical processes. For
example, in terms of climate, we would like to know how and
why water varies from one location to another. A relevant
example in terms of annual average precipitation distribution
over California is shown in Figure 1.1. The map shows clear
gradients in precipitation from North to South as well as a

function of elevation (i.e. high precipitation occurs in the


Sierra Nevada). Understanding how and why this variability
exists will be a key component of this course.

Significant connections also exist between water and
engineering due to the need to design water resources systems
(both for water supply and hazard mitigation) to manage and
allocate water resources. A map of population distribution in
California is shown in Figure 1.2. It is clear from the map that
there is a distinct mismatch between the areas of highest
precipitation and highest population. For example, Los
Angeles receives approximately 15 inches of precipitation per
year (Figure 1.1), which amounts to approximately 90
gallons/day/person. The average user in Los Angeles however
requires approximately 200 gallons/day, an amount that
cannot be met by local supply. Beyond urban water use, a
significant amount of water is used for agricultural irrigation,
much of which occurs in some of the areas of lowest
precipitation in the state (i.e. the Central Valley and the
Imperial Valley in the southeastern corner of the state). Such
mismatches in distribution of water availability and use is not
uncommon throughout the world and necessitates a significant
amount of water resources engineering to move water from
where it is plentiful to where it is needed.

Finally, the need for water and its scarcity implies
economic value, hence making water of high economic interest.
In fact, much of the development of the Western U.S. has
been and will be dictated by water (e.g. Reisner, 1993,
Hundley, 2001, Carle, 2004, Green, 2007). Questions like who
has the right to water and whether it should be treated like
9

F IGURE 1.1 Map of annual average precipitation over California (from www.ocs.orst.edu/prism).

other commodities are beyond the scope of this course, but


hugely important.

The focus on this course will be on the basic physical
understanding of processes related to hydrology as a basis for
ultimately engaging the issues and problems described above.

F IGURE 1.2 Population in California based on 2000 U.S. Census survey.

10

In terms of quantification of processes, many equations will be


used throughout the book. The majority of the equations used
will be physically based rather than empirical. Physically
based equations are derived from fundamental physical
equations and are therefore dimensionally consistent (and
most commonly use SI units). Empirically based equations are
not derived from physics, but instead fitted to experimental
data. As such, they may use and require that inputs be in
particular units and provide answers in prescribed units.
When empirical equations are presented throughout the book,
the necessary units for inputs and outputs are described
explicitly. For physically based equations, units are typically
not explicitly mentioned because they can consist of any
dimensionally consistent set.

11

S ECTION 3

The Hydrologic Cycle



A useful way to conceptualize hydrology is via the notion
of the hydrologic cycle as shown in Figure 1.3 and Movie 1.1.
One can envision this cycle occurring across spatial scales
from local to global and across all time scales. Some of the
key hydrologic reservoirs include: the atmosphere, soil water
(i.e. near-surface in the unsaturated soil zone and in deeper

M OVIE 1.1 Conceptual movie of the hydrologic cycle.

F IGURE 1.3 Conceptual picture of the hydrologic cycle.

groundwater reservoirs), surface water (i.e. rivers, lakes), the


cryosphere (including seasonal snowpacks, glaciers, icecaps),
the biosphere (e.g. water stored in vegetation), and the
oceans. Across all reservoirs, water can be stored in any of its
three phases (liquid water, solid ice, and water vapor). For
example, in the atmosphere, water can exist in all three
phases, while the cryosphere consists mostly of solid ice and
soil and surface water is primarily liquid. Table 1.1 shows
estimates of the volume of water stored in the key reservoirs
in the global system. Note that these are very rough estimates
and dierent cited sources will yield dierent estimates. The
numbers are meant to illustrate the order of magnitude of
each reservoir. The ocean is by far the largest of the reservoirs
and contains saline water. The next biggest reservoir is the ice
caps and glaciers which contain mostly bound-up (albeit
increasingly changing) fresh water in frozen form. For the
12

T ABLE 1.1. E STIMATES OF STORAGE IN PRIMARY


GLOBAL HYDROLOGIC RESERVOIRS (B RAS , 1990).
RESERVOIR
Oceans

VOLUME
(km3)

% TOTAL WATER

1,322,000,000

97.2

Ice caps & glaciers

29,199,700

2.1

Groundwater (near-surface)

4,171,400

0.31

Lakes & Rivers

130,700

0.017

Soil Moisture

66,700

0.005

Atmosphere

12,900

0.0009

dynamic (mobile) freshwater reservoirs, groundwater is by far


the largest. The relatively small fraction of water stored in
lakes and rivers, soil moisture, the atmosphere, and biosphere
misrepresents their importance as they tend to be the most
dynamic parts of the hydrologic cycle as will be described in
more detail below.

Water moves between these various reservoirs via fluxes.
Some of the key hydrologic fluxes include: precipitation (either
in liquid or solid form), evaporation and transpiration
(together referred to as evapotranspiration), infiltration,
recharge, and runo. Precipitation and evapotranspiration are
the key fluxes between the atmosphere and surface (land and
oceans). Precipitation may accumulate when it falls as snow
while rainfall is partitioned at the surface into infiltration and
surface runo. Percolation of water through the unsaturated
soil zone recharges groundwater aquifers which ultimately

feeds surface water bodies via lateral flow and runo. The
atmospheric water is replenished via evaporation from the soil
and open water surfaces and transpiration from vegetation.
Fluxes and storage are directly linked via mass balance as
described in more detail below.

A key aspect of the hydrologic cycle is the fact that it is
driven by energy inputs (primarily from the sun; Figure 1.3).
At the global scale, the system is essentially closed with
respect to water; negligible water is entering or leaving the
system. In other words, there is no external forcing in terms of
a water flux. Systems with no external forcing will generally
eventually come to an equilibrium state. So what makes the
hydrologic cycle so dynamic? The solar radiative energy input,
which is external to the system, drives the hydrologic cycle.
Averaged over the globe, 342 W m-2 of solar radiative energy
is being continuously input to the system at the top of the
atmosphere. This energy input must be dissipated, and this is
done, to a large extent, via the hydrologic cycle. Due to this
fact, the study of hydrology is not isolated to the study of
water storage and movement, but also must often include
study of energy storage and movements.

13

S ECTION 4

Unique Properties of Water



Many of the reasons for the importance of water in the
Earth system (and for its necessity in life) have to do directly
with its unique properties. Some of the key properties relevant
to hydrology are discussed in more detail below.

The key properties of water are ultimately a function of
its molecular structure, which consists of two hydrogen atoms
and one oxygen atom (i.e. H2O). In particular, a H2O molecule
is polar in nature where the hydrogen side of the molecule
has a positive charge and the oxygen side has a negative
charge (Figure 1.4). The polarity of the molecule controls how
water molecules arrange themselves in liquid and ice form,
where the positive end of one molecule is attracted to the
negative end of another molecule via so-called hydrogen
bonds. The arrangement of molecules determines the density
of water which for liquid water is approximately 1000 kg m-3
(with a slight variation with temperature as shown in Table
1.2) and for ice is 917 kg m-3.

The attraction from these bonds leads water to have a
high viscosity and surface tension among other unique
properties. Figure 1.5 shows the thermodynamic phase
diagram for pure water, which is simply a plot showing how
regions of pressure-temperature space correspond to the
particular phase of water. The curves represent equilibrium
conditions where two water phases can coexist and the

F IGURE 1.4 Basic structure of a water molecule showing its


polarity (from bioweb.uwlax.edu/bio203/2010/olson_moll/
polarwatermoleculeGOOD.jpg).

intersection of all three curves is the triple point. The


equilibrium curve between solid and liquid represents melting
(or fusion), the curve between liquid and vapor represents
vaporization (or condensation) and the curve between solid
and vapor represents sublimation (or deposition). Of
particular relevance is that the pressure and temperature
conditions for Earth on the phase diagram are located in a
region that is near boundaries between all three phases, and
in fact phase changes on Earth are a common occurrence. For
context, conditions on most other planets in the solar system
have pressure and temperature conditions that are not
14

T ABLE 1.2. P ROPERTIES OF P URE L IQUID W ATER AS


A FUNCTION OF T EMPERATURE A BOVE F REEZING

F IGURE 1.5 Phase diagram showing phases of pure water as a


function of pressure and temperature (from

msduncanchem.com/Unit_11/phase_diagrams_ws_files/image001.gif).

suitable for such phase transformations (if water was even


present). As described further below, phase changes for water
are associated with considerable energy consumption/release.
This is one of the primary reasons why water and the
hydrologic cycle is of primary importance in the climate
system.

Of particular importance with respect to climate on
Earth are the latent heat and thermal heat capacity
properties of water. The specific heat capacity of a substance
is the amount of energy required to raise a given mass of the
substance by 1 degree. For water, the specific heat capacity
(cp) at T=0C is 4216 J kg-1 K-1. Like many properties of

SPECIFIC HEAT

TEMPERATURE

DENSITY

(C)

(kg m-3)

999.87

4216

15

999.13

4184

30

995.67

4177

CAPACITY
(J kg-1 K-1)

water, this can vary with temperature as shown in Table 1.2.


This is the highest specific heat capacity of any known
substance (Mays, 2005). So why is this property so important
to climate on Earth? The primary eect is one of buering
temperature changes. Much of the solar energy entering the
top of the Earths atmosphere reaches the surface, a
significant fraction of which is covered by water. The high
heat capacity of water means that much of this energy can be
absorbed with relatively small temperature changes. A planet
without water would warm much more significantly for the
same energy input. This eect can also be seen in regional
climate patterns. Coastal areas near large bodies of water
generally have a more temperate climate (i.e. relatively small
temperature variations) compared to those inland due to the
buering capacity of the water body. In another example, the
human body takes advantage of the high heat capacity of

15

water to regulate the bodys temperature within the small


range needed for healthy functioning.

Another key set of properties of water are the so-called
latent heats associated with phase transformation. While heat
capacity is the energy required to change the temperature of a
given substance (i.e. liquid water), the latent heats are the
energy associated with a constant temperature (isothermal)
phase change. The latent heat of vaporization (Lv) is the
energy consumed in transforming liquid water to water vapor
(or released in converting vapor to liquid) and has a value (at
T = 0C) of Lv = 2.5106 J kg-1. This property has a slight
variation with temperature (e.g. equals 2.25 x 106 J kg-1 at
100C), but for most applications we can use this nominal
value. The latent heat of fusion (Lf) is the energy consumed in
transforming solid water (ice) to liquid water (or released in
converting liquid to ice) and has a nominal value of Lf = 3.34
105 J kg-1. Note that this is almost an order of magnitude
less than the latent heat of vaporization. The latent heat of
sublimation (Ls) is the energy consumed in transforming ice to
water vapor (or released in converting vapor to solid) and has
a nominal value of Ls = 2.85106 J kg-1. The large amounts of
energy required to break the hydrogen bonds between water
molecules make these properties very high compared to other
substances. The fact that these latent heats are so large for
water, coupled with the fact that phase transformations are
quite common in the Earth system, result in significant energy
sources and sinks in the hydrologic cycle. Specifically, latent
heating is critical in energy transfer between the surface and
atmosphere and in global heat transport.

E XAMPLE 1.4.1
How much energy would be released if all of the
atmospheric water vapor were condensed?
According to Table 1.1, the volume of water in the
atmosphere is approximately equal to 12,900 km3. The
equivalent amount of mass (using a density of water of
1000 kg/m3) is then given by:

(1000)3 m 3
Mass = (12,900 km )(1000 kg m )
1 km 3
= 1.29 1016 kg
3

-3

The equivalent amount of released energy is obtained by


multiplying the mass by the latent heat of vaporization
(Lv). Hence the amount of energy is:

Energy = (1.29 1016 kg)(2.5 10 6 J kg -1 )


= 3.225 10 22 J
This is a large amount of energy. Despite the atmosphere
only containing relatively little water compared to
other reservoirs, this amounts to enough energy to power
over 1 trillion 100 W light bulbs for over 10 years! The
key point is that the water in the atmosphere is a
significant potential latent heat source (or sink) when
phase changes occur. This energy, which is due to the
high latent heat of water, is one of the primary
mechanisms for energy transport in the atmosphere.
16

E XAMPLE 1.4.2
If the energy released in the previous example
were absorbed by the global oceans, how much of
a temperature change would there be?
The specific heat capacity of a substance dictates how its
temperature will change (per unit mass) given an energy
input. According to Table 1.1, the volume of water in the
global oceans is approximately equal to 1,322,000,000
km3. The equivalent amount of mass (using a nominal
density of water of 1000 kg/m3) is then given by:

(1000)3 m 3
Mass = (1,322, 000, 000 km )(1000 kg m )
1 km 3
= 1.322 10 21 kg
3

-3

The temperature change that would result from the


energy input would then be:

Temperature change =
(3.225 10 22 J)/(1.322 10 23 kg)/(4216 J kg -1K-1 )
= 5.8 10 5 K
Due to the large heat capacity of the ocean (and its large
mass) there would be an imperceptible change in
temperature due to this large energy input. This is
illustrative of the large buering capacity of water.

17

S ECTION 5

Mass Balance, Fluxes, and


Units

The key starting point of most hydrologic analyses
involve application of mass balance. A mass balance equation
for any system (as well as many other conservation laws) can
be derived using Reynolds transport theorem (Mays, 2005).
The theorem starts with the definition of a control volume in
space defined by a control surface (Figure 1.6). Implicit in any
mass balance application is a pre-specified control volume;
without one the mass balance is not meaningful.


Properties of the fluid are denoted as either extensive or
intensive. Extensive properties are those that are in some way
related to the total mass of the system, including mass (m),
momentum (mV), and energy (E). Intensive properties are
generally normalized by mass (e.g., momentum per unit mass,
which is simply velocity, energy per unit mass, etc.). In the
control volume approach, we can represent an extensive
property as B and an intensive property as B (dB/dm). The
general control volume equation can be expressed as (Mays,
2005):

dB d
=
B dV + CS BV dA
dt
dt CV

(1.5.1)

where the individual terms are:

dB
total rate of change of extensive property
dt
d
of change of extensive property in control
B dV rate
dt CV
volume (where is density of fluid and dV is a
differential volume)
rate of outflow of extensive property across
BV dA netcontrol
surface (where V is velocity and dA

CS

F IGURE 1.6 Basic schematic of a control volume with fluxes


across the control surface (adapted from Mays, 2005).

is differential area of the control surface)


The mass balance (or continuity) equation is derived
from the Reynolds transport theorem by setting B to mass
(making B = 1) and noting that for mass conservation the
left-hand-side term equals zero, which yields:

d
dM

dV
=
= V dA

dt CV
dt
CS

(1.5.2)
18

which simply states that the rate of change of mass storage in


the control volume (mass/time) is exactly balanced by the net
rate of mass outflow across the control surface. Several
simplifying assumptions can be made to Equation (1.5.2),
including constant density (often appropriate in the case of
liquid water flow/storage problems) in which case:

d
dV = CS V dA
dt CV

(1.5.3)

kg 1
m3

=
;
s
w
s

which can then simply be written as:

dS
= I i Oi
dt
i
i

(1.5.4)

where S is the volume storage and the right-hand-side is the


sum of volumetric flux inflows (Ii) minus the sum of
volumetric flux outflows (Oi) across the control surface of the
control volume. A key point is that only fluxes across the
control surface need be considered in the mass balance
equation. Internal fluxes do not contribute to mass/volume
storage changes. This fact is often used in the construction of
a control volume to eliminate fluxes that are unknown by
making them internal fluxes. Hence for the equation to be
meaningful, a control volume must be explicitly defined.
Another simplified case is that of steady-state in which case
the storage term is equal to zero, yielding:

I = O
i


In the context of mass-balance (or other) applications,
several forms of fluxes are used and it is important to be able
to dierentiate between them and the units associated with
them. Mass fluxes (i.e. used in Equation (1.5.2)) have units of
mass/time (e.g., kg s-1). Volume fluxes (i.e. used in Equation
(1.5.4)) have units of volume/time (e.g., m3 s-1). To convert
between the two, the density of the fluid can be used, i.e. for
water:

(1.5.5)

w = 1000 kg m 3

Often a flux density is used in place of a flux, which is simply


either a mass or volumetric flux normalized by an appropriate
cross-sectional area across which the flux is occurring, i.e.

kg 1
kg
2
s A
ms
m3 1
m
mm m
Volume:
(or
,
, etc.)
s
A
s
d
y
Mass:

where as shown above, if the mass flux density is known,


volume flux density can be obtained by dividing by the
density of water. The volume flux density is a common way of
expressing fluxes in hydrology. For example, the annual
average rainfall in Los Angeles is 15 inches/year, which is
implicit over a given area. To get the actual volume flux
would require multiplying this flux density by the surface area
of Los Angeles over which the flux density occurs.

19


As will be discussed in much more detail later, and has
been alluded to above, water and energy fluxes are often
directly connected. The most relevant example is that the
evaporation flux is tied directly to a phase change energy flux.
For example, the latent heat flux associated with phase
change is simply the mass flux density multiplied by the latent
heat of vaporization:

Energy flux density = Lv Mass flux density

J kg
= Wm 2
2
kg m s

where the units of W m-2 are the most commonly used for
energy fluxes (actually flux densities) in hydrology.

E XAMPLE 1.5.1 ( CONTINUED )


be lined with concrete to prevent any drainage or
seepage from the bottom. Using a water balance,
what volume (on average) of groundwater (in
cubic meters) must be added to the lake each
year to keep the pond at a steady-state level?
The first step to any mass balance is the definition of a
control volume. For this problem the control volume and
relevant fluxes are shown below:

E XAMPLE 1.5.1
A golf course has requested a permit to install a
40,000 ft2 pond to enhance the beauty of its
facilities. It is hypothesized that due to high
evaporation rates, the water in the pond will
have to be supplemented with pumped
groundwater. There is a small creek that
discharges an average of 0.0005 m3/s into the
pond. The outlet valve from the pond releases an
average rate of 0.0003 m3/s to keep the pond
from getting stagnant. Precipitation on the pond
is 260 mm/year, and the annual evaporation is
estimated to be 105 inches/year. The pond will

where the mass balance for the pond can be written as:

dS pond
dt

= P +Qin +Gin E Qout

The steady-state form can be invoked in which the


storage change would be zero so that:

20

E XAMPLE 1.5.1 ( CONTINUED )

Gin = E +Qout P Qin


Next, the fluxes must be converted to proper and
consistent units. Converting everything to SI units, the
area of the pond is given by:
2

Apond

! 1m $
2
2
= 40000 ft #
& = 3718 m
" 3.28 ft %

The known fluxes are given by:

mm 1 m
3718 m 2 = 967 m 3 /y
y 1000 mm
in 2.54 cm 1 m
E = 105
3718 m 2 = 9916 m 3 /y
y 1 in 100 cm
P = 260

Qout

m3
Qin = (0.0003 0.0005)

s
(3600 24 365) s
= 6307 m 3 /y
1y

The necessary influx from the groundwater pumping is


then given by:

Gin = (9916 967 6307) m 3 /y = 2642 m 3 /y

21

S ECTION 6

Global Hydrologic Cycle


and Average Mass Balance

To begin to understand timescales of hydrologic fluxes
and mass balance it is instructive to start at the global scale.
If we consider a control volume covering the entire surface of
the globe (Figure 1.7), the relevant fluxes across the control
surface are the precipitation and evaporation fluxes over both
land and ocean. Note that the runo between land and ocean
is an internal flux and therefore not relevant to the mass
balance of the control volume we have chosen. The
instantaneous global mass (or volume) balance can then be
written as:

dS
= Pland Eland + Pocean Eocean
dt

(1.6.1)

where S simply represents the mass/volume stored in the


ocean and land water reservoirs and consistent units must be
used (either in terms of mass or volume). Given the above
equation, if all instantaneous fluxes are known, then the
storage change can be determined. Alternatively, if the storage
change is known and all fluxes but one are known, the
equation can be used to solve for the unknown flux as a
residual.

In practice, the instantaneous mass balance terms are

F IGURE 1.7 Estimates of global long-term average fluxes and


reservoir storage over land and ocean. Storages are indicated
by boxes (in units of 103 km3) and fluxes are indicated by arrows (in units of 103 km3/year) (adapted from U.S. National Research
Council, 1991; Mays, 2005).

dicult, if not impossible, to estimate accurately (especially


over large scales). Therefore to simplify hydrologic analysis we
often take the long-term average of the above mass balance
equation, where we can use the time-averaging operator:

X =

1
T

t +T

X dt

(1.6.2)

where T is an averaging window in time and the above


equation can be applied to each term in the mass balance
22

equation:

dS
= Pland Eland + Pocean Eocean
dt

(1.6.3)

The question then becomes, if the averaging period is long


enough, can we say anything about any of these terms? Over
the long-term we can reasonably say that the (average)
storage change should be close to zero or at least much
smaller than the average fluxes. This implicitly assumes some
stationarity in the system (i.e. it is steady-state in the longterm). In general, the longer the averaging period the more
accurate the assumption will be. At a minimum it should be
at least a full year to average over the seasonal cycle (i.e. most
regions have a wet season where water storage is increasing
and dry season where water stores are depleted). Anything
less would be expected to violate the steady-state assumption.
Taking the long-term average results in:

dS
0 = Pland Eland + Pocean Eocean = Pglobal E global
dt

(1.6.4)

Pglobal = E global

(1.6.5)

which should make sense because if these fluxes were not in


balance, water would either be accumulating over the longterm in the atmosphere (if evaporation exceeded
precipitation) or in the land/ocean (if precipitation exceeded
evaporation). From data it is estimated that:

Pglobal 1 m yr -1 = E global

where this is expressed in terms of a volumetric flux density


(i.e. normalized over the surface of the Earth). Expressed as a
volumetric flux this would equal 1415 km3 day-1.

We can think of the above flux as the average global rate
of the hydrologic cycle. In other words this is the average
amount of water moving through the various hydrologic
reservoirs. Note that the instantaneous fluxes at any given
point and time may be very dierent than this (as will be
discussed later). Given this rate we can estimate average
residence time in the various reservoirs using the relationship:

Residence Time =

Volume of reservoir
Average volumetric flux rate

(1.6.6)

which provides an estimate of the turnover time of water in a


given reservoir. Given the estimates of reservoir volumes
(shown in Table 1.3), the residence times can be estimated
and are shown in Table 1.3. These residence times are
indicative of how dynamic the dierent reservoirs are. For
example, a molecule of water will stay in the atmosphere for 9
days on average before precipitating out, while a molecule of
water will spend 8 years on average flowing through
groundwater. So while the atmosphere and soil moisture
reservoirs are among the smallest, they are the most dynamic
based on their residence times and contribute to much of the
variability in the system.

It is very important to keep in mind that the above
numbers are global long-term averages. Significant variability
exists in most hydrologic fluxes in both space and time. In
23

T ABLE 1.3. E STIMATES OF RESIDENCE TIME IN


PRIMARY GLOBAL HYDROLOGIC RESERVOIRS . T HESE
VALUES ASSUME A GLOBAL AVERAGE HYDROLOGIC
RATE OF 1 METER / YEAR (1415 km 3 /day)
RESERVOIR
Oceans

VOLUME
(km3)

RESIDENCE
TIME

1,322,000,000

2500 years

Ice caps & glaciers

29,199,700

Groundwater (near-surface)

4,171,400

8 years

Lakes & Rivers

130,700

88 days

Soil Moisture

66,700

47 days

Atmosphere

12,900

9 days

fact it is this variability that must be accommodated for in


the design of water systems. Some interesting examples to
illustrate spatial variability include consideration of the
driest and wettest places on Earth. For the purposes of
illustration we can compare long-term average precipitation as
a representative measure of the local hydrology. For example,
Los Angeles receives an average of approximately 15 inches
(38 cm) of rainfall per year making it a relatively dry place.
However the driest place on Earth is the Atacama Desert in
Chile (Figure 1.8). In parts of the Atacama there has never
been recorded rainfall, and over the whole desert, rainfall
averages less than 1 mm/year! The lack of water makes for a
particularly stark landscape with little to no life to the point
that it is often compared to the surface of Mars (Figure 1.9).

F IGURE 1.8 Location of Atacama Desert in Chile (the driest


place on Earth).

In contrast, the wettest place on Earth is generally


considered to be Cherrapunji, India (Figure 1.10), which
receives an annual average rainfall of 11.1 m/year. Moreover,
due to its monsoon climate, most of this rainfall occurs within
the 3-6 month rainy season. Keeping in mind this is the longterm average, another interesting example is the maximum
observed annual rainfall in Cherrapunji, which was 26.5
meters!

These two examples provide interesting ends of the
hydrologic spectrum, where in one water is extremely scarce,
24

F IGURE 1.9 Illustrative photo of the arid terrain of the Atacama desert (from en.wikipedia.org/wiki/File:Paranal_360-degree_
Panoramic.jpg).

making water supply a problem. In the other, water is too


plentiful causing extreme flooding problems. Beyond these two
examples of extreme spatial variability in hydrology, it is
important to keep in mind the temporal variability. For
example, most of the 38 cm (15 in.) of annual average rainfall
in Los Angeles is isolated to the winter season (DecemberJanuary-February-March), which is in contrast to
Cherrapunji, which receives almost all of its rainfall during the
Summer. In doing any hydrologic analysis, it is important to
clearly define the relevant spatial and temporal scales of
interest.

F IGURE 1.10 Location of Cherrapunji India (the wettest


place on Earth).

25

S ECTION 7

Watershed Mass Balance



The last section used a global control volume to gain
some insight into long-term hydrologic fluxes at large scales
and the corresponding residence times in key hydrologic
reservoirs. In many hydrological analyses we are instead
interested in some local or regional scale. The question then
becomes: How do we choose an appropriate control volume for
such analyses? While any control volume can be chosen, one
that is physically-meaningful involves the concept of a
watershed (or catchment or river basin). The watershed is in
recognition of the fact that topographic slopes are the primary
drivers of many (especially surface) fluxes. Physically
speaking, water is expected to flow down hill into a nearby
stream channel where that water will then flow downstream
through the river network. Smaller tributary streams merge
with other streams forming larger and larger streams that will
ultimately tend to flow into the ocean or other large water
bodies. The river network is simply the organization of
drainage patterns over the landscape.

The definition of a watershed starts with the choice of a
point in space that coincides with a location on a particular
river network, typically referred to as the watershed outlet.
The outlet may be the terminal outlet (i.e. to an ocean or
other large body of water) or some location on the river
upstream from the terminal outlet (perhaps where a stream

gauge exists). A watershed is simply the set of all points


(upstream) that would ultimately route water to the defined
outlet point. If one has topographic (i.e. elevation) maps, the
watershed can be defined using a relatively simple set of rules.
This process is generally referred to as watershed delineation
and traditionally was done by hand using hard-copy
topographic maps. Now it is more commonly done in an
automated way using digital representations of topography
(discussed in more detail below). Watersheds come in all
shapes and sizes and any given watershed is made up of many
smaller watersheds (sub-watersheds or sub-basins). Examples
of some of the large watersheds covering the continental U.S.
are shown in Figure 1.11. Note that the full Mississippi River
basin covers approximately 50% of the land area of the

F IGURE 1.11 Major watersheds covering the United States


(from maps.howstuffworks.com/united-states-watersheds-map.htm).

26

continental U.S. The continental divide (Great Divide) along


the peak of the Rocky Mountains separates the watersheds
that drain to the Pacific Ocean or Gulf of California vs. those
that drain to the Atlantic Ocean or Gulf of Mexico. Of
particular relevance to California are the primary basins
draining the Sierra Nevada (Figure 1.12) and the Colorado
River (Figure 1.13), both of which provide the major water
supply to California. The western Sierra Nevada basins drain
into a large series of dams controlled by the California
Department of Water Resources (CADWR), which provides a
major supply of water to California, while two basins on the
eastern side of the Southern Sierra Nevada (Mono and Owens)
supply the Los Angeles Aqueduct. Water from the Colorado
River is conveyed via the Colorado River Aqueduct to
Southern California (Figure 1.14).

Digital topographic information can be used to
automatically delineate a watershed. These digital
representations are most commonly referred to as digital
elevation models (DEMs), where a map of elevation is
provided on a regular grid at some nominal spatial resolution.
Some DEMs are simply digitized versions of old topographic
maps and are available from agencies like the USGS (http://
eros.usgs.gov/#/Guides/dem), while others have been
developed from satellite-based algorithms (i.e. the SRTM
[http://www2.jpl.nasa.gov/srtm/] and ASTER [http://
asterweb.jpl.nasa.gov/gdem.asp] products). Based on these
various sources of data, much of the topography of the globe
is now available at spatial resolutions of less than 100 meters.
A DEM can be thought of as simply an array of elevation

F IGURE 1.12 Key watersheds in Sierra Nevada responsible


for water supply to California.

27

data on some coordinate grid system. (Such gridded data is


often called raster data). From the gridded data and a
specified outlet coordinate, one can derive the upstream pixels
that will ultimately flow to the specified outlet. A MATLAB
code for doing such watershed delineation is included in the
Modular Distributed Watershed Educational Toolbox (MODWET) provided as part of the textbook. The specific code is
called: watershed_area_and_stream_delineation.
Geographic Information System (GIS) software (e.g. ArcGIS)
also generally have built-in functions for watershed analysis.

The primary reason a watershed is a convenient control
volume is the fact that by construct it has one lateral flux (i.e.
runo at the outlet), where it is commonly assumed that there
is negligible groundwater flow between watersheds. The
instantaneous mass balance applied to a watershed control
volume is given by:

dS
= I i Oi = P E Q
dt
i
i

(1.7.1)

where S is the total storage of water within the basin (i.e.


snow, soil moisture, groundwater), P is the precipitation, E is
the evaporation and Q is the streamflow (runo at the outlet).
As can be done for any mass balance, if we take the long-term
average, this yields:

F IGURE 1.13 The Colorado River watershed (from


en.wikipedia.org/wiki/Colorado_River).

dS
0 = P E Q
dt

(1.7.2)

so that over the long-term the average basin yield (i.e. the
28

maximum amount of water that could be extracted) is given


by:

Q = P E

(1.7.3)

This simple form of mass balance is useful for a first-order


estimate of basin yield assuming long-term average
precipitation and evaporation are known. Other fluxes
discussed previously (i.e. infiltration, percolation, groundwater
flow, etc.) are internal fluxes when the entire watershed is
used as a control volume and therefore do not contribute to
the mass balance in this context.

It is important to keep in mind that the basin yield
shown in Equation (1.7.3) provides an estimate of the total
amount of available local water. In many regions, California
chief among them, local water supply is not enough to meet
local demand. In such instances water must be imported from
other watersheds. This is precisely what is done in California
where the Sierra Nevada (Figure 1.12) and Colorado River
(Figure 1.13) yield is conveyed to populated and agricultural
areas via a complex series of water resource systems. Figure
1.14 illustrates this system, where many dams are used (in
both the western Sierra Nevada and along the Colorado
River) to store water and aqueducts are used to convey water
from one location to the next. As seen in Figure 1.14,
Southern California gets water via the California, Colorado
River, and Los Angeles Aqueducts.
E XAMPLE 1.7.1

F IGURE 1.14 Map showing the key elements of the California


water system (from

en.wikipedia.org/wiki/File:California_water_system.jpg).

A new dam is built to create a municipal


reservoir at the outlet of a watershed. The
average annual rainfall for the region is 1.0 m y-1.
29

E XAMPLE 1.7.1 ( CONTINUED )

E XAMPLE 1.7.1 ( CONTINUED )

Evaporation pans in the area indicate an average


open-water evaporation rate of 1.1 mm d-1. The
river system brings an average inflow of 1.75 m3
s-1 into the reservoir.

mass or volume since the water density is eectively


constant. The storage volume can be expressed as
the(constant) area multiplied by the reservoir height.
The equation can then be written as:

a) How long would it take to fill the reservoir to


a depth of 3 m if no water is released through
the dam gates during the filling period? Assume
that the reservoir has an approximately constant
surface area of 10,000 hectares (1 ha = 10,000
m2).
b) Once the reservoir is filled, what is the
average reservoir release (in m3 s-1) that can be
made that will maintain the given reservoir level
at steady-state?
The relevant control volume for this problem is the
reservoir itself. Assuming no groundwater seepage into or
out of the reservoir, the mass balance equation can be
written as:

dS
= P E +Q
dt
where the storage and fluxes (precipitation, evaporation,
and inflow respectively) can be expressed in terms of

dh
= P E +Q
dt
where the fluxes on the right-hand-side can be written in
terms of volume flux densities, which are just the volume
fluxes per unit area:

P = 1.0 m y -1
! 1 m $ ! 365 d $
-1
E = 1.1 mm d #
&#
& = 0.4 m yr
" 1000 mm % " 1 y %
!
$ ! 365 d $ ! 1 $
3
-1 86400 s
-1
Q = 1.75 m s #
&#
& # 8 & = 0.55 m yr
" 1 d % " 1 y % " 10 m %
-1

The rate of storage change in the reservoir is then given


by:

dh
= (1.0 0.4 + 0.55) m y -1 = 1.05 m y -1
dt
Based on this rate, with no outflows, the reservoir will
take 2.9 years to fill up to a depth of 3 meters.
If there is a reservoir release (Qout) the (steady-state)
30

E XAMPLE 1.7.1 ( CONTINUED )


mass balance equation changes to:

dS
= P E +Q Qout = 0
dt
Qout = P E +Q = 1.05 m y -1
which, not surprisingly, is the same as the filling rate if
there is no outflow. The above flux can be converted
back to volumetric flux units by multiplying by the
reservoir area, which yields: Qout = 3.3 m3 s-1.
In practice, most reservoirs are primarily responsible for
smoothing out natural variability so that in wet years
the reservoir may add storage and in dry years storage
may be depleted. The size of the reservoir is a key design
variable that needs to be large enough to meet demands
or attenuate floods with a specified reliability criteria.
Note that in this simple example the reservoir area is
constant, whereas in most cases, the area will increase as
the reservoir fills.

31

S ECTION 8

MOD-WET Codes

Throughout the course, several of the equations and/or
concepts that can be applied numerically are implemented for
student use in MATLAB code. Together these functions are
referred to as the Modular Distributed Watershed
Educational Toolbox (MOD-WET). The MOD-WET codes
are available as a companion to the book and are provided to
help cement basic applied concepts for students. They are
generally implemented as functions (i.e. with inputs/outputs)
rather than scripts. This makes them modular in structure so
they can be used in conjunction with other MOD-WET codes
or additional codes you develop on your own. They are listed
at the end of each chapter where the relevant concepts are
introduced. The actual functions provide documentation
about the inputs/outputs/units and other details. The user
should carefully examine each code before applying it so that
you are aware of the details behind the calculations.

Relevant functions based on concepts introduced in this
chapter include:
Watershed delineation code:

watershed_area_and_stream_delineation.m

32

S ECTION 9

Conceptual Questions

9. What property of water is used in the context of


evaporation to convert between a mass flux density and an
energy flux density?
10. What are the SI units of an energy flux density?

1. Name at least four hydrologic reservoirs.


2. Name at least four hydrologic fluxes.
3. Is water a polar or non-polar molecule? Explain.
4. How does one convert between a volumetric flux and a
volumetric flux density? What are the typical units of each
type of flux?
5. What is the largest global water reservoir? What is the
largest freshwater global reservoir? It it an active or
relatively static component of the hydrologic cycle? What is
the largest active freshwater reservoir in the global system?
6. What is the nominal density of liquid water? What is the
nominal density of ice? Which is bigger? What is the
implication for a freezing lake?
7. Are the specific heat capacity and latent heats for water
generally large or small compared to other substances?
What are some of the implications of this?

11. If a watershed is used as a control volume in a mass


balance problem, what generally are the relevant hydrologic
fluxes in the mass balance?
12. If you wanted to estimate the long-term water yield of a
basin using mass balance, what assumption is often
invoked?
13. What is the average rate of the global hydrologic cycle (in
meters/year)? How does that compare to the driest and
wettest places on Earth?
14. What is the average residence time of water in the global
atmosphere? How does it compare to the residence time in
the groundwater system?
15. Describe what a watershed is. What is a sub-watershed?
16. What is a DEM?

8. What are the SI units for a volumetric flux? a mass flux? a


volumetric flux density? a mass flux density?

33

S ECTION 10

Sample Problems
Problem 1.1. Based on this chapter and background
reading references (e.g the chapter Tapping into a Planetary
Cycle from the book Introduction to Water in California
[Carle, 2004]) answer the following questions:
a) Name Earths hydrologic reservoirs where the great water
wheel takes place? What powers the water cycle (i.e. what is
the energy input that drives the hydrologic cycle)? What
fraction of the Earths freshwater is considered active and
which hydrologic reservoirs are associated with this water?
What fluxes transport water through the hydrologic cycle and
connect one reservoir to another?
b) What is the average annual precipitation rate in Los
Angeles, CA? In Eureka, CA? What period of the year does
California receive most of its precipitation? Which direction
do winds typically blow in California? How are these winds
related to the dry region of the Mojave Desert?
c) Why is the Sierra Nevada snowpack considered a water
reservoir? What technique has been used by water supply
planners to forecast water availability during spring-summer?
d) What technique has been used by scientists to reconstruct
periods of drought or wet years when a historic rainfall record
is not available? Using this method, experts have been able to

identify two epic drought periods in Californias history.


How long did these periods last? How does Californias
wetness in the 20th century compare with this longer
historical record?
Problem 1.2. Hydrologists throughout the world use a
variety of units. Although SI units are fairly universal, it is
useful to be familiar with other common unit conventions.
The following exercises should help achieve this familiarity.
a) The concept of volumetric fluxes is commonly expressed in
terms of a volume per unit area (flux density), or a depth.
This is generally the case in measuring rainfall or evaporation
over a known area, like a river basin, a lake, or a reservoir.
Los Angeles receives approximately 15 inches of rainfall per
year. How many meters of rainfall does Los Angeles receive
each year? How does this compare to a city in Northern
California, such as San Francisco, which receives about 559
mm of rain annually?
b) An acre-foot (AF) is a common unit to measure volumes of
water, such as reservoirs, canals, aqueducts, etc. It is defined
as the volume of water required to cover an acre of land with
a depth of 1 foot of water. The Los Angeles Basin covers an
area of 460 mi2. Using the average rainfall for Los Angeles,
what is the volume of water received over the basin in acre-ft?
In US gallons?
c) Streamflow is a measure of the volume of water per unit
time that passes a given point in a river or stream.
Streamflow is commonly measured in cubic feet per second
34

[cfs]. What is the discharge in the Saint Lawrence River, in


cfs, if the average discharge is 10,400 cubic meters per second
[m3 s-1]?
d) For water resources planning and management, it is
important to have an estimate of how much water one person
uses in one day. On average, one acre-ft is estimated to serve
the annual domestic needs of approximately six people (i.e.,
1AF/[6 people]/yr). How many gallons of water does one
resident use per year? Per week?
Problem 1.3. The relatively unique properties of water are
among the reasons why it plays such a key role in the Earth's
climate. Among those properties are the ones that relate
energy and water processes. Provide all answers in standard
SI units.
a) Define the dierences (in meaning) of heat capacity of a
substance and latent heat (i.e. of vaporization or fusion) of a
substance. Which one is used in the context of evaporation to
convert between a mass flux density and an energy flux
density?

e) Suppose you are making yourself 0.50 liters of tea but its
temperature is 75C, which is too hot for you to drink. In
order to cool its temperature, you immerse three ice cubes of
the same dimension as the one in part c) in your tea. If the
energy used to melt the ice came from the internal energy (i.e.
temperature) of the tea, what is the lowered temperature of
your tea?
f ) Suppose you do not drink all of the tea, and you leave 50
ml in the cup. How much energy would be required to fully
evaporate the remaining tea?If this energy came from the air
in a room surrounding the cup of tea, how would the air
temperature change, i.e. by what amount and via an increase
ordecrease in air temperature? Assume the room has
dimensions of 5 m x 5 m x 3 m and the air density is 1.2 kg/
m3.
Problem 1.4. The reservoir shown in cross-section below
receives inflow from a river with annual mean discharge Q =
2.55 m3s-1. Observations from a nearby precipitation station

b) What are the typical values of these properties for water at


0C?
c) What is the nominal density of liquid water? What is the
nominal density of ice? Which is bigger?
d) Suppose you have a 3 cm x 3 cm x 3 cm ice cube that is
frozen at a temperature of 0C.How much energy input would
be required to melt it?
35

yield an estimate of P = 1.0 m yr-1 as the mean annual


precipitation over the reservoir area of A = 80 km2.
For water planning purposes you need to estimate the losses
from the reservoir via evaporation (E) and leakage (G). There
is a proposal to estimate the losses by performing a chemical
tracer experiment. A harmless chemical tracer that has no
nearby natural sources (i.e. no occurrence in the stream or
large-scale groundwater flow) is mixed within the lake and its
concentration in the seepage water flux (S) below the dam is
measured. The seepage S is the combined leakage from the
reservoir (G) and large-scale lateral groundwater flow (U)
below the reservoir. The concentration of the tracer in the
lake is CL = 50 ppm (where ppm is parts per million, i.e. m3
tracer/million m3 water). The seepage flow is measured to be
1.27 m3 s-1 with a tracer concentration of CS = 10 ppm.
Estimate the two loss terms from the reservoir (in m yr-1)
through consideration of mass balance of water and chemical
tracer in the system. Make sure to clearly identify your
control volume in any mass balance application for the
analysis to be meaningful.
Problem 1.5. A new dam is built to create a municipal
reservoir. The average annual rainfall for the region is 1.7 m
yr-1. Evaporation pans in the area indicate an average
evaporation rate of 2.2 mm day-1. The river system brings an
average inflow of 3.5 m3 s-1 into the reservoir.
a) How long would it take to fill the reservoir to a depth of 10
m if no water is released through the dam gates during the
filling period? Assume that there is negligible groundwater

inflow and leakage from the reservoir and that it has an


approximately constant surface area of 10,000 hectares (1 ha
= 10,000 m2).
b) Once the reservoir is filled, what is the average reservoir
release (in m3 s-1) that can be made that will maintain the
given reservoir level?
c) From historical data, it is determined that the worst
drought on record reduced the annual precipitation and
streamflow in the region by approximately 50%. Given a
necessary yield (reservoir release) equal to your answer to part
b), would the reservoir drop below a level of 8 meters in such
a drought year?
Problem 1.6. Annual average data from three large river
basins around the world is given below:
AMAZON
RIVER
(BRAZIL)

YANGTZE
RIVER
(CHINA)

OB RIVER
(RUSSIA)

WATERSHED
AREA (km2)

7,180,000

1,970,000

2,950,000

PRECIPITATION
(mm/yr)

1,777

1,120

563

STREAMFLOW
(m3/s)

190,000

35,000

12,500

a) Compute the mean annual evapotranspiration (in mm/yr)


for each watershed. Clearly state any assumptions made in
your calculations.

36

b) For each watershed, compute the runo coecient (annual


river discharge as a fraction of precipitation) in %.

37

Chapter 2

Atmospheric
Thermodynamics

S ECTION 1

Learning Objectives
By the time you finish this chapter you should be able to:
1. List the key constituents of the atmosphere, specifically
including the radiatively active components

8. Sketch the typical profiles of temperature, pressure, and


density in the atmosphere
9. State the typical lapse rate of temperature in the
troposphere
10. State typical values for air pressure and density near the
surface

2. Write down the ideal gas law (equation of state) for moist
air in terms of air temperature, density, pressure, and
water vapor content
3. Use the ideal gas law to compute the density of air from
the other states
4. Define the concepts of saturated vapor pressure and
saturated specific humidity and how they are dependent
on temperature (i.e. the Clausius-Clapeyron equation)
5. List and define the typical descriptors used to identify the
actual water vapor content of air and be able to convert
between the dierent metrics
6. Describe the dierence between specific humidity or vapor
pressure and saturated specific humidity or saturated vapor
pressure.
7. Use the hydrostatic equation to relate atmospheric
pressure and density as a function of height

39

S ECTION 2

Atmospheric Composition

The atmosphere is a gaseous mixture of several
constituents, some of which are relatively constant, while
others are quite variable. The composition of the atmosphere
is shown in Table 2.1. By volume, the air in the atmosphere is
composed almost entirely of nitrogen and oxygen (together
equaling 99%). These two constituents and many others
(argon, neon, helium, krypton, hydrogen, and nitrous oxide)
represent the vast majority of the atmosphere and are
relatively constant in their concentration. Several components
vary as a result of natural or anthropogenic processes. The
most variable constituent is water vapor, which varies
naturally as a result of the hydrologic cycle. Carbon dioxide
varies both naturally (due to photosynthesis and respiration)
and anthropogenically (due to the burning of fossil fuels). Of
particular importance to climate and the hydrologic cycle are
those gases that are radiatively active (i.e. water vapor,
ozone, methane, carbon dioxide, and to a lesser extent
oxygen).

A radiatively active gas is one whose molecular structure
is such that it absorbs (and emits) radiative energy (this will
be discussed in much more detail in the next chapter).
Radiatively active gases are often referred to as greenhouse
gases. The primary hypothesis for anthropogenic climate
change has to do with the combined facts that 1) humans

T ABLE 2.1. P RIMARY C ONSTITUENTS OF THE


S TANDARD A TMOSPHERE IN THE T ROPOSPHERE
GAS

% BY VOLUME

Nitrogen (N2)

78.084

Oxygen (O2)*

20.946

Argon (A)

0.934

Carbon Dioxide (CO2)*

0.033

Neon (Ne)

0.00182

Helium (He)

0.000524

Methane (CH4)*

0.00016

Krypton (Kr)

0.0014

Hydrogen (H2)

0.00005

Nitrous Oxide (N2O)

0.000035

Ozone (O3)*
Water (H2O)*

0 - 0.000007
0-4

*radiatively active gases


have added a significant amount of carbon dioxide to the
atmosphere and 2) it is a greenhouse gas. In fact, water vapor
and methane are more radiatively active, but carbon dioxide
gets the most attention because it is the one that is increasing
rapidly due to anthropogenic sources. The vertical profiles of
some of these constituents will be discussed below. The clear
signature of both natural and anthropogenic eects on carbon
dioxide can be seen in measurements. An observatory on
40

in Figure 2.1 which is attributable directly to the


anthropogenic source from the burning of fossil fuel. As shown
in Figure 2.1, this source has led to a dramatic increase in
CO2 concentrations of approximately 25% in the last 50 years.
Moreover the trend is not linearly increasing, but accelerating
under present emission conditions. A major topic of ongoing
research is what impact does this significant increase in
carbon dioxide in the atmosphere have on the global climate,
weather patterns, and extremes.

F IGURE 2.1 Time series of measured atmospheric CO2 at

Mauna Loa Observatory in Hawaii shown in red along with a


moving average of the long-term trend shown in black (from
esrl.noaa.gov/gmd/webdata/ccgg/trends/co2_data_mlo.png).

Mauna Loa, Hawaii has provided relatively long-term records


of the carbon dioxide concentration (Figure 2.1). Two key
modes of variability are seen in the data. The first is the
natural seasonal variability, where CO2 decreases due to
increased uptake by deciduous plants in the Spring/Summer
season and increases due to respiration as a result of fallen
leaf (and other biomass) decay during the Fall/Winter. This
natural pattern should have some interannual variability, but
is expected to be relatively steady in terms of the long-term
average. The second mode of variability is the clear trend seen
41

S ECTION 3

Atmospheric States

The state of the atmosphere is the set of variables that
can be used to quantitatively describe it. The thermodynamic
states include temperature, pressure, density, and some metric
of water (vapor) concentration. It is important to keep in
mind that the atmosphere is simply a moving fluid (with the
above characteristics varying in space and time) so another
key state variable is the wind speed (which has components in
three orthogonal directions). Knowledge of the state of the
atmosphere at a given location/time will ultimately provide a
mechanism for quantifying many of the water and energy
movements in the atmosphere and between the atmosphere
and land surface. Some key questions that will be addressed in
this and other sections of this chapter include:
1. How are the variables that comprise the thermodynamic
state of the atmosphere related (at a given location)?
2. What metrics for water vapor concentration are used, and
related to some of these metrics, what is the upper limit to
the concentration of water vapor?
3. How do key variables vary with height and what
relationships govern this variation?

To start with the first question, it is worth pointing out
that the atmosphere behaves like an ideal gas (something

you should have covered in your physics course). Simply put,


an ideal gas is one whose density is relatively low, and the
atmosphere generally satisfies the necessary criteria. In
particular, one could treat each constituent of the atmosphere
(Table 2.1) as an individual ideal gas, i..e.

pi = i RiT

(2.3.1)

pi partial pressure of gas i [Pa]

i density of gas i [kg m -3 )


Ri gas constant for gas i [J kg -1 K-1 ]

T temperature of gas (mixture) [K]

Having an equation for each gas is generally unnecessary since


we are primarily interested in the bulk properties of the
atmosphere (not of individual gases). To simplify things, we
generally consider an air parcel in the atmosphere as
essentially a two-component mixture:
Moist Air = Dry Air (N2, O2, ...) + Water Vapor (H2O)
This idea is in recognition of the fact that most of the
constituents of the atmosphere are relatively fixed, especially
when compared to the high variability seen in water vapor.
Given this, the density of moist air is given by:

= d + v

(2.3.2)

where the d subscript refers to dry air and the v subscript


refers to (water) vapor. Substituting the ideal gas law
expressions for each of the two components yields:

42

pd
p
p
e
+ v = d +
RdT RvT RdT RvT

(2.3.3)

where pd and pv are the partial pressures of dry air and vapor
respectively, and Rd and Rv are the dry air gas constant (287 J
kg-1 K-1) and the vapor gas constant (461 J kg-1 K-1)
respectively. In many texts related to hydrology, the partial
pressure for water vapor (i.e. vapor pressure) pv is replaced
by e, which is what we will do here. Equation (2.3.3) can be
rearranged noting that the total pressure is equal to the sum
of the two partial pressures (i.e. p = pd + e) to form a more
compact equation of state (i.e. ideal gas law) for moist air:

Rd
p "
e%
1

(1

)
;

=
= 0.622
$
'
RdT #
p&
Rv

(2.3.4)

where the second term on the right-hand-side essentially


represents a correction for the presence of vapor. The ratio of
vapor pressure to total pressure (e/p) represents the fraction
of total pressure made up by the partial pressure of the water
vapor molecules. What Equation (2.3.4) also illustrates is that
an air parcel with more water vapor (i.e. higher e/p) will
actually be less dense than a parcel of dry air with the same
temperature and pressure. Without changing the above
equation, we can also define a new quantity called virtual
temperature (Tv) so that we can write:

p = RdTv ; where Tv =

e
1

(1

(2.3.5)

where in this case the moisture correction is embedded in the


virtual temperature definition. While virtual temperature can
be thought of simply as a definition, if one wishes to attach a
physical meaning to it, it denotes the temperature a parcel of
dry air would have to be at to have the same density as one
with a vapor pressure e.

As will be discussed in more detail below, the total
pressure of the atmosphere varies greatly with height, but a
typical value near the surface (at sea level) is 100,000 Pa or
1000 mb. Also note that the units required in the above
equations must be in SI.
E XAMPLE 2.3.1
What is the density of moist air at a pressure of
1000 mb, air temperature of 25C, and vapor
pressure of 25 mb? How does the density change
as a function of vapor pressure?
The virtual temperature under these conditions is:

Tv =

"
e%
$1 (1 ) '
p&
#
= 300.99 K

(25 + 273.15) K
"
2500 Pa %
$1 (1 0.622)
'
100000 Pa &
#

and the density is then given by:

43

E XAMPLE 2.3.1 ( CONTINUED )

p
100000 Pa
=
= 1.16 kg m -3
-1
-1
RdTv (287 J kg K )(300.99 K)

These answers can be confirmed with the MOD-WET


functions: virtual_temperature.m and air_density.m.
Note that this moist air density is about three orders of
magnitude smaller than the density of water. Nearsurface air density is typically between 1 and 1.2 kg m-3.
Based on the equation it is clear that an increase in
humidity (i.e. vapor pressure) will actually decrease the
density of air. Hence moister air will generally be less
dense and therefore more buoyant.

44

S ECTION 4

Metrics for Water Vapor


Concentration in Air

As shown in the previous section, the presence and
amount of water vapor in air impacts the thermodynamic
properties of air. Several metrics for quantifying vapor
concentration are used and are discussed here. It is important
to note that all are dierent ways of expressing the same thing
and therefore knowledge of one can always be used to get
another. In some cases, one may be measured and another
may be needed in a model. The most common descriptors are:
1. Vapor pressure (e): As described above, this represents the
partial pressure of water vapor. Pressure is simply the force
(per unit area) imparted by the kinetic motion of molecules,
so the more water vapor molecules present, the higher the
vapor pressure. The common units used for vapor pressure are
Pascals (Pa) and millibars (mb) where 100 Pa = 1 mb. The
vapor pressure amount clearly depends on the humidity of the
air (and varies with height in the atmosphere), but typical
values of vapor pressure in the atmosphere near the surface
are in the range 1000 - 3000 Pa.
2. Mixing ratio (w): This is a mass-based metric representing
the Mass of vapor/Mass of dry air and can be expressed as:

w=

v
e
=
d
p e

(2.4.1)

The units of mixing ratio are kg H2O/kg dry Air. It is


important to keep these explicit underlying units in mind, but
because it is a ratio of masses the mixing ratio is often given
as a unitless quantity.
3. Specific humidity (q): This is another mass-based metric
representing the Mass of vapor/Mass of moist air and can be
expressed as:

q=

v
e
=
v + d
p (1 )e

(2.4.2)

The units of specific humidity are kg H2O/kg Air. Note that


in many cases, and especially near the surface, p >> e (i.e.
100,000 Pa vs. 2000 Pa) in which case the above equations
could be written as:

q w

e
p

(2.4.3)

Typical values for specific humidity (or mixing ratio) near the
surface are 5-1510-3 kg kg-1, where the small values simply
represent the small fraction of vapor in the atmosphere (Table
2.1). Because of the small values, units of g/kg are often used,
in which case the typical values near the surface are 5-15 g
H2O/kg Air.

Several other metrics exist that are relative in nature.
Before defining these, the upper limit (water vapor holding
capacity) of the atmosphere must be discussed. From
thermodynamic principles one can define the saturated vapor
45

pressure (es), which simply represents the amount of vapor


that would be in equilibrium with a bulk water surface. In
fact, the curve between liquid water and vapor shown in the
pure water phase diagram (Figure 1.5) is illustrative of this
quantity. Note that the pressure in the phase diagram is for
pure water meaning that the pressure on the y-axis is in fact
the vapor pressure of water (since for pure water the partial
pressure is equal to the total pressure). It can be shown that
the saturated vapor pressure is given by the ClausiusClapeyron equation, which is defined by:

bottle with some air space above the liquid water. Given the
temperature of the system in the bottle, the equilibrium
amount of vapor in the air space would have a vapor pressure
equal to the saturated vapor pressure. In truth, some water
molecules will escape the liquid state and others will leave the
vapor state and condense on the liquid. The saturated vapor
pressure is the equilibrium condition where the rate of
vaporization of liquid water molecules is equal to the rate of
condensation of vapor molecules. An analogous quantity can
be defined over bulk ice and yields:

des Lv es
=
dT Rv T 2

L
ei (T ) = es 0 exp s
Rv

(2.4.4)

1 1

T0 T

(2.4.6)

This dierential form (i.e. an ordinary dierential equation


[ODE]) shows that the saturated vapor pressure is dependent
solely on temperature. If one ignores the relatively slight
temperature dependence of Lv, the above can be solved via
separation of variables, yielding the integrated form of the
Clausius-Clapeyron equation:

with the primary dierences being the substituted use of the


latent heat of sublimation and that this is only relevant for
sub-freezing temperatures. The focus of the equations below
are on the saturated vapor pressure over water, but that for
ice becomes relevant in cold clouds and snow processes as
described in later chapters.

L
es (T ) = es 0 exp v
Rv


In practice the saturated vapor pressure is to a very
good approximation the upper limit to the actual vapor
pressure in air, i.e.:

1 1

T0 T

(2.4.5)

where T0 and es0 are integration constants (273.16 K and 611


Pa respectively) chosen in the integration. Note that empirical
equations also exist that dier in form and take into account
the temperature dependence of the latent heat of
vaporization. The physical interpretation of es can be thought
of as the water holding capacity of air. Imagine a closed water

0 e es (T )

(2.4.7)

Based on the previous definition of specific humidity, we can


also define the saturated specific humidity:

qs

e ; with 0 q q s (T, p)
p s

(2.4.8)
46

This upper limit to vapor content or water vapor holding


capacity will turn out to be very important in both
evaporation and cloud/precipitation processes discussed later.
A key point that should be understood is that the actual
amount of vapor (i.e. e or q) in an air parcel is in general
dierent (less than) the saturated amount of vapor (i.e. es or
qs) except in the case of saturation. The saturated vapor
pressure or specific humidity are theoretical (not measurable)
upper limits on the measurable quantities e or q. In other
words, if the temperature of the air is known, you know its
water vapor holding capacity, but you have no knowledge of
the actual humidity in the air (unless saturation conditions
exist).

diurnal variability (even if the amount of actual vapor (i.e. the


numerator) is relatively constant).
5. Dew point temperature (Td): This metric is defined as the
temperature at which air of a given vapor concentration (e or
q) would need to be cooled (at constant pressure) to reach
saturated conditions. This can be defined via the integrated
Clausius-Clapeyron equation:

L
e = es (T = Td ) = es 0 exp v
Rv

1
1

T T
0
d

(2.4.10)

which can then be solved for Td :


1


These newly defined saturation quantities allow for the
definition of other metrics of vapor concentration in air:

1 Rv e
Td =
ln
T
L
es 0
0
v

4. Relative humidity (RH): This quantity is simply the ratio of


actual to saturated concentration, i.e.:

Based on the definition of dew point temperature, you should


be able to convince yourself of the following condition:

RH =

e
q
=
es q s

(2.4.9)

The above definition gives relative humidity as a fraction,


however, RH can also be defined as a percentage. Based on
the above definitions, it is expected that the RH of air will
generally vary between 0 (dry) and 100% (fully saturated air).
An important point to keep in mind is the strong temperature
dependence of RH due to the denominators dependence on T
via the Clausius-Clapeyron equation. Since temperature varies
strongly throughout the day, the RH generally has a strong

Td T

(2.4.11)

(2.4.12)

6. Vapor pressure deficit (VPD): This metric is another


relative one that identifies how close the air is to saturated air
via:

VPD = e = es e

(2.4.13)

which by definition is equal to zero under saturated conditions


and is positive under non-saturated conditions (when e < es).
It also highlights the fact that saturated vapor pressure and
actual vapor pressure are two dierent quantities.
47

7. Wet-bulb temperature (Tw): This metric is defined as the


temperature air would be cooled to by evaporating water into
it (at constant pressure) until saturation is reached. It is
defined from the first Law of Thermodynamics as:

T Tw
L
= v
q s (p,Tw ) q c p

(2.4.14)

where due to the nonlinearity in the saturated specific


humidity, this is an implicit equation that needs to be solved
iteratively. The metric gets its name from the a wet-bulb
thermometer which is used to measure it. A wet-bulb
thermometer is a regular thermometer with a saturated wick
covering the thermometer sensor. The evaporative cooling
from the wick cools the air in the immediate vicinity of the
sensor. When used in conjunction with a regular (dry-bulb)
thermometer (often together called a sling psychrometer), one
can deduce the humidity of the air based on the dierence
between the dry-bulb and wet-bulb temperatures.

Together the metrics defined above represent the most
commonly used variables for expressing the concentration of
water vapor in air. It is important to remember that each is
simply a dierent way of representing the humidity of air and
that it is often necessary to convert between two dierent
metrics.

E XAMPLE 2.4.1
For a parcel of air with a vapor pressure of 1500
Pa, a temperature of 20C, and a pressure of
98000 Pa, determine the following alternate
humidity metrics for the parcel:
a) vapor density, b) specific humidity, c) relative
humidity, d) dew point temperature, and e)
vapor pressure deficit
a) The vapor density is given by the ideal gas law for
water vapor (Equation (2.3.1)), i.e.:
v =

e
1500 Pa
=
RvT (461 J/kg/K)(293.15 K)

= 0.01 kg(H 2O)/m 3 (Air)

where note that this is about 1% of a typical moist air


density.
b) The specific humidity provides a mass-based
concentration, i.e.:

q =
=

e
p (1 )e
1500 Pa
98000 Pa (0.378)(1500 Pa)

= 0.0095 kg(H 2O)/kg(Air) = 9.5 g(H 2O)/kg(Air)

48

E XAMPLE 2.4.1 ( CONTINUED )

E XAMPLE 2.4.1 ( CONTINUED )

c) The relative humidity requires that the saturated


vapor pressure first be calculated. Using Equation
(2.4.6), the saturated vapor pressure is:

e) The vapor pressure deficit is given by:

VPD = e = 2368 Pa 1500 Pa = 868 Pa

) 2.5 10 6 J/kg #
&,
1
1
es = (611 Pa)exp +

%
(.
461
J/kg/K
273.16
K
293.15
K
$
'*
= 2368 Pa

where the larger the vapor pressure deficit the drier the
air is (i.e. relative to saturation). Air with a relative
humidity of 100% will have a vapor pressure deficit equal
to zero.

The relative humidity is just the ratio of the actual


vapor pressure to the saturated vapor pressure, which in
this case is equal to 63.4%.

These answers can be confirmed with the MOD-WET


functions: vp_to_specific_humidity.m,
sat_vapor_pressure.m, vp_to_RH.m,
vp_to_dew_point_temperature.m, and
vapor_pressure_deficit.m.

d) The dew point temperature is defined as the


temperature that air would have to be cooled to so that
it reached saturation and is given implicitly by solving
for the temperature at which e = es, which can be
inverted to yield:
1

)
# 1500 Pa &,
1
461 J/kg/K
Td = +

ln
%
(.
6
* 273.16 K 2.5 10 J/kg $ 611 Pa '= 286.1 K = 13. C

This means that if the air were cooled by 7C it would


reach saturation. By definition, at the dew point
temperature the air would have a relative humidity of
100%.

49

S ECTION 5

Vertical Profiles of Atmospheric States



It is important to discuss the primary variability of the
atmospheric states with height, specifically typical profiles for
temperature, pressure, density and humidity. The collection of
typical state profiles are often referred to as the standard
atmosphere. At any moment in time and space the profiles
may vary greatly from these standard profiles.

The typical temperature profile is used to define dierent
layers of the atmosphere (Figure 2.2). The lowest layer of the
atmosphere is called the troposphere and is approximately 10
km in depth. This is the layer most relevant to hydrologists
since it is where most of the weather processes occur and
where most of the water vapor exists. The temperature in the
troposphere is reasonably linear (decreasing with height), i.e.:

T(z) = T0 z,

T0 surface temperature

(2.5.1)

lapse rate in temperature


z altitude

The lapse rate in the troposphere is approximately 6.5 K/km


and the surface temperature is approximately 290 K. The
reason most of the water vapor and weather processes are

F IGURE 2.2 Plot of standard profiles of temperature, density,


and pressure in the atmosphere (from en.wikipedia.org/wiki/
File:Comparison_US_standard_atmosphere_1962.svg).

limited to the troposphere is due to the increasing


temperature with height in the stratosphere. This stratified
temperature profile is very stable (i.e. warmer air overlying
colder air) and acts as a lid to vertical motions. Exceptions
to this include extreme thunderstorms and volcanic eruptions,
50

both of which can penetrate the tropopause into the


stratosphere.

Pressure and density profiles are tightly coupled via the
hydrostatic equation, which can be derived from the balance
of pressure and atmospheric weight forces to yield:

dp
= g
dz

(2.5.2)

where the density itself depends on pressure, temperature, and


vapor pressure via the Ideal Gas Law (Equation (2.3.4)).
Combining Equations (2.3.4) and (2.5.2) along with the
decreasing temperature with height (Equation (2.5.1)) leads to
decreasing (near exponential) profiles for both pressure and
density with height (Figure 2.2). The typical values for
pressure and density at the surface are approximately 100,000
Pa and 1.2 kg m-3 respectively. At the top of the troposphere
the pressure and density drop o to about 25,000 Pa (250 mb)
and 0.4 kg m-3 respectively. The air density is negligible above
40 km (mid-stratosphere). Also, because pressure decreases
monotonically with height, it is often used interchangeably
with altitude as a vertical coordinate.

There is no standard water vapor profile, but there are
key characteristics that can be described. Figure 2.3 shows
representative profiles for three dierent latitude bands. In
general, the higher latitudes have less vapor since they are
colder. The lower humidity is largely controlled by the
saturated specific humidity governed by Clausius-Clapeyron
(i.e. colder air cannot hold as much vapor). The key feature is

F IGURE 2.3 Typical profiles of specific humidity in the troposphere at different latitudes.

that the specific humidity quickly drops o with height and is


essentially zero at the top of the troposphere. The reason for
this profile is that the primary source of water is at the
surface and the stratosphere prevents vertical movement
beyond the lower layer of the atmosphere.

51


To characterize the bulk amount of vapor in the
atmospheric column, we often use the so-called total
precipitable water, which is simply the integration of water
vapor mass (per unit volume, i.e. vapor density) with height:

Wp =

dz

(2.5.3)

which using SI will have units of kg (H2O)/m2. The physical


interpretation of precipitable water is that this would be the
amount (mass) of water that would be yielded (per unit area)
if all vapor were condensed out of that column of air. Since
vapor density is diculty to measure, the above expression
can be rewritten in a more convenient form using the
definition of specific humidity and the hydrostatic equation as
given by:

Wp =

1
v dz = q dz =
g
0

1
q dp = g
ps

ps

q dp

(2.5.4)

Note that in cloudy atmospheric columns the liquid and ice


water could also be considered. Often precipitable water is
expressed in depth units (i.e. equivalent depth of water) using
the density of liquid water as a conversion. Typical values of
precipitable water are on the order of 0 - 5 cm. In cases where
only near-surface data is available (i.e. full profiles of specific
humidity are not available), empirical equations exist, i.e.
(Dingman, 2008):

Wp = 1.12 exp(0.0614Td )

(2.5.5)

E XAMPLE 2.5.1
Estimate the precipitable water from the tropical
specific humidity profile in Figure 2.3. Assume a
standard atmospheric pressure profile. How does
that estimate compare to the one from the
empirical expression shown in Equation (2.5.5)?
Without data for the specific profile, it can be
approximated via a piecewise linear profile where at
altitudes of 0, 3, 5, 7, and 9 km, the corresponding
specific humidities are approximately given by 16, 5.5,
2.5, 1.5, and 0.5 g/kg respectively. The atmospheric
pressures corresponding to each level can be roughly
estimated from the standard atmosphere shown in Figure
2.2 as 101300, 70120, 54050, 41110, and 30800 Pa
respectively for the altitudes listed above.
The integral in Equation (2.5.4) can then be evaluated
numerically (i.e. using the trapezoidal rule):

1
Wp =
g

ps

q dp g 2 (q
p

i+1

+ qi )(pi+1 pi )

= [(16 + 5.5) 10 3(101300 70120)Pa +


(5.5 + 2.5) 10 3(70120 54050)Pa +
(2.5 + 1.5) 10 3(54050 41110)Pa +
(1.5 + 0.5) 10 3(41110 30800)Pa] / 2 / 9.81 m s -2
= 44.4 kg m -2 = 4.44 cm

where in this empirical equation [Wp] = cm and [Td] = C.


52

E XAMPLE 2.5.1 ( CONTINUED )

E XAMPLE 2.5.1 ( CONTINUED )

Hence the integrated amount of water in the tropical


profile is equivalent to a depth of 4.4 cm of liquid water.

In this particular case the two estimates are relatively


close (i.e. within 10%).

Now we can compare this estimate to one obtained using


the empirical equation. It should be noted that whenever
you are using an empirical equation, it was developed
from the specific dataset analyzed and therefore is not
always generalizable to other cases. Also, in this case the
empirical expression relies only on near-surface humidity.
This makes its application much easier since a full profile
is not needed, but it implicitly depends on an assumed
profile which may dier from the one being used. Based
on the surface specific humidity, the surface vapor
pressure is given by:

These answers can be confirmed with the MOD-WET


functions: specific_humidity_to_vp.m,
precipitable_water.m, and Wp_from_Td.m.


An illustrative example of the variability in total
precipitable water on the timescale of days is shown in Movie
2.1. The key features are that the highest precipitable water
occurs over equatorial regions (where oceans are prevalent and

pq (101300 Pa)(15.5 10 3 kgkg -1 )


e
=
= 2524 Pa

0.622
The dewpoint temperature is then given by:
1

)
# 2524 Pa &,
1
461 J/kg/K
Td = +

ln %
(.
6
273.16
K
611
Pa
2.5

10
J/kg
$
'*
= 294.2 K = 21. C

The precipitable water from Equation (2.5.5) is then


given by:

Wp = 1.12 exp(0.0614(21)) = 4.1 cm

M OVIE 2.1 Illustration of hourly precipitable water pattern

evolution (over the ocean) as measured from the AMSU satellite (from sos.noaa.gov/videos/PrecipitableWater.mov).
53

evaporation is highest). Also visible are so-called atmospheric


rivers, which represent relatively narrow bands of high
precipitable water values that often bring vapor on-shore. A
specific example of an atmospheric river is the so-called
Pineapple Express, which consists of intermittent events that
bring a significant amount of vapor (and often subsequent
precipitation) from the region of the Pacific Ocean near
Hawaii to the northwestern U.S. coastal regions. Movie 2.2
shows another example in terms of the monthly average
precipitable water over the globe. The key features seen in the
animation are the latitudinal dependence of water vapor, with
highest values near the equator, and the seasonal north-south

shift, which is driven by seasonal eects of the sun as will be


described in the next chapter.

M OVIE 2.2 Illustration of monthly averaged precipitable water over the globe from July 2002-July 2012 (from
earthobservatory.nasa.gov/GlobalMaps/).

54

S ECTION 6

MOD-WET Codes

Relevant functions based on concepts introduced in
this chapter include:
Air density:

air_density.m
Bisection (root finding) function:

bisect.m
Vapor pressure (from dew point temperature):

dew_point_temperature_to_vp.m
Script with definitions of key physical constants used in
other functions:

physical_constants.m

Saturated vapor pressure (Clausius-Clapeyron) over water:



sat_vapor_pressure.m
Vapor pressure (from specific humidity):

specific_humidity_to_vp.m
Vapor pressure deficit:

vapor_pressure_deficit.m
Virtual temperature:

virtual_temperature.m
Dew point temperature (from vapor pressure):

vp_to_dew_point_temperature.m
Relative humidity (from vapor pressure):

vp_to_RH.m
Specific humidity (from vapor pressure):

vp_to_specific_humidity.m

Precipitable water (from profile data):



precipitable_water.m

Wet-bulb temperature (from vapor pressure):



vp_to_wet_bulb_temperature.m

Vapor pressure (from relative humidity):



RH_to_vp.m

Vapor pressure (from wet-bulb temperature):



wet_bulb_temperature_to_vp.m

Saturated vapor pressure (Clausius-Clapeyron) over ice:



sat_vapor_pressure_ice.m

Precipitable water (empirical function from dew point):



Wp_from_Td.m

55

S ECTION 7

Conceptual Questions
1. Name at least two radiatively active gases in the
atmosphere.
2. What gas in the atmosphere is most variable?
3. What is the general trend of carbon dioxide in the
atmosphere. What are its two key modes of variability?
Explain.
4. Name the atmospheric state variables that are related to
each other via the equation of state for moist air (i.e. ideal
gas law).
5. With all else equal, is an air parcel with more vapor
concentration less dense or more dense than a parcel with
less vapor?

9. What is the dew point temperature a metric of? Does


knowledge of the dew point temperature generally tell you
anything about the actual temperature? Explain.
10. Name the layer of the atmosphere that is most relevant to
hydrologists. Explain why.
11. How does the temperature profile generally vary within
the troposphere?
12. How do the pressure and density profiles generally vary
within the troposphere? What equation governs the
pressure profile?
13. How does the specific humidity profile generally vary
within the troposphere?
14. Describe in words what precipitable water is. Where on
the globe is precipitable water generally highest/lowest?
15. What is an atmospheric river?

6. What, if anything, does the saturated vapor pressure of the


atmosphere tell you about the actual vapor pressure of the
atmosphere?
7. If the relative humidity of air is 100%, what do you know
about the specific humidity?
8. If the specific humidity is constant throughout the day, how
would you expect the relative humidity to change (if at all)
throughout the day. Explain.
56

S ECTION 8

Sample Problems
Problem 2.1. Several atmospheric constituents, most notably
water vapor, greatly aect many of the processes we will
study. Therefore, it is important to be able to estimate how
much water is in the atmosphere over an area of interest.
Additionally, it is important to become familiar with various
temperature and humidity relationships. Consider the
following:
a) A parcel of moist air is at 870 mb, with a temperature of
12C and a relative humidity of 95%, what is the moist
density of the air in kg m-3?
b) What is the saturated vapor pressure of air (in Pa) at a
temperature of 37C? Explain the physical meaning of
saturated vapor pressure. How does it change as the
temperature increases? How does it change as the humidity
level of air changes?
c) What is the specific humidity (in g/kg) of air in part a)?
d) What temperature would a dry parcel of air need to have
in order to have the same density of the air in part a)?
e) What is the dew point temperature (in C) of the air in
part a)? What is the meaning of the dew point temperature?

Problem 2.2. A parcel of moist air has a temperature of


31C, a pressure of 950 mb, and a specific humidity of 8 g/kg.
Using the appropriate equations, compute the following
properties of the air parcel:
a) vapor pressure
b) virtual temperature
c) density
d) mixing ratio
e) relative humidity
f ) dew point temperature
g) vapor pressure deficit
h) wet bulb temperature
i) Qualitatively, how (if at all) would each of the above
quantities [i.e. in parts a)-h)] change if the air temperature
decreased (with air pressure and specific humidity fixed)?
Problem 2.3. A large room in a museum has a 3-meter high
ceiling and its lateral dimensions are 10 by 15 meters. To
preserve the artwork, it is imperative that the specific
humidity be kept at 10 g/kg and air temperature maintained
between 20 and 25 degrees Celsius. Due to a power outage,
the climate control system responsible for maintaining
conditions in the room fails.
You are called to help fix the problem and by the time you
arrive your measurements indicate that the air relative
humidity is 80% with an air temperature of 21 degrees Celsius
and air pressure of 1000 mb. You have a portable dehumidifier
57

that takes in air from the closed room, removes water vapor
by condensing the vapor into liquid, and then discharges the
dry air back into the closed room. You run the dehumidifier
until the specific humidity is lowered to the required value.
a) What mass of water (in kilograms) will need to be
condensed out to reach the required humidity level?
b) What is the amount of energy released (in Joules) as a
result of condensing the humidity?
c) Assuming the energy computed in part b) goes into
warming the air, what is the expected rise in air temperature
due to the dehumidification. Note: The specific heat capacity
of air is 1004 J/kg/K. For your calculation, you may assume
that the change in air density (or change in air mass) is
negligible as a result of the heating/dehumidification.
d) Will the air temperature still be in the necessary range as
a result of the dehumidification or will it need to be
additionally cooled?
Problem 2.4. In winter-time, heated indoor air reduces the
relative humidity inside buildings. This is especially a problem
inside hospitals in cold climates where the air temperature
and relative humidity must be kept at 69F (20.6C) and 75%
respectively for health and hygiene reasons.
Consider a cold winter day in Chicago when the outside air
temperature is 20F (-6.7C) and the relative humidity is 90%.
(Assume air density is 1.2 kg m-3 and air pressure is 1000 mb).

a) What is the relative humidity inside the hospital building if


the air is brought from outside and heated to the required
temperature, but not humidified?
b) Consider a hospital building with 1500 m3 volume. It has
a humidifier system that vaporizes 1 gallon hr-1 of water
(264.2 gallons = 1 m3). How many hours should the humidifier
be in operation to increase the relative humidity of the indoor
air to the regulation limit?
Problem 2.5.
a) Combine the hydrostatic equation and the equation of
state (ideal gas law) to show that the pressure profile for a
general virtual temperature profile (Tv) is given by:

# g
p(z) = p0 exp %
$ Rd

dz &
T (z)(
'
0
v

b) Derive the expression for the pressure profile for the case
with a linear virtual temperature profile:

Tv (z) = Tv 0 z
which is a reasonable approximation to the profile in the bulk
troposphere (i.e. with a constant temperature lapse rate).
Problem 2.6. Several atmospheric constituents, most notably
water vapor not only impact water fluxes, but greatly aect
the radiative properties of the atmosphere. Therefore, it is
important to be able to estimate how much water is in the
atmospheric column over an area of interest. The
58

thermodynamic state of the atmosphere is often characterized


using radiosondes, which are essentially weather balloons that
take measurements as they ascend through the atmosphere.
Radiosonde data collected at Hilo, Hawaii on June 21, 2012
showed that the virtual temperature of the air decreases
linearly with a lapse rate of:

equation shown above? How would you convert from these


units to a depth (e.g. cm of water)?

T = 5 K km -1

" z%
v (z) = v 0 exp $ '
# H&

as the radiosonde altitude increases from the surface (0 km


above sea level [asl]) to 8 km asl. The observed surface virtual
air temperature is equal to 28 C. (Note: A linear decrease is
usually a reasonable approximation to the virtual air
temperature profile in the troposphere). Surface air pressure is
equal to 1000 mb.
a) Based on the derived pressure profile in the previous
problem, plot the pressure profile for this case. What is the air
pressure at an altitude of 8 km?
b) A common metric that is used to characterize the total
amount of vapor is precipitable water:

Wp =

(z)dz
0

c) Suppose the observed water vapor density from the surface


to an altitude of 8 km is described by the following
exponential decay function:

where the surface vapor density is equal to 11 g (H2O) m-3,


and H is a length scale describing how quickly the variable
decays with height. What is the precipitable water (in cm of
water) of the 8 km atmospheric column observed by the
radiosonde in Hilo on June 21st?
d) Increased greenhouse gases are expected to result in an
increase in atmospheric temperature. The Intergovernmental
Panel on Climate Change (IPCC) reports that the increase in
the next 100 years may be between 0.6- 3.4 K. Conceptually,
what is the eect that the increase in temperature will have
on maximum water storage in the atmosphere (e.g. Is air
capable of holding more or less water?)? Comment on how
this might aect precipitation (frequency and rates).

which simply measures the integrated amount of water vapor


in the vertical column from height 0 to height z. To get the
total amount of water you would integrate to the top of
atmosphere. What are the units of precipitable water in the

59

Chapter 3

Radiation
Processes

S ECTION 1

Learning Objectives
By the time you finish this chapter you should be able to:

9. Describe what factors at the surface will contribute to


variability in incoming shortwave radiation at a point (in
addition to the basic Earth-Sun geometry eects that
determine TOA fluxes)
10. Define net radiation at the land surface and its
components

1. Write down the integrated radiation equation for emission


from a blackbody

11. Use the empirical Idso formula to estimate atmospheric


emissivity

2. Write down the integrated radiation equation for emission


from a graybody

12. List the necessary inputs (measured/estimated


quantities) needed to compute net radiation

3. Define emissivity, albedo, transmissivity, and absorptivity

13. Compute incoming solar radiation at the land surface

4. List the primary factors controlling spatial/temporal


variability in incoming solar radiation at the top of
atmosphere (TOA)

14. Compute incoming/outgoing longwave radiation at the


land surface

5. Draw a schematic of the radiation budget of the Earth

15. Compute net radiation at the land surface

6. Write down typical values for emissivity and albedo of the


Earths land surface
7. Describe how (and based on what variables) the
atmosphere impacts the amount of TOA shortwave
radiation that reaches the surface (in both clear/cloudy
skies)
8. Describe how (and based on what variables) the
atmosphere impacts the amount of longwave radiation
that reaches the surface (in both clear/cloudy skies)
61

S ECTION 2

Basics of Radiation

Radiation is a form of energy that plays a crucial role in
hydrology. Radiative energy provided by the sun is the
external forcing of the system that drives the global
hydrologic cycle and atmospheric and oceanic motions. The
Earths surface and atmosphere reflect and absorb energy
from the sun as well as emit radiative energy of their own.
The absorbed energy and energy fluxes tends to drive water
fluxes and transformations (i.e. phase changes). Due to the
importance of radiative fluxes we would like to understand the
mechanisms responsible for the variability in radiative forcing
and develop simple models for these fluxes. We will start with
a quick refresher on basic radiation physics. For a much more
detailed treatment of radiation processes, the reader is
referred to Liou (1992, 2002).

Radiation is a form of energy carried by electromagnetic
(EM) waves (or photons). These waves travel at the speed of
light and can be characterized by either wavelength (the
length of a single period of the wave) or frequency (the
number of periods of the wave passing a given point per unit
time). The relationship between the two is given by:

= c
wavelength [m]
c speed of light [3.0 10 8 m/s]

(3.2.1)

F IGURE 3.1 Electromagnetic spectrum showing shortwave


and longwave radiation regions as well as sub-regions of the
spectrum (i.e. visible, infrared, microwave, etc.).

which shows that the speed of light can be used to convert


between wavelength and frequency. A schematic illustrating
the electromagnetic spectrum, which simply corresponds to
the characterization of such radiation across the spectrum of
wavelength/frequency is shown in Figure 3.1. Of particular
relevance to hydrology are those waves between ultraviolet
(nanometer scale) and microwave (centimeter scale) that
contain non-negligible amounts of energy.

From physics we know that the amount of energy
radiated per unit area per unit wavelength (i.e. spectral
radiant emittance) by a so-called blackbody (i.e. a perfect
emitter) is given by the Planck function:

frequency [s 1 = Hz]
62

R = B =

2 hc 2

(3.2.2)

5 ehc/(K T ) 1

which has SI units of W m-2 m-1 and where h is the Planck


constant (6.62610-34 J s), K is the Boltzmann constant
(1.380610-23 m2 kg s-2 K-1), and T is the physical
temperature of the body. What this equation tells us is that:
i) the emitted radiation occurs over a spectrum of wavelengths

or frequencies and ii) is controlled solely by the physical


temperature of the emitting body. An illustrative example of
this dependence is shown in Figure 3.2. The key
characteristics of the figure are that the peak is a strongly
increasing function of temperature (note log-scale on figure),
while the center of mass of the spectrum shifts to shorter
wavelengths with increasing temperature. For illustration, the
blackbody emission spectrum is shown for the Sun and Earth.

If we are interested in the bulk (integrated) radiative
energy flux from a body, then we can integrate the spectral
radiant emittance across all wavelengths, i.e. for a blackbody:

R = B d

(3.2.3)

Based on the analytical form of the Planck function one can


show that the above integral yields the so-called StefanBoltzmann equation:

R = T 4 [Wm 2 ]

(3.2.4)

Stefan-Boltzmann constant = 5.67 10 8 W m 2 K4

F IGURE 3.2 Planck function as a function of temperature and


wavelength. The yellow and red curves corresponds to the approximate blackbody emission that would occur from the Sun
and Earths surface respectively (from en.wikipedia.org/wiki/
File:BlackbodySpectrum_loglog_150dpi_en.png).

It turns out that the sun, to a good approximation, behaves


like a blackbody. However, many other natural media (i.e. the
Earths surface and atmosphere) are not. Instead they are
referred to as graybodies, where the spectral irradiance is
given by:

R = B ;

spectral emissivity []

(3.2.5)
63

The spectral emissivity represents the eciency of emission


(varying between 0 and 1), i.e. how close the body is to a
blackbody, at a given wavelength. Note that in general the
emissivity can be a strongly varying function of wavelength.
By definition a blackbody has an emissivity equal to 1.0 at all
wavelengths. The integrated flux from a graybody can be
obtained by integrating across all wavelengths as above
introducing a broadband emissivity:

R = T 4
broadband emissivity []

(3.2.6)

This can be thought of as a general equation for the


integrated radiative flux with a blackbody being a special
case.

64

S ECTION 3

Shortwave vs. Longwave


Radiation

Based on the Planck function it is evident that
temperature plays the key role in the intensity and
distribution of emitted radiative energy from a given body. In
the Earth system, we are generally dealing with two sources of
radiative energy: solar radiative fluxes (emitted by the sun)
and terrestrial radiative fluxes (emitted by the Earths surface
and atmosphere). The temperature of these two sources are
vastly dierent, with the temperature of the Suns surface
approximately 5800 K, while terrestrial temperatures are on
the order of 250 - 300 K. Based on these temperatures, the
relative spectral irradiance for solar (at the top-ofatmosphere) and terrestrial sources are shown in Figure 3.3.
As expected, the location of the peaks are dierent (i.e.
approximately 0.6 microns for solar and approximately 10
microns for terrestrial). Moreover, what becomes clear is that
there is very little overlap in the two spectra; i.e. the right tail
of the solar spectrum becomes negligible around 4 microns
and is approximately coincident with the left tail of the
terrestrial spectrum.

The solar irradiance distribution covers three primary
parts of the EM spectrum: ultraviolet (in the left tail), visible
(around the peak), and near-infrared (on the right side of the
distribution). The human eye is only sensitive to visible light;

Shortwave: Rs =

=4 m

=0.1 m

R d

F IGURE 3.3 Distribution of shortwave and longwave radiation.

when you see something, what your eye is really sensing is the
solar (visible) radiation being reflected from that object. The
objects color corresponds to the wavelengths of the visible
light that are reflected (rather than absorbed).

The terrestrial spectrum covers two dierent parts of the
EM spectrum: infrared (thermal infrared and far infrared) and
microwave (in the tail of the distribution beyond 50 microns).
The peak occurs in the thermal infrared (around 12 microns).
The basic principle of night-vision goggles takes advantage of
this fact where the sensors are tuned to be sensitive to the
thermal infrared region. So even though little visible light may
be available, the thermal infrared signature is still apparent.
65


Due to these dierences, we often treat the two fluxes
separately, in terms of both measurement and modeling. The
solar fluxes are referred to as shortwave radiation (spanning
approximately 0.1-4 microns) and terrestrial fluxes are
referred to as longwave radiation (spanning approximately
4-100 microns). Formally, the integrated fluxes can be written
as:
Shortwave: Rs =

Longwave: Rl =

=4 m

=0.1 m

=100 m

=4 m

R d

(3.3.1)

R d

(3.3.2)

Measurements of these fluxes are made by looking at these


distinct bands.

66

S ECTION 4

Radiative Properties of
Media

As shown above, an emitted radiative flux from a
particular body depends on its physical temperature and its
emissivity. The broadband emissivity strongly depends on the
media characteristics. For land surfaces, the surface type
largely dictates the broadband emissivity (see Table 3.1) and
typically ranges between 0.9-1.0. The broadband emissivity of
the atmosphere strongly depends on the concentration of
radiatively active gases, i.e. conceptually:

= f ([H 2O],[CO2 ],[CH 4 ],...)


Since the gas concentrations can vary in time and space, the
emissivity itself varies. The most variable radiatively active
gas is water vapor and hence most models for the broadband
emissivity of the atmosphere depend explicitly on the water
vapor concentration. Examples will be discussed in more
detail in Section 7 where longwave flux modeling is discussed.

While the emissivity is a radiative property that
describes the eciency of radiation emission, other radiative
properties are important in terms of how incident radiation
interacts with media. For example, at the top of the
atmosphere, shortwave radiation is incident on the atmosphere
in a downwelling direction. Another example is upwelling
longwave radiation emitted by the land surface that is

T ABLE 3.1. T YPICAL BROADBAND EMISSIVITY FOR


VARYING LAND SURFACE TYPES ( FROM A RYA , 2001)
SURFACE TYPE

EMISSIVITY

Water

0.92-0.97

Snow (old)

0.82-0.89

Snow (fresh)

0.90-0.99

Ice

0.92-0.97

Bare soil

0.84-0.97

Grass (long, 1 m)

0.90

Grass (short, 0.02 m)

0.95

Agricultural crops

0.90-0.99

Forests

0.97-0.99

incident on the bottom of the atmosphere. We are interested


in what happens to these radiative fluxes as they interact with
media. To begin to quantify these interactions we define three
additional radiative properties of a media: the transmissivity
(t), absorptivity (a), and reflectivity (r). Each of these
represent an eciency of a particular radiative process.

Transmissivity represents the fraction of incident
radiation that gets transmitted through a media. On a
molecular level this represents photons (radiation) that do not
directly interact with the molecules in the media and therefore
directly passes through it. Absorptivity represents the fraction
of incident radiation that gets absorbed by the media, i.e. the
time-averaged fraction of incident photons (radiation) that hit
individual molecules and are absorbed. Reflectivity represents
the fraction of incident radiation that is scattered in all
67

directions, i.e. the time-averaged fraction of incident photons


that hit individual molecules and are reflected. It is important
to note that these properties are not independent, but linked
via energy conservation at a particular wavelength, i.e.:

t + a + r = 1

(3.4.1)

which is simply a statement that all incident energy on a


given media (at a particular wavelength) is either transmitted,
absorbed, or reflected.

Applying the above conservation principal to the
atmosphere, we can write for the shortwave band (i.e. using
an s subscript for shortwave):

ts + as + rs = 1

(3.4.2)

and for the longwave band (i.e. using an l subscript for


longwave):

tl + al + rl = 1

F IGURE 3.4 a) Normalized blackbody radiance spectrum for


(3.4.3)

the sun and earth and b) absorption as a function of wavelength and molecular absorption (from homepages.ius.edu/
kforinas/ClassRefs/ElectroMagnetic.html).

The shortwave reflectivity is often referred to as albedo.



As with emissivity in the atmosphere, each of these
properties are functions of the greenhouse gas concentrations
(where absorption by radiatively active gases is highly
dependent on wavelength). An illustration of the absorptivity
of the atmosphere across both the shortwave and longwave
spectra is shown in Figure 3.4 along with the primary
molecules responsible for absorption. One can see that even
over relatively short wavelength intervals, the absorption can
vary from close to 100% to close to 0%. Key features seen in

the figure include almost complete absorption of UV radiation


by ozone (O3), relatively little absorption in the visible region,
with the highest absorption occurring near 7 microns due to
water vapor, and key absorption features in the longwave due
to carbon dioxide, methane, and water vapor (among others).
Not shown on the figure is the microwave region (above 100
microns), which has relatively low absorption. The
determinant of the atmospheric windows (i.e. relatively
transparent due to low absorption) versus regions where the
68

atmosphere is opaque (high absorption) depend primarily on


the radiatively active gases in the atmosphere along with
scattering due to molecules and aerosols.

Some rough approximations can be made for the
radiative properties in dierent parts of the spectrum and for
atmosphere vs. land surface. In general in the atmosphere,
scattering within the longwave spectrum is considerably
smaller in the clear sky atmosphere compared to
transmissivity and absorptivity, such that:

tl + al 1 tl 1 al

(3.4.4)

which simply states that the fraction of longwave radiation


transmitted through the atmosphere is essentially equal to the
fraction that is not absorbed. This simple model explains one
of the primary mechanisms for global warming. As the
concentration of greenhouse gases increase, the (longwave)
absorptivity increases thereby decreasing the amount of
transmitted surface longwave radiation through the
atmosphere (that would otherwise ultimately leave the Earth
system at the top of the atmosphere). This absorbed radiation
is then re-emitted by the atmosphere, a fraction of which is
emitted toward the surface ultimately increasing the surface
temperature.

The land surface is essentially opaque to both shortwave and


longwave radiation (i.e. the transmissivity is small) and
similar to the atmosphere, longwave reflectivity is small so
that:

as + rs 1 as 1

(3.4.7)

al 1

(3.4.8)

surface albedo

which simply states that the fraction of incident shortwave


radiation that is not reflected is absorbed and most of the
incident longwave radiation is absorbed.


The same conservation principle can be applied to the
land surface for shortwave and longwave fluxes:

ts + as + rs = 1

(3.4.5)

tl + al + rl = 1

(3.4.6)
69

S ECTION 5

Modeling Top-ofAtmosphere Shortwave


Fluxes

The solar radiation entering the top of the Earths
atmosphere is the key external driver of the Earth system.
Understanding how much of the radiation at the TOA is
absorbed by the atmosphere, reaches the surface, and is
reflected back to space provides a key basis for understanding
the temporal and spatial variability of the hydrologic cycle. As
stated earlier, the sun acts like a blackbody emitter at a
temperature of approximately 5780 K. Hence the amount of
radiation emitted by the surface of the sun is given by:

F IGURE 3.5 Schematic showing general picture of a location

on Earths TOA relative to the incoming solar beam and the resulting solar zenith angle.

4
Tsun
(5.67 10 8 W m 2 K4 )(5780 K)4 = 6.33 10 7 W m 2

This amount of energy is radiated in all directions. The Earth


only intercepts a small fraction of this amount (the ratio of
the surface area of the Sun to the surface area of a sphere
defined by the Earths distance from the Sun). The solar
beams are essentially parallel by the time they reach the top
of the Earths atmosphere and the incoming radiation is
generally referred to as the solar constant (I0):

I 0 = 1367 W m 2

which would be the amount of radiative energy intercepted by


a surface perpendicular to the incoming beam (Figure 3.5).
This amount of radiation then gets projected dierently at
dierent locations on the globe as locations are more or less
perpendicular to the incoming beam. Figure 3.6 illustrates
this where the same flux density (i.e. W m-2) is projected over
an area with a length scale of 1000 km near the equator, but
double that nearer the poles. Hence the amount of energy
(perpendicular to the TOA) is much less at higher latitudes
than it is at lower latitudes. This simple eect of geometry is
70

F IGURE 3.7 Schematic showing the solar position in terms of


zenith (angle from vertical) and azimuth (angle in horizontal
plane from North) angles.

F IGURE 3.6 Illustration of how an equal amount of energy

(i.e. the solar constant -- here scaled by the ratio of the intercepted area to the global surface area) is distributed over larger
areas as one moves further from the location normal to the incoming beam.

to first-order responsible for most of the spatial and temporal


variability seen in hydrology, weather, and climate.

The position of the sun relative to a location on the
Earth (either at the surface or TOA) can be uniquely defined
by two angles: the solar zenith angle and the solar azimuth
angle (Figure 3.7). The zenith angle is a measure of the
departure of the incoming solar beam direction from the
vertical at the point of interest. The solar zenith angle is 0 if
the sun is directly overhead (i.e. exactly perpendicular to the

tangent plane) and 90 at the horizon (i.e. exactly parallel to


the tangent plane). It is this angle which describes how
perpendicular the surface is to the incoming beam and
therefore how much the solar constant is reduced. The
azimuth angle is an angle in the horizontal plane, usually
referenced to be 0 (or 360) at due North and 180 at due
South.

Recall that the Earth is tilted (at approximately 23.5)
relative to the orbital plane around the Sun (Figure 3.8). This
tilt combined with the orbit around the Sun is the reason for
seasonal variability over the course of a year. Additionally, the
orbit around the sun is not circular, but elliptical which
introduces another (relatively minor) mode of seasonal
71

provides an illustration of the diurnal progression of day-night


(the so-called terminator) over the globe on the winter
solstice. On this day, the Earths tilt (relative to the sun) is at
its maximum of 23.5 which shows up directly in the
terminator line angle. On either equinox, the terminator
would line up directly with the Earths axis of rotation. Of
particular note is that the south pole experiences daylight
throughout the day and the southern hemisphere experiences
a longer day compared to the northern hemisphere (the
opposite would be true for the summer solstice).

F IGURE 3.8 Plane of the elliptical orbit of the Earth around

the Sun indicating the Earths tilt as well as its orientation as a


function of day of year (from physicalgeography.net/fundamentals/
images/ecliptic_plane.jpg).

variability as the actual distance between the Sun and Earth


varies throughout the year. Nominally, the Summer season
corresponds to the time of year when the sun is closest and
more directly overhead, with the opposite being true for
Winter. As a result, the seasons are at opposite times of the
year for the Northern and Southern Hemisphere due to the
tilt of planet. For the Northern Hemisphere, the summer
solstice represents the longest day of the year since the
Earths Northern hemisphere is pointing most directly at the
Sun and vice versa for the winter solstice (Figure 3.8). The
opposite is true for the Southern Hemisphere. Movie 3.1

M OVIE 3.1 Animation showing the minute-by-minute evolution of the day-night terminator for the winter solstice (from
sos.noaa.gov/videos/oneday.mov).

72


Figure 3.9 provides an illustration of the diurnal and
seasonal evolution of the solar position for a given location (in
this case 40 North latitude). The details of how this position
is calculated is given below; here only the qualitative aspects
are discussed. The sun is generally lowest in the sky at the
winter solstice: only 26.5 above the south horizon (i.e. a solar
zenith angle of 63.5) at solar noon. Solar noon is simply the
time corresponding to the highest point in the sky on a given
day (and is generally dierent that local noon which is
impacted by time zones, daylight saving time, etc.). At this
location the sun is 73.5 above the south horizon (i.e. a solar
zenith angle of 16.5) at solar noon on the summer solstice.

For the Spring and Fall equinoxes, the sun is 50 above the
south horizon (i.e. a solar zenith angle of 40) at solar noon
since there is no relative tilt between the Eartha axis and the
axis of the orbital plane on those days.

Based on the above factors, the amount of incoming
shortwave radiation at any location/time at the TOA is a
function of Earth-Sun geometry which is completely defined
by: i) Latitude (i.e. location), ii) Hour of day (due to rotation
of the Earth), and iii) Day of year (due to tilted axis and
elliptical orbit around Sun). Several models for the TOA flux
based on these inputs are available at varying levels of
precision. A relatively simple model for the flux at the
location perpendicular to the TOA is:

cos 0
I0
, daytime: 0 90
2
Rs 0 =
d

0,
nightime

(3.5.1)

where the cosine of the solar zenith angle is given by:

cos 0 = sin sin + cos cos cos

F IGURE 3.9 Schematic showing the seasonal evolution of solar position in the sky at a latitude of 40 degrees North.

(3.5.2)

# 2
&
23.45
declination angle =
cos %
(172 DOY )(
180
$ 365
'
latitude
T 12
hour angle = 2 h
24
# 2
&
d = 1 + 0.017 cos %
(186 DOY )(
$ 365
'
73

where in the above equations: DOY represents the day of year


(i.e. January 1st: DOY = 1 and December 31st: DOY = 365),
Th represents the solar hour of the day (midnight = 0, solar
noon = 12, 11pm = 23), and d represents the distance
between the sun and Earth normalized by the mean distance.
In applications, data are often tied to dierent time
coordinates including Universal Time Coordinate (UTC; same
as Greenwich Mean Time) or local time (which may or may
not include daylight saving time) rather than solar time. In
such cases, the time should be first converted to solar hour
before applying the above equations. MOD-WET functions for
Equations (3.5.1) and (3.5.2) are provided in solar_geometry
and TOA_incoming_solar.

The above model shows that the projection of the solar

F IGURE 3.11 Mean daily insolation (in W m-2) on a horizon-

tal surface at the top of the atmosphere as a function of the day


of the year and the latitude. White areas correspond to the polar night (from stratus.astr.ucl.ac.be/textbook/images/image(2).jpg).

F IGURE 3.10 Seasonal evolution of the solar declination angle (from physicalgeography.net/fundamentals/images/declination.jpg).

constant is primarily controlled by the solar zenith angle,


which is governed by the latitude of the location of interest,
the declination angle, and the hour angle. The elliptical orbit
plays a smaller role via the denominator in Equation (3.5.1).
The declination angle is the latitude at which the Sun is
directly overhead. Due to the tilt of the Earth, it varies
throughout the year between 23.5N (the Tropic of Cancer)
and 23.5S (Tropic of Capricorn) latitude (Figure 3.10). So the
declination angle is the primary seasonal control on the
incoming flux perpendicular to the TOA. As will be discussed
74

further below, the declination angle is also relevant to


radiation received by sloped surfaces. In the Northern
Hemisphere above 23.5N, the Sun is always in the Southern
sky. Hence south facing slopes will get exposed to more
radiation than north facing slopes. The opposite is true for
the Southern Hemisphere. Finally, the diurnal variability in
TOA solar flux comes through the hour angle term.

To illustrate the seasonal and spatial variability of the
TOA insolation, the daily-averaged amount can be computed

and plotted as a function of latitude and DOY (Figure 3.11).


The main features are that the Northern Hemisphere gets the
highest insolation in the summer and lowest in Winter with
the opposite true for the Southern Hemisphere. It is also clear
that the seasonal variability is least at the equator and
highest at the poles (where one can see the signs of perpetual
daytime during the summer and perpetual nighttime during
the winter). Implicit in Equation (3.5.1) is that the day length
itself varies considerably throughout the year. An example of
the variation in day length (for a location at 45 latitude) is
shown in Figure 3.12. A primary take-home message is that
many hydrologic fluxes will be directly aected by solar
forcing and as such will exhibit many of the variabilities seen
in the characteristics discussed here.
E XAMPLE 3.5.1
Compute the top-of-atmosphere solar flux over
Los Angeles (latitude of 34 degrees North) at
solar noon on the summer solstice (June 21;
DOY 172) and winter solstice (December 21;
DOY 355).
The earth-sun orbital parameters for DOY 172 (June 21)
are given by:

F IGURE 3.12 Illustration of seasonal evolution of day length


at a particular latitude.

" 2
%
23.45
cos $
(172 172)' = 0.409 rad
180
# 365
&

75

E XAMPLE 3.5.1 ( CONTINUED )

= 2

12 12
= 0 rad
24

" 2
%
d = 1 + 0.017 cos $
(186 172)' = 1.0165
# 365
&
The cosine of the solar zenith angle is then given by
(using radians for all angles, where the latitude is equal
to 0.593 radians):

cos 0 = sin(0.409)sin(0.593) +

cos(0.409)cos(0.593)cos(0)

= 0.983 0 = cos 1(0.983) = 0.185 rad = 10.6


where because this is the summer solstice and at solar
noon the solar zenith angle is relatively low. Note: This
day/time corresponds to the lowest solar zenith angle
over the entire year. The TOA solar flux is then given
by:

Rs 0 = (1367 W m -2 )

0.983
-2
=
1300.5
W
m
(1.0165)2

The earth-sun orbital parameters for DOY 355


(December 21) are given by:

E XAMPLE 3.5.1 ( CONTINUED )

" 2
%
23.45
cos $
(172 355)' = 0.409 rad
180
# 365
&

12 12
= 0 rad
24
" 2
%
d = 1 + 0.017 cos $
(186 355)' = 0.9835 rad
# 365
&

= 2

The cosine of the solar zenith angle is then given by


(using radians for all angles, where the latitude is equal
to 0.593 radians):

cos 0 = sin(0.409)sin(0.593) +

cos(0.409)cos(0.593)cos(0)

= 0.539 0 = cos 1(0.539) = 1.002 rad = 57.4


where because this is the winter solstice and at solar
noon the solar zenith angle is much higher. Note: This
day/time corresponds to the highest noon-time solar
zenith angle over the entire year. The TOA solar flux is
then given by:

Rs 0 = (1367 W m -2 )

0.539
-2
=
761.7
W
m
(0.9835)2

Note that this is less than 60% of the summer solstice


value. The projection of the solar constant at the much
higher zenith angle results in a significant reduction.
76

E XAMPLE 3.5.1 ( CONTINUED )


This large dierence is due solely to the dierence in
earth-sun geometry between the two dates. How these
energy inputs are attenuated by the time they reach the
surface will depend on the properties of the atmosphere.
These answers can be confirmed with the MOD-WET
functions: TOA_incoming_solar.m and
solar_geometry.m.

77

S ECTION 6

Modeling Surface Shortwave Fluxes



The previous section introduced the basic concepts
controlling the amount of shortwave flux incident at the TOA.
Often of particular interest in hydrology is how much of this
TOA flux actually reaches the surface. It is the surface
incident radiation that will drive processes like evaporation.
The TOA flux is attenuated (reduced) by greenhouse gases,
aerosols/dust, and clouds before it reaches the surface. The
specific processes which cause the attenuation are absorption
and scattering. A schematic example of these processes during
daytime conditions is shown in Figure 3.13. Note the figure
shows the solar constant at the TOA, implying the location is
at a latitude corresponding to the declination angle. In general
the amount at the TOA will not be I0, but rather Rs0.
Nevertheless, the relative fractions attenuated due to
absorption and scattering provide a qualitative description of
the various sources and relative magnitudes of attenuation.

As shown in Figure 3.13, there are two primary
components of shortwave radiation reaching the surface. One
is the so-called direct beam radiation. This is the amount of
beam radiation that does not interact with the atmosphere
and is therefore completely transmitted. The direct beam
transmissivity varies with optical depth of the atmosphere (a
function of greenhouse gases and clouds). As shown in Figure

F IGURE 3.13 Schematic showing how solar direct beam radiation is absorbed and scattered in the atmosphere (from
powerfromthesun.net/Book/chapter02/chapter02.html).

78

3.13, this transmissivity varies very roughly from 33-83%. The


low end of this range would be for cloudy skies, moist air,
and/or high zenith angle (i.e. during the winter and closer to
sunrise/sunset). The high end of this range would be expected
to correspond to clear sky conditions with a low zenith angle
(i.e. near solar noon and during the summer). The second
component is the so-called diuse radiation. This is radiation
that originated as direct beam at the TOA, but was scattered
(generally multiple times and in varying directions) before
reaching the surface. The fraction of diuse is generally
smaller than direct beam (5-26%), but also depends on
atmospheric conditions and solar zenith angle. Before sunrise
or after sunset or in overcast conditions, the illumination
provided at the surface is exclusively composed of diuse solar
radiation (Figure 3.14; in this case very evident as diuse flux
from clouds). One additional term contributing to incident
radiation at the surface, and not shown in Figure 3.13, is the
backscattered flux. This is the amount of direct and diuse
flux that is reflected from the surface and then backscattered
by the atmosphere. This terms is always smaller than the
direct beam and diuse fluxes.

To model these eects and get the total incident
shortwave at the surface we can define:

Rs = Rs,dir
+ Rs,dif
+ Rs,bs

Rs total downwelling shortwave flux

Rs,dir
downwelling direct beam flux

Rs,dif
downwelling diffuse flux

Rs,bs
backscattered flux

(3.6.1)

F IGURE 3.14 Photograph showing diffuse radiation being reflected off of clouds near sunset.


In terms of estimating these fluxes, models of varying
complexity can be used. Of high complexity are those models
generally referred to as radiative transfer codes. These require
a full accounting of the profiles of atmospheric constituents
and surface conditions and attempt to model the absorption,
scattering and transmission as radiation interacts with each
discretized layer of the atmosphere. These types of models are
commonly used as modules within global climate models,
where the atmospheric state profiles are explicitly modeled at
every time step. Measurements of such profiles are not
generally available with sucient sampling density in space/
79

time and so for oine hydrologic analysis, simpler semiempirical bulk models are often used instead. A general
example of such a bulk model is given by:

Rs,dir
= tsRs 0

T ABLE 3.2. T YPICAL BROADBAND ALBEDO FOR


VARYING LAND SURFACE TYPES ( FROM A RYA , 2001)
(3.6.2)

ts direct beam shortwave transmissivity


R

s,dif

= Rs 0

(3.6.3)

diffuse scattering coefficient

Rs,bs
= (Rs,dir
+ Rs,dif
)

(3.6.4)

surface broadband albedo


The coecients in each equation represent the bulk
broadband (i.e. integrated across the entire shortwave
spectrum) atmospheric transmissivity, scattering eciency,
and surface reflectivity. The broadband albedo of the surface
is primarily dependent on surface type (Table 3.2). The
transmissivity and scattering coecient are expected to vary
with atmospheric characteristics.

Putting the three equations together with the definition
of the total flux yields:

Rs = ts + + ts + 2 Rs 0

(3.6.5)

which shows that the total incident shortwave at the surface is


simply the TOA flux reduced by a multiplicative coecient
that varies between 0 - 1. It is very important to keep in mind
that the coecients shown in these equations must depend in
some way on atmospheric conditions and other factors alluded

SURFACE TYPE

ALBEDO

Water (small zenith angle)

0.03-0.10

Water (large zenith angle)

0.10-1.0

Snow (old)

0.40-0.70

Snow (fresh)

0.45-0.95

Ice

0.20-0.45

Bare soil

0.05-0.40

Grass (long, 1m)

0.16

Grass (short, 0.02m)

0.26

Agricultural crops

0.10-0.25

Forests

0.05-0.20

to above. This is generally where the semi-empirical nature of


the models show up, where models for these coecients have
been developed based on experimental data and that may be
used with data available at (or nearby) a given study site.

Many examples of such empirical models exist. Generally,
clear-sky models are easier to develop so that clouds are dealt
with via an additional attenuation factor, although physically
clouds tend to simply change the above defined coecients
(i.e. decrease transmissivity and increase scattering). Models
are also generally developed for flat horizontal surfaces, with
complex topography eects creating additional complexity.
One such example (Dingman, 2008) uses:

ts = sa dust

(3.6.6)
80

sa = exp(asa + bsa M opt )

optical depth, precipitable water, or dust).

asa = 0.124 0.0207Wp


The model shown above is applicable to clear sky
conditions. To take into account clouds, an additional
attenuation must be included. For the purposes of generality,
Equation (3.6.5) could be rewritten as:

bsa = 0.0682 0.0248Wp

dust attenuation due to dust ( 0-0.13)


= 0.5(1 s + dust )

(3.6.7)

s = exp(as + bsM opt )

(3.6.9)

where fsc(cloud) represents an empirical attenuation factor for


clouds in the shortwave spectrum that depends on some
identifiable cloud characteristics. Some examples of this
empirical function include (Bras, 1990):

as = 0.0363 0.0084Wp
bs = 0.0572 0.0173Wp

where in the above equations Mopt is the optical depth of the


atmospheric air mass and Wp is the previously defined
atmospheric precipitable water (here in units of [cm]). It is
very important to note that empirical functions such as these
are often not dimensionally consistent as a result of regression
to data, so that care must be taken with respect to the units
that are used. A first-order approximation used for optical air
mass depth is the inverse of the cosine of the solar zenith
angle (Bras, 1990), i.e.:

M opt = 1 / cos

Rs = fsc (cloud)"#ts + + ts + 2 $% Rs 0

(3.6.8)

More accurate representations can be found in Dingman


(2008) or other sources. These equations show that for an
increase in optical depth, precipitable water, or dust (i.e. more
molecules for interaction including greenhouse gases) there is
a decrease in the amount of transmission and an increase in
the amount of scattering (and vice versa for the case of less

fsc (C ) = 1 0.65C 2

(3.6.10)

where C is areal cloud cover fraction (i.e. 0 for clear-sky and 1


for completely overcast) or (U.S.A.C.E, 1956; Dingman, 2008):

fsc (C ) = 1 (1 ksc )C

(3.6.11)

where,

ksc = 0.18 + 0.0853Z c

(3.6.12)

where Zc is the cloud base height (in kilometers). The


attenuation factor fsc should vary between 0 - 1. Models based
solely on cloud cover fraction do not make any attempt to
incorporate the possible variability due to other cloud
characteristics (i.e. cloud thickness, water content, etc.).
MOD-WET functions for Equations (3.6.6)-(3.6.8) and
(3.6.10) are shown in Section 9 which can be used (along with
the TOA flux) to compute surface solar radiation fluxes.
81

E XAMPLE 3.6.1

E XAMPLE 3.6.1 ( CONTINUED )

For the summer solstice TOA value computed in


Example 3.5.1, estimate the incident shortwave
radiation on a short-grass surface for a clear-sky
day with an atmospheric precipitable water equal
to 3.5 cm and dust coecient of 0.1.
Qualitatively, how would things change if the
atmosphere were more humid, more smoggy, or
cloudy?

= 0.5(1 0.83 + 0.1) = 0.27

Based on the solar noon zenith angle, the optical depth


for this time is approximately equal to 1/0.983 = 1.02.
The atmospheric shortwave transmissivity can be
estimated by (Equation (3.6.6)):

asa = 0.124 0.0207(3.5) = 0.196

bsa = 0.0682 0.0248(3.5) = 0.155

sa = exp((0.196) + (0.155)(1.02)) = 0.70

ts = 0.70 0.1 = 0.6

which indicates about 60% of the TOA shortwave flux


will be transmitted directly. The atmospheric scattering
coecient can be estimated by (Equation 3.6.7)):

as = 0.0363 0.0084(3.5) = 0.066


bs = 0.0572 0.0173(3.5) = 0.118

s = exp((0.066) + (0.118)(1.02)) = 0.83

which indicates about 27% of the TOA shortwave flux


will be scattered toward the surface. The total flux
reaching the surface is then given by:

Rs = [0.6 + 0.27 + (0.27)(0.26)(0.6) + (0.27)2(0.26)]Rs 0


= [0.6 + 0.27 + 0.04 + 0.02](1300.5 Wm -2 )
= 1208 Wm -2
which shows that about an equivalent of 93% of the
TOA flux reaches the surface via transmission, forward
scattering or backscattering (in decreasing level of
importance). This fraction is relatively high because it is
the summer solstice. For the winter solstice, the optical
depth would be significantly larger which would decrease
the amount reaching the surface.
A more humid atmosphere would increase the
precipitable water and decrease the flux reaching the
surface. A more smoggy atmosphere would decrease the
transmissivity and increase the scattering, but likely lead
to an overall surface flux reduction. A cloudy atmosphere
would also generally decrease the amount of radiation
reaching the surface (see Equation (3.6.10)).

All of the models described above provide estimates of
the downwelling surface shortwave on a horizontal plane. In
areas of complex topography (i.e. mountainous regions)
82

additional factors must generally be considered. The primary


impact of terrain on incoming shortwave flux is due to varying
exposure to the sun. In the same way that the solar zenith
angle relative to a tangent plane modifies the solar constant at
the TOA (Equation 3.5.1), the local zenith angle due to a
sloping surface can enhance/reduce the amount predicted for
a horizontal plane (Figure 3.15). The local illumination at a
given point in space is a function of two parameters: the slope
(or strictly the slope gradient) and aspect (or slope aspect)
angles.


The slope (S) is the angle in the vertical plane between
the tangent plane in the steepest descent direction and a
horizontal plane. Based on this definition, a horizontal surface
has a slope of 0 degrees while a vertical surface has a slope of
90 degrees. The aspect (A) angle is the azimuth angle in the
horizontal plane between the steepest descent direction and an
arbitrarily defined reference direction (usually North). So a
point that is facing due North, East, South, or West is
generally labeled with an aspect of 0, 90, 180, or 270 degrees
respectively. Based on the slope and aspect angles, one can
define the local illumination (zenith) angle, which is simply
the zenith angle of the solar direct beam relative to the sloped
surface:

cos s = cos 0 cos(S) + sin 0 sin(S)cos(0 A)

(3.6.13)

s local illumination (zenith) angle


0 solar zenith angle

0 solar azimuth angle


S slope angle
A aspect angle

The cosine of the local illumination angle scales the direct


beam flux reaching the surface in the same way the cosine of
the solar zenith angle scales the solar constant at the TOA.

F IGURE 3.15 Schematic showing how slope/aspect can impact incident shortwave radiation.


In most applications, the slope and aspect are generally
computed from gridded DEM data. Several methods can be
used to compute the slope and aspect angles. The most
commonly applied computes the slope at a point by fitting a
plane to the neighboring nine cells (inclusive of the point of
83

interest) in the DEM and operating on the elevation data


matrix z(x,y) as:
2
2

dz dz
tan(S) = +
dx dy

dz dx
tan(A) =
dz dy

(3.6.14)

(3.6.15)

where the derivatives are usually computed via a finite


dierence approximation. An example of a DEM and the

F IGURE 3.16 DEM illustrating the distribution of elevation


(in meters) over an example terrain.

derived slope and aspect maps for a mountainous region are


shown in Figures 3.16-3.18 (using MOD-WET function:
generate_slope_and_aspect_from_DEM). Note, as expected
based on the definitions, the slope is smallest at valley floors
and ridge tops with some hillslopes having slopes over 50
degrees. For aspect, the north facing slopes are red/pink and
south facing slopes are blue/green. The key question is: why
are these variables important to hydrology? It is important to
keep in mind that together the slope/aspect control the
exposure of a given point to the sun. As mentioned above, in
most of the Northern Hemisphere the sun is in the southern
sky. This means that south facing slopes will generally get

F IGURE 3.17 Distribution of slope (in degrees) of the terrain


represented by the DEM in Figure 3.16.

84

snow accumulation. North facing slopes will generally have a


significantly longer persistence of snow on the ground
(sometimes weeks or months) due to the smaller snowmelt.

The other main influence that topography can have on
radiation fluxes is a result of shading. The solar zenith and
azimuth angles change continuously diurnally and seasonally.
For a given solar geometry (i.e. zenith/azimuth angle) a point

F IGURE 3.18 Distribution of aspect (in degrees) of the terrain represented by the DEM in Figure 3.16.

more direct beam radiation than north facing slopes. Hence


one would expect more evaporation or snowmelt on south
facing slopes and less on north facing slopes. A good
illustration of this can be seen in Figure 3.19. Despite
relatively close physical proximity to each other, the two
dierent facing slopes show a stark dierence in vegetation
density (which are driven in this case by soil moisture
patterns). The south facing slope would be expected to be
drier due to increased evaporation from the higher radiation
inputs. As a result it is much more sparsely populated by
vegetation than is the northern sloping face. The same can
generally been seen in mountainous regions that experience

F IGURE 3.19 Photograph illustrating the impact of slope and


aspect effects on radiation on other hydrologically relevant
properties, in this case soil moisture and vegetation distribution. The slope on the left is north-facing and the slope on the
right is south-facing (from en.wikipedia.org/wiki/File:
Slope_effect.JPG).

85

on the ground is either exposed to the sun or the suns


position is obstructed by neighboring topography. When the
suns position is obstructed the point will be in shade. This
means that the direct beam flux is zero and any shortwave
radiative flux input is due solely to diuse and backscattered
radiation (Equation (3.6.1)). Similar to the impacts described
above relative to aspect, these patterns of shading can directly
impact the patterns of hydrologically relevant variables (snow,
soil moisture, vegetation, etc.). So in areas of complex terrain,
the topography will have many direct impacts on hydrology
(in addition to flow direction which will drive runo as
described in Chapter 10).

to the horizontal resolution of the underlying DEM. In the


case of no shade and zero slope this reduces to Equation
(3.6.2). The shading mask is a dynamic quantity that is equal
to 0 when the sun is below the local horizon for a given pixel
and is equal to 1 when the sun is above the local horizon. The
MOD-WET function topo_shade_calc can be used to
generate these dynamic maps.

For diuse flux, the primary impact of topography has to
do with the so-called sky view factor (SVF), which is an
integrated measure of the amount of sky seen by the ground
at a given point. By definition, a horizontal plane has a SVF
of 1 and an infinitely deep and narrow canyon would approach


To quantify the impact of topography and shading, the
modified geometry can be accounted for to correct the
horizontal plane fluxes. Muller and Scherer (2005) formulated
the slope/aspect and shade eects into a single correction
term for direct beam flux received on a given unit area as:

s,dir

"
cos s
1 %
= tsRs 0 $maskshade
'
cos

cos(S)
#
&
0

(3.6.16)

where the term in brackets is the correction and maskshade is a


binary variable that is 0 when the pixel is in shade and 1
when in direct light. Note that the TOA flux contained the
cosine of the solar zenith angle which is in the denominator of
the correction term. So the correction removes that angular
dependence and instead multiplies by the cosine of the local
illumination (zenith) angle. The last term in the correction is
a geometric enlargement term that projects the sloped surface

F IGURE 3.20 Distribution of the sky-view factor (SVF) of the


terrain represented by the DEM in Figure 3.16.

86

a SVF equal to 0. An example of the SVF for the same


topography shown in Figure 3.16 is shown in Figure 3.20. The
ridges, which have a horizontal tangent plane, have a SVF of
1, while some of the valleys have values of 0.7 or lower. The
implication of having SVF less than 1.0 is that some of the
sky that would otherwise contribute to forward scattered
shortwave radiation is obscured. The diuse flux can be
modified to account for SVF via (Muller and Scherer, 2005):

Rs,dif
= Rs 0SVF + Rs (1 SVF)

(3.6.17)

where the first term represents the original scattered


contribution reduced by an obscured sky view, while the
second term represents a contribution from the reflected
shortwave radiation from the surrounding terrain (weighted by
the terrain-obscured fraction: (1-SVF)). An equivalent
backscattered term (i.e. Equation (3.6.4)) could then be
constructed. The SVF map can be computed using the MODWET function: compute_shade_lookup_table_and_SVF.
E XAMPLE 3.6.2
For the case shown in Example 3.6.1, how would
the incident surface radiation change if the
surface had: a) a slope of 10 degrees and an
aspect of 0 degrees (i.e. due North) and b) a
slope of 10 degrees and an aspect of 180 degrees
(i.e. due South)? In both cases assume there is
no shade and the sky view factor is equal to 1.0.

E XAMPLE 3.6.2 ( CONTINUED )


a) Note that at solar noon, the solar azimuth angle is
180 degrees (i.e. due South). The cosine of the local
illumination (zenith) angle for this case would be:

cos s = cos(10.6 )cos(10 ) + sin(10.6 )sin(10 )cos(180 0 )


= 0.936 s = cos 1(0.936) = 0.360 rad = 20.6
So for this case the topography changes the solar zenith
angle from 10.6 to 20.6 degrees. The direct beam flux is
then given by:

s,dir

" cos(20.6 )
1 %
= (0.6)(1300.5 W m )$(1)
'

# cos(10.6 ) cos(10 )&


-2

= (780.3 W m -2 )[0.967] = 755 W m -2


so that the direct beam flux is 97% of what it was for
the horizontal plane. Since the sky view factor for this
case is 1.0, the diuse flux will remain the same. In the
more general case where SVF < 1 the diuse flux may
be reduced or increased depending on the reflected
radiation from the surrounding terrain.
b) For the second case, the cosine of the local
illumination (zenith) angle for this case would be:
cos s = cos(10.6 )cos(10 ) + sin(10.6 )sin(10 )cos(180 180 )
= 0.999 s = cos 1(0.999) = 0.010 rad = 0.6

87

E XAMPLE 3.6.2 ( CONTINUED )


Note that because the slope is facing due south the local
illumination is exactly equal to the dierence between
the solar zenith angle and the slope angle. If the slope
had been 10.6 degrees the local zenith angle would be 0.
The direct beam flux is then given by:

s,dir

" cos(0.6 )
1 %
= (0.6)(1300.5 W m )$(1)
'

cos(10.6
)
cos(10
)
#
&
-2

= (780.3 W m -2 )[1.033] = 806 W m -2


So for this case the horizontal flux is actually increased
by 3%. This is because the sloped surface that is facing
southward is more directly oriented toward the sun. Of
particular importance is that the two surfaces of the
same slope but opposite aspects have a dierence of over
50 W m-2 of direct beam flux. This dierence will
propagate to other surface hydrologic fluxes (i.e.
evaporation, snowmelt, etc.) which will lead to
heterogeneity in surface states (soil moisture, snow water
equivalent, etc.). Again, the diuse flux will be the same
as for the horizontal case if SVF = 1.

88

Modeling Longwave Fluxes


at the Surface

At the surface there are two relevant longwave fluxes:
that incident at the surface from above (i.e. emitted
downwelling radiation by the atmosphere) and upwelling/
outgoing radiation emitted by the surface. The downwelling
longwave radiation can be written conceptually as:

Rl = f ([H 2O],[CO2 ],...),T(z), clouds

(3.7.2)

Rl = aTa4

S ECTION 7

(3.7.1)

which is to say that the downwelling flux depends on the


atmospheric emissivity (which itself depends on the
concentrations of radiatively active gases), the temperature
profile, and the characteristics of any clouds present. To fully
model these fluxes at a given point requires the use of a
radiative transfer model and inputs of all the necessary
profiles. Because the atmosphere is rather absorptive of
longwave radiation, most of the downwelling radiation
reaching the surface is that emitted from the lower layers of
the atmosphere. So semi-empirical models are often used to
model the downwelling flux in terms of more readily available
near-surface meteorological measurements. In particular, the
clear-sky flux, using the physically-based bulk flux equation
introduced above, can be written as:

a effective atmospheric emissivity

Ta effective atmospheric temperature


While the atmospheric emissivity is dependent on all
greenhouse gas concentrations, the most variable is water
vapor, so several empirical models have been developed to
estimate the emissivity based on humidity measurements.
Examples include that by Idso (1981):

a = 0.74 + 0.0049ea

(3.7.3)

ea reference-level vapor pressure [mb]

by Satterlund (1979):
(T 2016)

a = 1.08(1 exp(ea a

));

[ea ] = mb;

[Ta ] = K

(3.7.4)

by Brutsaert (1975):
0.14

!e $
a = 1.24 ## a &&
"Ta %

[ea ] = mb;

[Ta ] = K

(3.7.5)

and by Brunt (1932):

a = 0.605 + 0.048ea0.5 ;

[ea ] = mb

(3.7.6)

along with many others. Again, it is important to recognize


that empirical equations are not necessarily dimensionally
consistent (i.e. it is important to use specified units!) and are
most valid for the locations and conditions in which they were
developed.
89


The above model is for clear-sky downwelling radiation.
To account for clouds, a similar approach to that used in the
shortwave case can be used, i.e.:

Rl = flcaTa4

(3.7.7)

where flc represents an empirical factor. An important note is


that while clouds attenuate shortwave radiation, they actually
enhance longwave radiation. This is because clouds
significantly increase the concentration of water in the
overlying atmosphere which increases the eective emissivity.
One example of such an empirical augmentation factor is that
by Kustas et al. (1994):

flc = (1 + 0.22C 2 )

(3.7.8)

which indicates no augmentation for zero cloud cover (C = 0)


and an increase in longwave by 22% for full cloud cover (C =
1). This particular equation makes no eort to model dierent
types of clouds which will undoubtedly introduce some error.
For example, two conditions might exist with full cloud cover,
but one a very thin cirrus cloud at considerable height above
the surface and the other a thick fog cloud in contact with the
surface. Such a case would be expected to yield significantly
dierent cloud impacts on longwave flux.

For the upwelling longwave flux from a surface, a similar
bulk emission model is often used, i.e.:

Rl = sTs4

The surface upwelling radiation is generally simpler than the


downwelling radiation in that the emissivity is essentially
constant (since it depends on land surface type) or varies at
significantly longer time scales than the atmospheric
emissivity. Typically tabulated values are used depending on
the surface type (Table 3.1).

Strictly speaking, the longwave fluxes described above
are for a flat horizontal plane. While not generally as
important as with shortwave fluxes, the longwave fluxes can
also be impacted by terrain. The primary impact is due to
variable sky view factor (SVF) defined in the previous section.
The implication of having SVF less than 1.0 is that some of
the area contributing to downwelling longwave flux at a given
location is due to the emitted radiation of the surrounding
surfaces. This is often modeled using a weighted average of
the flux from the atmosphere itself and the surrounding
terrain, i.e for clear-sky flux.:

Rl = (SVF)aTa4 + (1 SVF) sTs4

(3.7.10)

where the brackets on the second term simply denote the


average emitted longwave flux from the surrounding terrain
(which may dier from the locally emitted flux). In the case of
flat terrain (SVF = 1) this reduces to Equation (3.7.2). This
can also be modified to account for cloud eects as described
above.

(3.7.9)

s surface broadband emissivity

Ts surface temperature

90

E XAMPLE 3.7.1

E XAMPLE 3.7.1 ( CONTINUED )

For the air with characteristics described in


Example 2.4.1, estimate the clear-sky
downwelling longwave radiation using the Idso,
Brutsaert, Satterlund and Brunt emissivity
models. How do the estimates dier and,
qualitatively, how would they change in cloudy
conditions? Also compute the upwelling longwave
radiation from the surface assuming it is covered
by a short-grass surface with a surface
temperature of 28 degrees Celsius.

a = 0.605 + 0.048(15 mb)0.5 = 0.79

First, note that all of emissivity models are empirical


and therefore not dimensionally consistent. So care needs
to be taken to use the specified units. Based on the Idso
model, the atmospheric emissivity for these conditions is:

a = 0.74 + 0.0049(15 mb) = 0.81


The Brutsaert model predicts the following emissivity:
0.14

! 15 mb $
a = 1.24 #
&
" 293.15 K %

= 0.81

The Satterlund model predicts:

a = 1.08(1 exp((15 mb)(293.15K 2016) )) = 0.84


while the Brunt model predicts:

so that for these conditions the estimates are reasonably


similar with a percent dierence of less than about 5%
across the models. This will directly propagate to the
clear-sky longwave radiation which for each of the
respective models are:

Rl = (0.81)(5.67 10 8 )(293.15 K)4 = 339 W m -2


Rl = (0.81)(5.67 10 8 )(293.15 K)4 = 339 W m -2
Rl = (0.84)(5.67 10 8 )(293.15 K)4 = 352 W m -2
Rl = (0.79)(5.67 10 8 )(293.15 K)4 = 331 W m -2
so that across all models there is a dierence of up to 17
W m-2. If conditions were cloudy the longwave fluxes
would increase due to increased emission by the clouds.
The upwelling longwave flux depends on the surface
emissivity (0.95 from Table 3.1):

Rl = (0.95)(5.67 10 8 )(301.15 K)4 = 443 W m -2


Note that the upwelling longwave flux is greater than the
downwelling flux. This is very often the case because
even when the overlying atmosphere is warmer than the

91

E XAMPLE 3.7.1 ( CONTINUED )


surface (i.e. at night or in winter), the surface emissivity
is often much larger than the eective atmospheric
emissivity.

92

(3.8.2)

S ECTION 8

Rl Rl

Net Radiation at the Surface

where in most cases the outgoing longwave flux is larger than


the incoming flux since the surface emissivity is generally
higher than the atmospheric emissivity and the temperatures
are comparable. Hence the net longwave radiation is often
negative.


The previous two sections provided physical descriptions
and relatively simple models for the key shortwave and
longwave fluxes occurring at the land surface. While it is
convenient to model each flux individually based on the
dierent physical relationships, the key radiative input is the
net (or integrated) radiation absorbed by the surface. It is this
net flux that will drive hydrologic processes at the surface.
The net radiation can be written simply as:

Rn net radiation net shortwave + net longwave


The net shortwave radiation is the downwelling shortwave (as
described by Equation (3.6.9) for horizontal terrain) minus the
upwelling shortwave radiation. By definition there is no
emission of shortwave by the land surface, so the only source
of upwelling shortwave is simply the amount of incident
radiation that is reflected. The reflectivity is exactly defined
by the albedo so that the net shortwave is given by:

Rs Rs = Rs Rs = Rs(1 )

(3.8.1)

which simply states that whatever fraction of incident


shortwave radiation is not reflected is being absorbed. The net
longwave radiation is the downwelling minus upwelling
longwave fluxes:


Putting the two together we can write the net radiation
at the surface as:

Rn (x,y,t) = Rs(1 ) + Rl Rl

(3.8.3)

where the spatial and temporal dependence is shown for


emphasis. The spatial/temporal variability comes through the
various factors that determine the individual fluxes (albedo,
emissivity, near-surface meteorological data, etc.). Since net
radiation is defined as the absorbed radiation by the surface, a
positive quantity denotes a net gain of energy input (that will
cause the surface to warm) and a negative quantity denotes a
net loss of energy (that will cause the surface to cool).

In terms of temporal variability, it is important to keep
in mind two key periodicities embedded in Equation (3.8.3).
The seasonal cycle is driven by the Earths orbit around the
sun, which yields seasonal variability in the TOA shortwave
flux as a function of latitude (Figure 3.11). This seasonal cycle
generally then imprints itself on the meteorological and
surface states that control the longwave fluxes. Diurnal
variability also plays a key role. The incident solar radiation is
zero at night and varies throughout the day with a peak at
solar noon. The incoming longwave is determined by air
93

temperature, humidity, and clouds, which can also have a


strong diurnal cycle. Movie 3.2 shows an example of the
diurnal evolution of downwelling shortwave and longwave
fluxes over the central U.S. during the summer (on a cloudy
day). The diurnal cycle is clearly seen and is modulated by
the presence of clouds, which strongly control the patterns.
These patterns will directly propagate to the surface net
radiation. During the day and especially during summer, the
shortwave fluxes generally dominate the net radiation and
since the longwave fluxes are generally comparable in

magnitude, provide a positive energy input to the surface. At


night (and at high latitudes during winter) radiative cooling
generally occurs due to the absence (or limited amount) of
shortwave radiation inputs.

M OVIE 3.2 Illustration of the diurnal evolution of downwel-

ling shortwave and longwave surface radiation in the presence


of clouds over a region in the Southern Great Plains of the U.S.
(Bart Forman, personal communication).
94

S ECTION 9

MOD-WET Codes

Relevant functions based on concepts introduced in
this chapter include:
Calculation of diuse shortwave scattering coecient:

diffuse_sw_scattering_coefficient.m
Calculation of direct beam shortwave transmissivity:

direct_sw_transmissivity.m
Computation of incoming solar radiation over complex
terrain:

disaggregate_SW.m
Conversion between easting/northing UTM vectors and
lat/lon vectors:

easting_northing_to_lat_lon.m
Slope and aspect determination for a given DEM:

generate_slope_and_aspect_from_DEM.m
Idso model for atmospheric emissivity:

idso_emiss.m
Satterlund model for atmospheric emissivity:

satterlund_emiss.m

Brutsaert model for atmospheric emissivity:



brutsaert_emiss.m
Brunt model for atmospheric emissivity:

brunt_emiss.m
Computation of clear-sky downwelling and upwelling
longwave fluxes:

longwave_flux.m
Calculation of longwave augmentation due to cloud fraction:

lw_cloud_augmentation_factor.m
Estimation of atmospheric optical depth:

optical_depth.m
Calculation of solar geometry (zenith and azimuth):

solar_geometry.m
Calculation of shortwave attenuation due to cloud fraction:

sw_cloud_attenuation_factor.m
Calculation of top-of-atmosphere solar radiation:

TOA_incoming_solar.m
Calculation of shading due to topography:

topo_shade_calc.m
Generation of a shade lookup table and SVF map from a
DEM:
95

compute_shade_lookup_table_and_SVF.m

Conversion between lat/lon and UTM coordinates:



utm2deg.m

96

S ECTION 10

Conceptual Questions
1. Name the respective sources of shortwave and longwave
radiation fluxes. What part of the spectrum is covered by
each? At what wavelength does the peak for each occur?
2. According to the Planck function, name the physical
variable controls the magnitude (and distribution) of
emitted (blackbody) radiation. Name the additional
physical parameter that describes how close an actual body
is to a blackbody.
3. Name at least two atmospheric radiatively active gases that
absorb solar radiation. Name at least two atmospheric
radiatively active gases that absorb longwave radiation.
4. Define in words what surface albedo represents. What land
surface types generally have the smallest and largest
broadband albedo and what are their typical values?
5. Name the variables are needed to determine the TOA
incoming shortwave flux.
6. Define in words what the declination angle represents.
7. What is the maximum possible value of incoming shortwave
radiation at the TOA? Where on the globe will this
generally occur?

8. Describe in words the processes that attenuate downwelling


solar radiation between the TOA and surface.
9. Name the three components that make up the incoming
shortwave radiation at the surface. At midday during the
summer, which component would you expect to be the
largest? Just before sunset, which component would expect
to be largest? Explain.
10. Describe in words the meaning of aspect.
11. In the northern hemisphere (above 23.5 degrees North),
which aspect (north or south facing) will generally get more
incident solar radiation? Explain. How do things change in
the Southern Hemisphere? Explain.
12. What atmospheric variable is primarily responsible for
determining the eective atmospheric emissivity? What
additional variable is primarily responsible for surface
downwelling longwave flux?
13. What land surface types generally have the smallest and
largest broadband emissivities and what are their typical
values?
14. Name the four components that make up the net radiation
at the surface. What surface and/or atmospheric variables
and parameters are generally needed to compute net
radiation at a given location/time.
15. When/where will net radiation generally be positive?
negative?
97

S ECTION 11

Sample Problems
Problem 3.1. In this problem, you are asked to examine the
influence of the atmosphere on the radiative flux incident at
the earths surface. Incoming solar radiation in the shortwavelength spectra (shortwave) at the top of the atmosphere
(TOA) is partially absorbed and scattered (reflected) by
atmospheric molecules and particulates. As a result, only a
fraction of the TOA flux actually reaches the surface. Answer
the following questions using the equations provided in
Sections 5 and 6.
a) Compute declination angle, zenith angle, and incoming
top-of-atmosphere (TOA) solar radiative flux for solar noon
on December 21st, 2011 (winter solstice), and June 21st, 2012
(summer solstice) in Hilo, Hawaii (19.73 deg. N). For this
latitude and days of the year (and time of day), is the sun in
the southern sky, northern sky, or directly overhead? [Hint:
You can answer this by comparing the latitude to the
declination angle]. Is your answer generally true throughout
the year? Justify your answer.
b) What is the predicted shortwave direct beam
transmissivity of the atmosphere on June 21st based on the
precipitable water value you estimated in problem #1 part c)?
Assume attenuation due to dust Based on your answer,
approximately what fraction of the incoming direct beam solar

radiation is attenuated by the atmosphere? Qualitatively, how


would this change if the amount of water vapor increased? So
if all else is equal, would you expect more direct beam solar
radiation to reach the surface in an area with a dry climate or
one with a humid climate?
c) What is the predicted scattering coecient of the
atmosphere on June 21st based on the precipitable water
value you estimated in the sample problem in the last
chapter? How do the transmissivity and scattering coecients
compare, i.e. which is larger/smaller, etc.? Qualitatively, how
would the scattering coecient change if the amount of water
vapor (i.e. cloud cover) increased?
d) Using Equation (3.6.5) and the coecients and variables
computed above, compute the incoming shortwave radiation
at the surface in Hilo, at noon on June 21st. Assume a surface
albedo of 0.24.
Problem 3.2.
a) Compute the top of atmosphere (TOA) solar radiation over
a location in the Sierra Nevada mountains (latitude of 36.535
degrees) for 9am, noon, and 3pm (local solar time) on October
17th (Day of Year 290). Note: The TOA flux we use (given by
the equation) is that perpendicular to the TOA. As part of
your solution, clearly indicate the solar zenith angle
corresponding to each time. What fraction of the solar
constant is coming in perpendicular to the TOA at these
times? Explain (using physical arguments) why the fractions
are less than 1.0. How does solar zenith angle vary throughout
the day?
98

For each of the three times listed above, answer the following
questions. Note: In reality the atmospheric parameters will
vary throughout the day due to the solar flux passing through
more or less atmosphere. Below we will assume they are
constant for simplicity.
b) Assuming that a transmissivity of 0.52 is representative for
this day over the Sierras, what is the magnitude of the direct
beam flux incident on a horizontal plane at the surface?
c) Assuming a scattering coecient of 0.185, what is the
magnitude of the diuse flux incident on a horizontal plane at
the surface?
d) Assuming the average surface albedo is 0.25 in this region,
what is the magnitude of the backscattered shortwave flux
incident on a horizontal plane at the surface?
e) What is the total shortwave radiation incident on a
horizontal plane at the surface? How do the three components
compare in magnitude?
f ) What is the net shortwave radiation absorbed by the
surface?
g) Describe conceptually what locations (latitudes) on the
globe would have increased solar radiation on south facing
slopes vs. those that would have increased solar radiation on
north facing slopes. [Hint: This is related to the declination
angle.]

Problem 3.3. The following information are measured over a


vegetated surface during a clear sunny day:
Net solar radiation = 510 Wm-2.
Air Temperature at 2 meters: 22.7C
Air relative humidity at 2 meters: 80.0%
Ground surface temperature = 25.5C
Air pressure = 980 millibars
a) What is the outgoing solar radiation at the surface?
Assume a surface albedo of 15%.
b) What is the outgoing longwave radiation from the surface
(assume a surface thermal emissivity of 0.96)?
c) What is the incoming atmospheric longwave radiation? Use
the Idso model for atmospheric emissivity.
d) What is the net radiation at the surface?
e) This radiation must be dissipated at the surface. Assuming
half of this net radiation goes into vaporization of water at
the surface, what would the evapotranspiration rate be in
mm/day? If the latent heat later condenses in the
atmosphere, how much energy (in W m-2) will be released/
consumed? Specify whether the condensation corresponds to a
release or consumption of energy.
Problem 3.4. In order to plan the days irrigation rate, a
farmer needs to estimate the maximum potential water loss
rate by estimating the absorbed radiative energy at the
99

surface. Over the irrigated field at noon, the ground


temperature is generally 41C and the air temperature is
33C. The air dew point temperature is 16C. Assume a
surface albedo of 30% and surface emissivity of 0.95. Based on
location and day (Day 186), the cosine of the solar zenith
angle at noon is estimated to be 0.974. Assume the
atmospheric direct beam shortwave transmissivity at this time
is 0.6, the diuse shortwave scattering coecient is 0.1 and
the sky is cloud-free.
a) Determine the net radiation at the surface.
b) Qualitatively, how would cloud presence impact your
estimate? Be specific.
c) Assuming 65% of the net radiation goes into
evapotranspiration (i.e. latent heat flux) at the surface, how
much water should the farmer apply (in mm/day) to balance
the evapotranspiration flux?
Problem 3.5. Some estimates of global warming contend
that surface air temperatures will increase by 3C over the
next 100 years. What would the change in incoming thermal
radiation at the surface be under this global warming scenario
if the nominal air temperature before the warming eect is
20C and the relative humidity of the surface air remains at
60%?
Problem 3.6. In a particular desert region observed by a
satellite, a localized thunderstorm covering 15% of the area
causes the soil in this area to become very moist (relative
saturation s =1). The dry area in the rest of region has s = 0.

After the storm, the contrast between dry/moist areas is


detectable from a satellite because soil moisture can have a
significant impact on surface albedo. Suppose a simple
equation for the dependence of albedo on soil relative
saturation (s) is given by:

= 0.3 0.1s
Simultaneous satellite measurements indicate that the areaaveraged outgoing longwave radiation from the entire region is
536 W m-2. A ground based measurement indicates that the
dry area surface temperature is 315 K. Incoming solar
radiation at this time is equal to 900 W m-2 and average air
temperature and relative humidity over the entire region is
312 K and 35% respectively. Assume a representative value of
0.97 can be used for the land surface emissivity in this area.
a) What is the surface temperature of the moist area?
b) What is the net radiation over the dry and moist areas?
Explain why they are dierent.
c) What is the area-averaged net radiation?
Problem 3.7. To a large degree, the climate of Earth is
controlled by radiative fluxes. Radiative equilibrium consists
of the balance of net solar (shortwave) radiation, which is an
external input to the system, and net terrestrial (longwave)
fluxes, which generally depend on the Earths surface and its
atmosphere. A schematic picture of the global radiative energy
balance is shown in the figure below. Radiative equilibrium
refers to the condition where the net radiative energy is equal
100

(i.e. in W), not the flux densities (which are in W m-2). In


other words, you will need to consider the area over which
each respective flux occurs. As indicated in the figure shown
above, an area equal to the projected cross-sectional area of
the Earth intercepts the incoming solar radiation, while the
outgoing longwave flux is emitted across the full surface area
of the Earth. The reflected shortwave radiation occurs over
the same area as the intercepted shortwave.

to zero (i.e. total energy in minus total energy out equals


zero).
Related to this, and of particular relevance, is the greenhouse
eect which is generally used to refer to the impact of
(radiatively active) atmospheric gases on the Earths
equilibrium temperature. On the global scale there is generally
only one incoming flux (solar radiation) and two outgoing
fluxes (outgoing (reflected) solar radiation, and emitted
longwave radiation). Here you will examine the impact of the
atmosphere on the radiative equilibrium.
a) First suppose there was no atmosphere overlying the
surface of the Earth. For this case, write the radiative
equilibrium equation at the surface in terms of the solar
constant, the Earths surface: temperature, albedo, and
emissivity, and the radius of the Earth. The radiative
equilibrium equation should consider the total energy fluxes

b) Using the above equation, derive an expression for the


Earths surface temperature. Your expression should be in
terms of the variables listed above and (and the StefanBoltzmann constant). Given the known value for the solar
constant, and assuming typical values for the Earths surface
albedo and emissivity are given by: 0.3 and 0.97 respectively,
what would the predicted surface temperature be for this noatmosphere case?. How does this compare to the observed
(global average) value of ~290K that exists as a result of
having an atmosphere?
Next, you will consider the impact of having atmosphere with
greenhouse gases. For simplicity, suppose the atmosphere can
be conceptualized by one layer characterized by a single
atmospheric temperature and atmospheric emissivity (both of
which are generally dierent then the surface values). Also, for
simplicity, assume the atmosphere is transparent to shortwave
radiation (i.e. shortwave transmissivity equal to 1.0), but
absorbs longwave radiation (with negligible longwave
scattering). Recall according to Kirchos Law that the
longwave absorptivity is equal to the longwave emissivity.
101

c) Based on this simplified representation of the atmosphere


layer, write expressions for the upwelling/downwelling
atmospheric longwave radiation atmospheric temperature/
emissivity, and the Stefan-Boltzmann constant) and for the
atmospheric longwave transmissivity (in terms of atmospheric
emissivity). Assume that the downwelling and upwelling
longwave radiation by the atmosphere are the same.
d) Based on the longwave transmissivity of the atmosphere,
what is the expression for the longwave radiation emitted by
the Earths surface that makes its way to the top of the
atmosphere?
e) With the presence of the atmosphere, there are radiative
equilibrium equations at both the TOA and the Earths
surface. Write both equations in terms of the variables defined
above. You will again need to consider the respective areas
across which the fluxes are intercepted/emitted. You can
assume the atmospheric surface is the same as the earth
surface.
f ) Combining the surface and TOA equations you can
eliminate the dependence on atmospheric temperature. Do so
to derive an expression for the Earths surface temperature in
terms of the other parameters in the problem. What is the
predicted surface temperature for an atmosphere with an
emissivity (absorptivity) equal to 0? Equal to 1.0? What must
the atmospheric emissivity equal to match the observed
globally-averaged surface temperature mentioned above?

102

Chapter 4

Atmospheric
Circulation

S ECTION 1

Learning Objectives
By the time you finish this chapter you should be able to:
1. Sketch a conceptual picture of the annual average net
radiation absorbed at the TOA by the Earth system
2. Describe how and why the net radiation distribution
causes atmospheric motions
3. Sketch a conceptual picture of the general circulation of
the atmosphere
4. List the key characteristics of the general circulation and
discuss its impact on climate patterns (including
precipitation patterns)
5. Identify how high/low pressure cells are connected to
circulation and vertical motion
6. Identify which circulation feature is primarily responsible
for equator to mid-latitude energy transport
7. Identify which circulation feature is primarily responsible
for mid-latitude to pole energy transport

104

S ECTION 2

Global Distribution of Topof-Atmosphere Net Radiation



This chapter focuses on the key aspects of atmospheric
circulation as they relate to other hydrologic concepts covered
in the book. For a more detailed treatment of atmospheric
dynamics the reader is referred to Marshall and Plumb (2007).

The net radiation at the TOA is the key driver of the
global atmospheric circulation and hydrologic cycle. Based on
the definitions and notation from the previous chapter, the
TOA net radiation can be written as:

Rn 0 (x,y) = Rs 0 (1 0 ) Rl0

(4.2.1)

0 planetary albedo

Rl0 outgoing longwave flux at TOA

where the spatial dependence is shown for emphasis. The


planetary albedo is the bulk reflectivity and is a composite of
the atmospheric reflectivity and surface reflectivity impacts on
outgoing shortwave flux. Note that there is no incoming
longwave flux, as there is at the surface, since there is no
atmosphere above the TOA to emit downwelling radiation.
The upwelling longwave flux is a composite of that emitted

M OVIE 4.1 Illustration of monthly evolution of TOA net radiation (from earthobservatory.nasa.gov/GlobalMaps/
view.php?d1=CERES_NETFLUX_M).

upward by the atmosphere as well as that emitted by the


surface and transmitted through the atmosphere. The
animation in Movie 4.1 shows maps of monthly net radiation
highlighting the spatial patterns and seasonality. Note both
positive and negative values representing a net absorption and
loss respectively.

It is instructive to average the above equation in various
ways to gain insight into the overall radiation budget and its
implications. If the above equation is globally averaged one
expects the basic conceptual picture shown in Figure 4.1. The
incoming shortwave flux is approximately 340 W m-2, which is
simply the solar constant multiplied by the ratio of the crosssectional area of the disk intercepting the incoming solar beam
to the Earths surface area, i.e.:
105

contribute to the balance, namely downwelling longwave


radiation (labeled backradiation in Figure 4.1; already
discussed in Chapter 4) and surface latent and sensible heat
fluxes (evaporation and convection in Figure 4.1; to be
discussed in Chapters 6 and 8).

Perhaps of more relevance is the insight obtained when
averaging the net radiation across all longitudes at a given
latitude band (i.e. zonally-averaged) to get the annual mean
net radiation at the TOA as a function of latitude. Figure 4.2

F IGURE 4.1 Long-term average global radiative energy balance.

" R2 %
2
E
'
1367 Wm $$
=
342
W
m
2 '
# 4 RE &
2

where RE is the radius of the Earth. Of this incoming


shortwave flux, approximately 29% (approximately 99 Wm-2)
gets reflected back to space. When considering the global
average, the TOA net radiation must equal zero for the planet
to be in radiative equilibrium, which is shown in Figure 4.1
where the outgoing longwave flux is equal to 71% of the
incoming (approximately 342 W m-2). Note the net energy
must also equal zero at the surface, but additional fluxes

F IGURE 4.2 Zonally averaged incoming, outgoing, and net


radiative fluxes over the globe.

106

shows the incoming and outgoing radiation fluxes, which when


added together provide the net radiation. What is seen is a
relatively symmetric pattern around the equator, with a
surplus of energy being input to the system in the tropics (i.e.
between latitudes of +/- 30) and a deficit of energy from
midlatitudes to the poles. In other words, the net radiation is
positive near the equator and negative near the poles. This
imbalance is a result of most of the solar input occurring near
the equator and radiative cooling occurring at the poles. The
key question is what does this imply?

For the planet to be in overall equilibrium the
implication is that there must be an underlying set of
mechanisms for the net transport of heat poleward from the
equator. To convince yourself of this, imagine the atmosphere
as three boxes isolated from each other, one centered on the
equator where the surplus exists and one in each hemisphere
where the deficit exists. If the radiation budget were the whole
story, the air in the box with a surplus would continuously
warm and the air in the two boxes with a deficit would
continuously cool. This is known not to be the case. While
heat between boxes could be transferred via conduction, this
is a relatively inecient process. When fluids are involved, a
much more ecient mechanism is via heat carried by fluid
motions (and as we will discuss later: latent heating associated
with vapor transport/condensation). In other words, warm
tropical air must be moved poleward and cool polar air must
be moved equator-ward.

F IGURE 4.3 Inferred total latitudinal heat transport resulting

from global net radiation distribution along with estimated contribution by the atmosphere and ocean circulation. PW = 1015
W (used with permission from Marshall and Plumb, 2007).

the total transport is shown in Figure 4.3. Some of the


mechanisms for the transport will be discussed in more detail
below in following sections, but they are driven principally by
atmospheric motions and to a lesser extent by oceanic
circulation.


The heat flux necessary to yield equilibrium can be
inferred from the imbalance in net radiation. An estimate of
107

S ECTION 3

Atmospheric Motions
Driven by Latitudinal Energy Imbalance

Given the implied heat transport discussed in the
previous section, the primary question is how does the
atmospheric transport work? Here we will focus primarily on a
qualitative description of the key characteristics of the
atmospheric motion, i.e. the so-called general circulation. It
is important to note that the general circulation we will be
discussing is representative of the long-term average. Hence,
the general circulation is responsible for much of the
climatology on Earth, while individual weather events
(including extremes) can be quite dierent than the mean.
The instantaneous state and motion of the atmosphere can
dier markedly from the general circulation.

The primary impact of the net radiation distribution
shown in Figure 4.2 is that, within the troposphere, warmer
air exists in the tropics relative to the poles (Figure 4.4).
Thermal expansion of the tropical air columns (relative to
those at the poles) leads to a sloping pressure surface aloft
(also seen in the temperature isotherms) which would be
expected to drive air toward the poles aloft. Additionally, air
near the surface in the tropics would be expected to be
warmest, which due to buoyancy would generate an uplift

F IGURE 4.4 Zonally-averaged distribution of atmospheric

temperature as a function of height. Red arrows indicate expected vertical motions at equator and poles.

(warm air rises) with cold air aloft at the top of the
troposphere generating sinking motion at the poles. If these
were the only factors, they would imply a single circulation
cell as shown in Figure 4.5, with surface flows from pole to
equator, upward flow at the equator, return flow aloft, and
downward flow near the poles. This simple convection cell
model would have the desired trait of transporting warm air
toward the pole and cool air to the equator, eectively
redistributing the net radiation imbalance. In fact, early
conceptual models of the Earths atmospheric flow proposed
this model (Marshall and Plumb, 2007). However, an
108

apparent force associated with the rotating reference frame


that tends to deflect air parcels to the right in the Northern
Hemisphere (and to the left in the Southern Hemisphere). As
a result, wind moving toward the equator would be deflected
creating so called easterly surface winds (i.e. from the east
moving westward) in the polar and equatorial regions.
Another consequence of rotation is that air parcels moving
poleward are closer to the axis of rotation and tend to

F IGURE 4.5 Schematic of expected single-cell thermallydriven atmospheric general circulation on a non-rotating
planet. (from http://www.ux1.eiu.edu/~cfjps/1400/circulation.html).

important factor is neglected in this model, namely the


rotation of the Earth, which brings angular momentum eects
into the picture.

The impact of angular momentum, and other factors
related to the Earths rotation, introduces a more complicated
general circulation. One of these key factors is the Coriolis
force (discussed in more detail in Section 5), which is an

F IGURE 4.6 Schematic of expected three-cell thermally-

driven atmospheric general circulation on a rotating planet.


(from http://www.ux1.eiu.edu/~cfjps/1400/circulation.html).

109

therefore accelerate in order to attempt to conserve


momentum. The result is a general circulation that features
three overturning circulation cells of varying intensity (rather
than just one) as shown in Figures 4.6 and 4.7. The three cells
are referred to as the Hadley (in the tropics on either side of
the equator), the Ferrel (at midlatitudes) and the Polar (near
the poles) cells. These types of figures attempt to
simultaneously show the circulations occurring in cross-section
in the atmosphere as well as along the surface.

The Hadley cell consists of warm air rising from the
surface at the equator, moving poleward aloft until around
30N/S, and descending toward the surface before returning
along the surface. The upwelling limb of the Hadley cell near
the equator is the primary driver of persistent tropical clouds

F IGURE 4.7 Cross-section of northern hemisphere circula-

tion superimposed on net radiation distribution. Arrows on the


surface indicate direction of surface winds.

often referred to as the Intertropical Convergence Zone


(ITCZ). The Ferrel cell has descending air at latitudes around
30N/S and poleward movement along the surface before
rising around a latitude of 60N/S and returning toward the
midlatitudes aloft. Finally, the Polar cell has the same sense
of circulation as the Hadley cell, but exists between the poles
and latitudes around 60N/S. Movement of air from the
surface to aloft (i.e. at the equator and poles) is typically
associated with lower than average surface atmospheric
pressure, while regions of air moving toward the surface from
aloft are typically associated with higher than average surface
pressure. These so-called lows and highs tend to be
persistent features of the general circulation as seen in Figure

F IGURE 4.8 Mean annual surface pressure distribution showing areas of typically high and low pressure (used with permission
from Marshall and Plumb, 2007).

110

4.8. These areas of high/low pressure are directly associated


with certain climate zones as will be discussed below.

It is very important to note that not all three of the
circulation cells are of equal importance or strength. The
Hadley cell is far and away the most persistent and influential
in terms of the energy transport described above. Evidence of
the circulation cells, with the prominence of the Hadley cell is
shown in Figure 4.9. As can be seen, the Hadley cell in each
hemisphere is consistent with the net transport of energy
poleward (albeit only to the midlatitudes). In fact, the Ferrel
and Polar cells are much less important in terms of energy

F IGURE 4.9 Cross-section of the mean annual meridional

overturning motion showing the clear presence of the Hadley


cells and to a lesser extent the Ferrel cells (used with permission
from Marshall and Plumb, 2007).

transport. Another mechanism is chiefly responsible for the


additional transport between the midlatitudes and the poles
as will be elaborated on below.

The other key features to note in Figures 4.6 and 4.7 are
the surface winds. Due to the Coriolis force, the winds in the
surface branch of the Hadley cell are deflected, becoming
easterly winds (or easterlies, i.e. from the east). In the
tropics, the easterlies are often called trade winds as their
persistent nature was taken advantage of to construct trade
routes used by sailing ships. As part of the Ferrel cell, the
surface winds are deflected, becoming westerlies. It is these
winds that form the onshore flow seen in the western U.S.
Finally the surface winds associated with the Polar cell consist
of easterlies. Another important point has to do with the
relative speed of the winds. To conserve its angular
momentum, an air parcel at higher latitudes would generally
need to travel at a faster speed (since it is closer to the axis of
rotation) than it does at lower latitudes. Hence the winds at
higher latitudes tend to be faster than the trade winds, which
are persistent but relatively mild.

As mentioned above, the Hadley cell is the chief
mechanism for transport of energy between the equator and
midlatitudes. This leaves the question: What mechanism is
chiefly responsible for transport between midlatitudes and the
poles? Rather than an overturning cell, the transport
generally occurs via eddy mixing in the horizontal plane of the
atmosphere. Due to stark temperature dierences, the
midlatitudes are often characterized by cold air masses from
the poles interacting with warmer air masses at midlatitudes
111

and/or the tropics (Figure 4.10). These colliding air masses of

F IGURE 4.10 Schematic showing frontal systems at midlatitudes that result in mixing of cold and warm air via horizontal
eddies and frontal uplift (used with permission from Marshall and
Plumb, 2007).

diering thermal characteristics are usually referred to as


fronts. The result of these interactions are two-fold: i) air
masses tend to circulate around high/low pressure anomalies
forming eddies that stir cold air equator-ward and warm air
poleward, and ii) warm air is driven upward over the cold air
due to buoyancy dierences (Figure 4.10). So while this
mechanism does not involve an overturning circulation cell, it
has a similar eect. Warmer air is transported poleward and
upward into the atmosphere while the opposite is true of
cooler air masses. Additionally, these horizontal eddies
(storms) are intermittent in nature in contrast to for example
the Hadley cell which is very persistent. The eddies tend to

M OVIE 4.2 Animation of the wind field over the Northern

Hemisphere and over North America from the NASA MERRA


dataset. Wind vectors are shown at three levels in the atmosphere: 300 mb (blue), 500 mb (white), and 850 mb (black)
(from NASA/Goddard Space Flight Center Scientific Visualization Studio;
svs.gsfc.nasa.gov/vis/a000000/a003700/a003733/index.html).
112

form and dissipate, with life cycles on the order of days to a


week. These can be seen in Movie 4.2 which shows an
example of the dynamic wind fields over the Northern
Hemisphere (based on a combination of observations and
model output). The general circulation discussed above is
essentially a time average of what is seen in the animation and
is discussed in more detail in the next section.

113

S ECTION 4

Summary of Key Characteristics of Circulation



To summarize the key aspects of the circulation
(especially those relevant to hydrologic processes) we can
disaggregate the key characteristics into those near the surface
and those higher in the atmosphere (Figure 4.11). It is useful
to remember that in this context the upper atmospheric
circulation features are mostly confined to the upper
troposphere.

A summary of the key low-level (near surface)
atmospheric circulation characteristics that exist include:
Relatively weak (i.e. typically 0-10 m/s) low-level winds (due
to friction) that exhibit a distinct spatial distribution
Convergence of easterly trade winds toward the equatorial
low pressure zone as part of the near surface branch of the
Hadley cell
Sub-tropical high pressure zones at approximately 30 N/S
as part of the downward limb of the Hadley cell
Westerlies, and more importantly, intermittent horizontal
eddies in the mid-latitudes

F IGURE 4.11 Schematic showing key general circulation features aloft (top panel) and at the surface (bottom panel) (used
with permission from Marshall and Plumb, 2007).


In the Northern hemisphere, the high surface pressure
zones are associated with anti-cyclonic (clockwise) flow, while
the low surface pressure zones are associated with cyclonic
(counter clockwise) flow (Figure 4.12). All of these features
are relatively persistent throughout the year, but can
114

Fast (i.e. 30-50 m/s) westerly winds in the subtropics (i.e.


subtropical jet) and midlatitudes (i.e. polar or midlatitude
jet)

F IGURE 4.12 Map showing persistent surface pressure field


and the resulting cyclonic and anti-cyclonic flow features.

strengthen/weaken and migrate seasonally. In particular, the


Hadley cells tend to shift following the seasonal cycle of the
sun and hence may not be centered at the equator, but rather
centered around the region of maximum solar input. Similarly,
the position of high and low pressure zones tend to shift. For
example the high o the coast of California tends to move
northward during the summer deflecting storms and
minimizing cloud/rain formation (as will be discussed later)
over the region and moves southward during the winter
allowing for the observed seasonal storm cycles.

A summary of the key upper-level atmospheric
circulation characteristics that exist include:
Simplification of the flow patterns due to the lessening of
frictional (surface) eects


These so-called jet streams result from the geostrophic
balance (discussed in more detail in Section 5) between
latitudinal pressure/temperature gradients and the Coriolis
force and generally occur in regions of strong temperature
contrast (i.e. where circulation cells meet; Figure 4.13). These
jet streams circle the globe at fast speeds and are responsible
for much of the advection of storms. The annual average zonal
wind is shown in Figure 4.14. Note that since this is an annual
average and not a snapshot in time, the feature seen near the
top of the troposphere is really indicative of both the
subtropical and midlatitude jets which migrate latitudinally
throughout the year. In particular, the polar (or midlatitude)

F IGURE 4.13 Atmospheric cross section along a longitudinal

line showing the key circulation features including areas of uplift, downdraft, overturning and jets (from theairlinepilots.com/
forumarchive/met/atmospherecirculation.jpg).

115

F IGURE 4.14 Mean annual zonally averaged plot of wind

showing large jet streams (used with permission from Marshall and
Plumb, 2007).

jet migrates significantly (Figure 4.15) and its shift in the


winter to lower latitudes is primarily responsible for storms
(i.e. just the eddies discussed previously) migrating toward
lower latitudes. Since these jet streams are often responsible
for advection of storms they are often referred to as the
storm track. It should be noted that they not only migrate
seasonally, but can vary significantly from day-to-day. This is
shown more explicitly in Movie 4.3 which shows an animation
of the atmospheric flow with an emphasis on the jet stream.
The animation shows the large spatial extent of this

F IGURE 4.15 Illustration of typical position and magnitude


of the midlatitude jet stream over the continental U.S. (from
www.ux1.eiu.edu/~cfjps/1400/circulation.html).

circulation feature which spans the entire globe. It should also


be noted that its position is quite dynamic, the magnitude
varies in time, and there are many cyclonic circulation
features that are generated by and advected by the jet stream.

Figures 4.16, 4.17, and 4.18 further summarize/
conceptualize the key aspects of the general circulation. There
are two main mechanisms responsible for the transport of
energy needed to equilibrate the latitudinal distribution of net
radiation: the Hadley overturning cell that transports energy
between the tropics and midlatitudes and horizontal eddies
116

relative to evaporation tend to occur at the equator where


water vapor is plentiful and the upwelling limb of the Hadley
cells resides (Figure 4.18). Conversely, the downwelling limb of
the Hadley cells corresponds to the location of most of the
great deserts of the world, where evaporation far exceeds
precipitation. For example, this is the primary reason, for the
dry Southern California climate despite its proximity to a
large body of water and an onshore flow.

M OVIE 4.3 Animation of the jet stream (shown in red) over

the Northern Hemisphere and over North America from the


NASA MERRA dataset (from NASA/Goddard Space Flight Center Scientific Visualization Studio;
svs.gsfc.nasa.gov/vis/a000000/a003800/a003864/).

that transport energy between midlatitudes and the poles.


These are mostly a reiteration of what has been stated above,
but it is important that these are the key take-home messages.
While features like the Ferrel and Polar cells are interesting,
they play secondary or indirect roles in the redistribution of
energy. From a hydrological perspective, knowledge of the
general circulation is important as it ends up directly
connecting to key fluxes (i.e. precipitation and evaporation) as
noted in Figure 4.18. As will be discussed in the next chapter,
areas of upward vertical air motions tend to be associated
with cloud and precipitation formation, while areas of
downward vertical motion tend to suppress cloud and
precipitation formation. Hence areas of high precipitation


A point worth reiterating is that the atmosphere is a
highly dynamic fluid and the above discussion has primarily
focused on its mean behavior. Other atmospheric features like
Monsoons, El Nino/La Nina, hurricanes/typhoons, etc. can be
as important as the general circulation, especially in localized
regions. This should be kept in mind. Movie 4.4 shows a
simulation of atmospheric dynamics in an attempt to more

F IGURE 4.16 Conceptual picture showing the key mechanism responsible for global heat and momentum transport
(used with permission from Marshall and Plumb, 2007).

117

aspect is transport of energy via latent heat fluxes. Water


vapor evaporates via the input of net radiation (occurs most
prevalently in the tropics where net radiation is high). This
energy input is then carried by the atmosphere in the form of
vapor until the vapor condenses (i.e. forms clouds). At that
point the same amount of energy that was consumed in the
vaporization process is released in the condensation process.
This is precisely why the process is called latent heating, as

F IGURE 4.17 Schematic showing the dominant features re-

sponsible for heat transport: the Hadley cell and midlatitude


eddies/frontal systems (used with permission from Marshall and Plumb,
2007).

clearly visualize this point. Clouds are shown in the grey/


white colors, while precipitation is shown in orange. Some
readily visible features of this include (among others): the
persistent Hadley circulation in the form of the ITCZ,
intermittent horizontal eddies at midlatitudes, seasonal eects,
including: the shift of the ITCZ north/south of the equator
and localized features like the Indian monsoon, which shows
oshore flow with little rain half the year and onshore flow
with heavy rain for the other half of the year.

Finally, in the above discussion, the focus has been on
movement and interchange of warm and cool air as a
mechanism for transport. This is important, but another key

F IGURE 4.18 Summary figure of atmospheric circulation

with key features and the implication of these features on precipitation and evaporation distribution with latitude (from
sonoma.edu/users/f/freidel/global/207lec2images.htm).

118

M OVIE 4.4 Animation of model output showing seasonal evolution of precipitable water (white colors) and precipitation (orange) (from vets.ucar.edu/vg/CCM3T170/index.shtml).

latent refers to hidden. This latent energy is carried by the


the moving atmospheric fluid to later be released upon
condensation. This mechanism is important largely because of
the high latent heat of vaporization of water and represents
one of the key mechanisms (driven by circulation) for heat
transport between the tropics and midlatitudes.

119

S ECTION 5

Fundamental Equations of
Atmospheric Motion

The general circulation of the atmosphere described
qualitatively in the earlier sections is the product of the
governing equations of motion for the atmosphere. The full set
of governing equations are the coupled set of equations
consisting of: 1) the equation of state, 2) conservation of
momentum, 3) conservation of air mass, 4) conservation of
water mass, 5) conservation of energy. The equation of state
was already defined in Equation (2.3.4). In addition to that
form, it can be expressed in terms of specific humidity rather
than vapor pressure as:

p = RdTv ; where Tv = T !"1 + 0.608q #$

(6.5.1)

where this form is often used because the conservation of


water vapor equations will also be expressed in terms of q. In
considering the dynamics of a fluid parcel, the key state
variable is the parcel velocity. The fluid motions are generally
defined in terms of a three-dimensional velocity vector:

V = [u,v,w]T

(6.5.2)

where the three components are given by:

u=

dx
;
dt

v=

dy
;
dt

w=

dz
dt

(6.5.3)

where the variables x, y, and z denote the coordinate position


of an air parcel and T represents the transpose operator. As
such typically u, v, and w represent west-east, south-north,
and vertical velocity components.

Before considering any of the conservation (budget)
equations in a moving fluid we first need to consider: i) local
storage changes in the variable and ii) changes caused by the
air parcel being moved through a variable field. Consider a
general property of an individual parcel:

= (x(t),y(t),z(t),t)

(6.5.4)

which states that the variable depends on space and time, but
that the spatial coordinates are also changing with time due
to position changes. The so-called total or material
derivative is given by:

d
dx dy dz
( ) =
+
+
+
dt
t x dt y dt z dt

=
+u
+v
+w
t
x
y
z

(6.5.5)

where the first term on the right-hand-side is termed the local


storage change and the last three terms are changes via
advection through the fluid. Using vector notation, Equation
(6.5.5) can be written as a general total derivative operator:

d
()
() =
+ (V i )()
dt
t

(6.5.6)

where this notation will show up in the conservation laws to


120

follow.

The conservation of momentum is simply an expression
of Newtons 2nd law, i.e.:

" du dt %
F = a = dV = $$ dv dt ''
m
dt $
'
# dw dt &

(6.5.7)

where forces can be in any orthogonal direction due to spatial


pressure variations in those directions. The viscosity term is
given by:

V
t visc

"
2
2
2
2
2
2
$ u x + u y + u z
$
= $ 2 v x 2 + 2 v y 2 + 2 v z 2
$ 2
2
2
2
2
2
$# w x + w y + w z

(
(
(

= 2 V;
where F is the set of forces acting on the fluid, m is the air
parcel mass, and a is the acceleration. Note that the
derivative in Equation (6.5.7) is the total derivative so that
using Equation (6.5.6):

dV V
=
+ (V i )V
dt
t

(6.5.8)

where the above is a vector equation which in this case has


three components. In a non-rotating reference frame the
relevant forces (per unit mass) are those due to: pressure
gradient, viscosity (friction), and gravity eects so that one
can write:

dV V
V
V
V
=
+ (V i )V =
+
+
dt
t
t press t visc t grav

(6.5.11)

kinematic viscosity

which represents frictional forces due to fluid viscosity and


spatial variations in shear stresses. Finally, the gravity term is
given by:
T
V
= #$0, 0,g %& = g
t grav

(6.5.12)

which acts only in the vertical direction. Hence, in a nonrotating reference frame the governing momentum equation
can be written concisely as:

dV V
1
=
+ (V i )V = p + 2 V g
dt
t

(6.5.13)

(6.5.9)
In a rotating reference frame an additional apparent force,
i.e. the so-called Coriolis force, becomes relevant. The rotation
is characterized by the rotation vector:

where the pressure gradient term is given by:

V
1
1 $ p p p '
= p = & , , )
t press

% x y z (

)
)
)

%
'
'
'
'
'&

(6.5.10)

= = 2 rad/day

(6.5.14)

which points in the direction of the axis of rotation. For a thin


121

atmosphere such as on Earth, the Coriolis term acts almost


exclusively in the horizontal plane and is given by:

V
t coriolis

$ v sin
&
= 2 V 2 & u sin
&%
0

$ fv '
'
&
)
)
=
2

& fu )
)
&
)
)(
% 0 (

(6.5.14)

Ug =

where f is the Coriolis parameter. Putting this with Equation


(6.5.13) yields the full momentum equation:

V
1
+ (V i )V = p + 2 V g 2 V
t

(6.5.15)

which represents the governing equation of the velocity field in


the fluid. It should be kept in mind that this vector equation
has three component equations which can be written out
explicitly term-by-term. The importance of each term in the
above momentum equation depends on the context.

Before moving to the other conservation laws, some
useful special cases of the momentum equation can provide
insight on various flow phenomena. One special case is
geostrophic flow or geostrophic balance which corresponds to
the case where the acceleration terms (terms on left-handside) and the friction term (second term on right-hand-side)
are very small compared to the other terms. This is often the
case aloft in the atmosphere. Under these conditions the above
horizontal momentum equation can be written as:

2 U g =

1
p

where the subscript refers to geostrophic conditions and the


vector Ug contains only the horizontal velocity components (u
and v). The equation can be rearranged to show:

(6.5.16)

1
k h p
f

(6.5.17)

where k represents the vertical unit vector (i.e. in the z


direction) and the h subscript denotes that the gradient is in
the horizontal plane. Using the right-hand-rule for cross
products, this shows that the geostrophic flow is orthogonal to
the vertical unit vector and pressure gradient, which means
that the geostrophic flow is parallel to the pressure contours
in the horizontal plane. Hence, based on a map of pressure
aloft in the atmosphere (where geostrophic balance is
generally a good approximation) one can readily diagnose the
large-scale flow patterns. Equation (6.5.17) is often recast in
terms of the so-called geopotential height (Zg), which is
eectively the height corresponding to a given pressure level,
and can be written as:

Ug =

g
k hZ g
f

(6,5.18)


An example is shown in Figure 4.19 in terms of
geopotential height (at 400 mb in the atmosphere) over the
U.S. At any given location, the direction of the geostrophic
flow can be diagnosed by the cross product. At the sample
point shown in Figure 4.19, the vertical vector is out of the
page, the geopotential gradient is southward and therefore the
122

geostrophic flow is westerly (i.e. toward the east). The


magnitude of the geostrophic flow (at a given latitude) is
determined solely by the magnitude of the geopotential height
(or pressure) gradient. Note that the geopotential height
varies in space and hence the direction and magnitude also
vary spatially. The flow is, as expected, mostly westerly with
some areas of cyclonic flow and will generally follow the
geopotential contours. The areas with highest magnitude
winds will be where the contours are closest, i.e. near
Washington/Oregon.

Another special case of the momentum equation is for
the vertical momentum equation when the acceleration and
friction terms are negligible in which case:

1 p
p
g = 0
= g
z
z

(6.5.19)

This illustrates that under these conditions the vertical


momentum equation can be replaced by the hydrostatic
equation. Note that this does not indicate that the vertical
velocity w is zero, rather just that vertical accelerations are
negligible. Hydrostatic models of the atmosphere generally
make this approximation to aid in the solution of the
governing equations.

F IGURE 4.19 Geopotential height map (in meters) at 400 mb


in the atmosphere over the continental U.S. The vertical vector
(out of the page), geopotential height gradient, and geostrophic
flow vectors (not to scale) are shown at a given point.


The conservation of air mass equation can be written as
the balance between local mass change and convergence of
mass flux:

= i ( V) = (V i ) i V
t

(6.5.20)
123

which can be rearranged in terms of the total derivative to


get:

d
= i V
dt

(6.5.21)

In the atmosphere it is often valid to assume that the lefthand-side is much smaller than the right-hand-side so that the
air mass conservation equation could be replaced by:

i V = [u x + v y + w z ] = 0

(6.5.22)

which is simply stating the flow is non-divergent.



The conservation of atmospheric water vapor can be
written in terms of specific humidity as:

dq
= body sources/sinks
dt

(6.5.23)

where the body source/sink terms typically correspond to


vapor diusion, an evaporative source from liquid/solid water
(Sevap, [Sevap]= kg m-3 s-1) and condensation sink from liquid/
solid water (Scond, [Scond]= kg m-3 s-1) from clouds respectively.
This can be written more explicitly as:

" 2q 2q 2q % Sevap S
dq
= q $ 2 + 2 + 2 ' +
cond
dt

y
z &
# x

(6.5.24)

Similarly, a conservation equation for liquid/ice water content


(ql) can be written:

Sevap Scond
dql
=
+
dt

(6.5.25)

Note that aside from the vapor diusion in Equation (6.5.24)


the two water conservation equations share the same source/
sink terms, but that they are of opposite sign. In other words,
evaporation from cloud droplets will be a source for vapor,
but will be a sink for liquid/ice. Similarly, condensation of
vapor will be a sink for vapor, but a source for liquid/ice.

Finally, based on the 1st Law of Thermodynamics one
can derive the conservation of energy equation in terms of the
so-called potential temperature:

T d
= 2 Rn LvSevap + LvScond ;
dt
R /c
#p& d p
= T %% ((
; p0 = 100000 Pa
p
$ 0'
thermal conductivity of air

c p

(6.5.26)

where the terms on the right-hand-side represent diusive


heat flux, convergence of net radiation, evaporative cooling,
and latent heating. Note that the latter two terms are the
same sources/sinks in the water conservation equations, but
converted to equivalent energy fluxes associated with phase
change. The net radiation convergence is essentially related to
the absorption/emission of radiative fluxes across the
atmospheric profile. Hence that term depends on the water
vapor profile (along with other radiatively active gases). The
potential temperature is a scaled temperature that accounts
124

for adiabatic temperature changes, i.e. the potential


temperature is conserved (constant) under adiabatic changes.

We have now defined the full set of governing equations:
Equations (6.5.1), the vector Equation (6.5.15), and Equations
(6.5.20), (6.5.24), (6.5.25), and (6.5.26). This represents eight
equations in terms of eight states (pressure, density,
temperature, three velocity components, specific humidity,
liquid/ice water content). The set of seven partial dierential
equations and one diagnostic equation (of state) are, among
other factors, coupled via velocity in the advection terms and
through phase changes in the body source/sink terms and the
impact of humidity and temperature on radiative fluxes.
Additionally, boundary conditions, e.g. fluxes at the surface,
play an important role in forcing the equations. Together
these equations form the basis of most numerical climate and
weather prediction models.

125

S ECTION 6

Conceptual Questions
1. Sketch the basic picture of net radiation at the top of
atmosphere as a function of latitude. (Make sure your axes
are labeled.)
2. What part (i.e. latitudes) of the globe absorbs the most net
radiation?
3. If averaged over the whole globe, what is the net radiation
at the TOA?
4. Which is ultimately more responsible for heat
redistribution: atmospheric or oceanic motions?

8. What high-level atmospheric circulation feature consists of


a high velocity westerly wind near the top of the
troposphere?
9. At what latitude bands do the persistent low surface
pressure anomalies occur? At what latitude bands do the
persistent high surface pressure anomalies occur?
10. What is the atmospheric circulation sense (i.e. clockwise
or anti-clockwise) around a low surface pressure zone? A
high pressure zone?
11. In what direction does a westerly wind blow?
12. What is an order-of-magnitude typical value of a nearsurface wind speed (in m/s)? What is an order-ofmagnitude typical value of a top of troposphere jet stream
wind speed (in m/s)?

5. Conceptually, how many circulation cells does the Earth


have? In practice, which cell is far and away the strongest,
most persistent, and therefore most important?
6. What atmospheric circulation feature is primarily
responsible for transporting energy between the equator
and midlatitudes?
7. What atmospheric circulation feature is primarily
responsible for transporting energy between the
midlatitudes and poles?

126

S ECTION 7

Sample Problems
Problem 4.1. Model output from a ten-year climate model
simulation yields the average values for outgoing TOA
longwave radiative fluxes and planetary albedo shown below.
a) Plot the planetary albedo as a function of latitude. Explain
the physical reason for the latitudinal dependence of planetary
albedo. Note: Planetary albedo is the bulk (combined eects
of surface and atmosphere) reflectivity representing what
fraction of incoming TOA shortwave radiation will be reflected
back into space.
b) Plot the outgoing longwave as a function of latitude.
Explain the physical reason for the latitudinal dependence of
outgoing TOA longwave flux. Note: The outgoing TOA
longwave flux is the combined flux emitted by the atmosphere
and the flux emitted by the surface that is transmitted to the
TOA.
c) Compute the annual mean TOA incident solar radiation for
all the latitudes tabulated below and plot the annual solar
incident radiation as a function of latitude. The MOD-WET
code TOA_incoming_solar.m may be useful.
d) Compute and plot the net TOA shortwave as a function of
latitude. On the same figure plot the net TOA longwave. Also
compute, and plot on the same figure, the total net TOA

LATITUDE

PLANETARY
ALBEDO (%)

OUTGOING
LONGWAVE
(W M-2)

-90

58.61

148.1

-84

63.39

150.9

-76

64.85

160.0

-68

57.92

186.6

-60

49.30

201.5

-52

43.54

209.9

-44

38.25

220.6

-36

31.46

235.3

-28

27.62

248.3

-20

24.79

254.6

-12

23.02

254.3

-4

20.50

250.6

19.39

248.9

12

21.26

250.5

20

25.50

254.6

28

29.47

249.5

36

33.26

235.7

44

37.75

225.0

52

40.63

213.8

60

44.05

206.1

68

48.04

197.3

76

58.19

191.6

84

60.92

182.9

90

59.47

192.6
127

radiation as a function of latitude. What are the maximum/


minimum net TOA radiations and where are they occurring?
e) If you were to integrate the total incoming and outgoing
fluxes over the areas over which they occur (to get the net
energy fluxes at the TOA), how would you expect the
magnitudes of the incoming and outgoing fluxes to compare?
Explain your answer.
Problem 4.2.
a) Explain how the energy absorbed from the Earth as a
function of latitude is related to the atmospheric general
circulation. Sketch and describe the key atmospheric
circulation features that are chiefly responsible for
redistribution of energy in the Earth system. Be specific.
b) What atmospheric circulation feature is primarily
responsible for transporting energy from the equatorial region
to the midlatitudes? Explain how the transport of water vapor
from the equator toward midlatitudes is related to energy
transport.

128

Chapter 5

Precipitation
Processes

S ECTION 1

9. Read and understand a typical surface weather map

Learning Objectives

10. Describe and explain the latitudinal dependence of


precipitation

By the time you finish this chapter you should be able to:

11. List and describe some typical ways of measuring


precipitation

1. Describe the primary cooling mechanism responsible for


condensation of liquid water in air

12. List and describe the primary ways of characterizing


areal average precipitation over a region/watershed

2. Define the dry adiabatic lapse rate and state its value
3. Define the moist adiabatic lapse rate and describe how/
why its typical value corresponds to the dry adiabatic
lapse rate
4. List and describe the two primary growth mechanisms for
water droplets in clouds
5. Use the Marshall-Palmer (or other) size distribution
function to compute the number of droplets in a cloud and
the liquid water content in clouds
6. Draw and label a schematic of the various processes
related to cloud/precipitation physics
7. Compute the precipitation reaching the surface using the
integral relating drop size distribution, uplift and terminal
velocities, and critical drop diameter
8. List and describe the four primary lifting mechanisms
involved in the meteorology of precipitation
130

S ECTION 2

Thermodynamics of Cloud
Formation

Precipitation generally refers to the flux of water
reaching the surface and is a result of a set of processes that
ultimately convert atmospheric water vapor to hydrometeors
(either liquid droplets or ice crystals) that can fall through the
atmosphere. Clouds are just a collection of hydrometeors that
are not large enough to fall to the surface. The processes
required to form clouds (and ultimately precipitation) include
both large scale meteorological phenomena as well as
microphysical processes on the molecular/cloud scale. For a
more detailed discussion of cloud and precipitation processes
the reader is referred to Yau and Rogers (1984) and
Pruppacher and Klett (2010).

The first necessary requirement for cloud formation is a
condensation mechanism whereby water vapor molecules are
converted to liquid or ice. To understand water vapor
transformations it is convenient to use parcel theory, which
is simply a conceptual representation used to follow parcels
of air as they move through the atmosphere. To a good
approximation, moving air parcels carry vapor around with
them. In other words, the specific humidity of an air parcel (q)
is conserved (i.e. q = constant) as a parcel moves, provided
there is no condensation. Given this fact, vapor will be carried
around by a parcel right up until the parcel becomes

saturated. Recall that the upper limit to specific humidity is


the saturated specific humidity (qs) as described by the
Clausius-Clapeyron equation (Equation (2.4.5)) and that this
water holding capacity is primarily dependent on air
temperature (and on pressure). As air cools the saturated
specific humidity decreases (and vice versa). So for a parcel to
reach saturation, the air parcel must be cooled in some way.

One can imagine dierent thermodynamic mechanisms
for cooling an air parcel, i.e. radiative cooling, cooling via
conduction, etc. However the most ecient mechanism for
cooling air is so-called adiabatic cooling as a result of vertical
uplift. The basic mechanism is as follows:
1) If an air parcel is lifted, it expands due to a decrease in the
ambient pressure (which the parcel equilibrates with quickly)
as described by the hydrostatic equation:

dp
= pg
dz

(5.2.1)

where the p subscript refers to the parcel property and the


density is described by the equation of state (Ideal Gas Law).
2) Expansion without an external heat input (adiabatic
process) causes cooling as described by the 1st Law of
Thermodynamics:

Energy input dQ = c pdTp +

dTp
Tp

Rd dp
cp p

1
dp = 0 (adiabatic)
p

(5.2.2)
(5.2.3)
131

3) Combining Equations (5.2.1) and (5.2.2) yields the socalled dry adiabatic lapse rate, which is defined as:

dTp

g
9.8 m s 2
1
d =
=

=
9.8
K
km
dz
c p 1004 J kg -1K-1

(5.2.4)

where the negative sign is used because it is understood that


the air parcels temperature cools with height. Here dry
refers to the fact that there is no condensation of water in the
adiabatic process (hence no heat input) and lapse rate often
refers to a temperature change with height.

What this tells us is that if an air parcel were lifted 1 km
in altitude its temperature would cool by almost 10 K.
Equivalently, if an air parcel is at the surface (with a surface
temperature Ts) and it is lifted upward, the parcel
temperature with height will be:

Tp (z) = Ts d z

(5.2.5)

which will be valid over the height of the atmosphere where


the air parcel is sub-saturated. This rapid cooling mechanism
is generally responsible for an air parcel (if lifted) ultimately
becoming saturated (i.e. reaching the state where q = qs(Tp,
p)), which then leads to cloud formation and potentially
precipitation. In parcel theory it is generally assumed that any
condensate is left behind, i.e. not carried further by the
moving parcel. The mechanisms responsible for the uplift will
be discussed in the next section. It should be noted that the
dry adiabatic rate also applies to descending air, only in

reverse (i.e. descending air will generally warm by


approximately 10 K/km).

The height at which a rising air parcel becomes
saturated generally corresponds to the cloud base and is
referred to as the lifting condensation level (LCL). Any
additional uplift will continue to cool the air, but at a
dierent rate. The air parcel will be kept saturated (i.e. q =
qs(Tp, p)), but since the saturated value continues to decrease,
water will be squeezed out of the air parcel into the form of
condensate (liquid water). The process of condensation
involves a release of energy, which counteracts some of the
cooling due to expansion. The rate at which air cools as it
ascends in the presence of condensation is referred to as the
saturated (or moist) adiabatic lapse rate and can be shown to
be:
d
s =
(5.2.6)
Lv dq s
1+
c p dT
Note that the last term in the denominator is just an
expression of the Clausius-Clapeyron equation (in the form of
saturated specific humidity), which is a function of
temperature and pressure. So the saturated adiabatic lapse
rate is not constant, but varies with the state of the parcel. In
the limit of dqs/dT = 0 (which would occur at cold
temperatures (high altitude) or when all the vapor is
removed) the saturated lapse rate would equal the dry
adiabatic lapse rate. In the lower to middle troposphere the
saturated adiabatic lapse rate typically is about 6-8 K/km,
132

3 km. It should be kept in mind that the location of the cloud


base depends on the initial temperature and humidity
contained in the near-surface air parcel. For example if the air
was starting at a warmer temperature at the surface (but had
the same specific humidity), it would take an additional
amount of uplift to reach saturation via adiabatic cooling.
Alternatively, if the air parcel was more moist to start (but
with the same temperature), it would reach saturation at a
lower altitude.
E XAMPLE 5.2.1
Suppose an air parcel near the surface has a
temperature of 25 degrees Celsius and specific
humidity of 10 g/kg. Suppose that the
temperature profile of the ambient air is given
by:

T(z) = Tsurf z

F IGURE 5.1 Conceptual picture of an air parcel rising through


the atmosphere. Prior to condensation it cools at the dry adiabatic lapse rate, while after condensation it cools at a moist
adiabatic lapse rate (from paos.colorado.edu/~toohey/Fig_10.jpg).

but can vary significantly. A schematic of the lapse rates


involved in air parcel movement and cloud formation is shown
in Figure 5.1. In this example the cloud base (LCL) occurs at

with an ambient lapse rate of 5 K/km and the


pressure profile of the ambient air is given by:
g

p(z) = psurf

!T(z) $Rd
#
&
#T &
" surf %

where Tsurf and psurf are the surface temperature


and pressure respectively with a surface pressure

133

E XAMPLE 5.2.1 ( CONTINUED )


value of 1000 mb. If the air parcel is lifted, what
is the expected cloud base height (or lifting
condensation level (LCL))? Qualitatively, how
will the answer change if the surface air were
moister? Or colder? What is the saturated
adiabatic lapse rate at the LCL?
It is commonly assumed that an air parcel will quickly
equilibrate with the ambient air pressure so that the
pressure profiles of the parcel and ambient air are the
same. The parcel air temperature profile however may
dier from the ambient profile since it is governed by the
dry adiabatic lapse rate. A special case would be where
the ambient air has a temperature lapse rate equal to the
dry adiabatic lapse rate, but this generally may not be
the case. It is also generally assumed that the specific
humidity of the air parcel will be constant until it is
lifted to the altitude where saturation occurs. Saturation
occurs due to a decreasing saturated specific humidity
with height and is defined by the condition:

q = q s (z LCL ) =

es (Tp (z LCL ))
p(z LCL )

E XAMPLE 5.2.1 ( CONTINUED )

#L
es 0 exp % v
%R
$ v

#1
&&
1
%
((
%T T z ((
$ 0
surf
d LCL ''
g

psurf

#T z &Rd
LCL
% surf
(
%
(
T
$
'
surf

which simply states that at the LCL the specific


humidity of the parcel will equal the saturated specific
humidity of the parcel. Any additional lifting will yield
condensation and hence clouds. Note that the saturated
vapor pressure profile is governed by the parcel (dry
adiabatic) lapse rate and the pressure profile is governed
by the ambient pressure profile.
The cloud base height or LCL (zLCL) is defined implicitly
by q = qs(zLCL) where q is the starting surface specific
humidity and is equal to 10 g/kg. Note that this
condition forms a nonlinear equation that cannot be
solved for in closed form. Hence this becomes a
rootfinding problem that must be solved numerically.
Doing so, yields an LCL equal to approximately 1740 m.
If the surface air were moister, it would be closer to
saturation and therefore will not need to be lifted as far
to reach saturation. For example, if the surface air had a
specific humidity of 12 g/kg, the LCL would be reduced
to approximately 1310 m. The opposite would be true if
134

E XAMPLE 5.2.1 ( CONTINUED )

E XAMPLE 5.2.1 ( CONTINUED )

the air were drier (i.e. it would have a higher LCL). For
colder air, the parcel will also be closer to saturation and
therefore have a lower LCL. For example, if the surface
air had a temperature of 20 degrees Celsius, the LCL
would be reduced to approximately 950 m.

(9.8 10 3 K/m)
s =
(2.5 10 6 J/kg)
1+
(6.84 10 4 kg/kg/K)
(1004 J/kg/K)

The saturated adiabatic lapse rate at the LCL can be


computed by first computing the temperature and
pressure at the LCL and then the slope of the derivative
of the saturated specific humidity with temperature:

which shows how the saturated lapse rate is significantly


reduced relative to the dry adiabatic lapse rate. As
lifting (and cooling) continues, the derivative of the
saturated specific humidity will approach zero and the
the saturated adiabatic lapse rate will approach the dry
adiabatic lapse rate.

TLCL = 298.15 K (9.8K/km)(1.74 km) = 281.1 K

= 3.6 10 3 K/m = 3.6 K/km

9.81 m s -2

! 281.1 K $(287 J/kg/K)(5103 K/m)


pLCL = (100000 Pa) #
&
298.15
K
"
%
= 66860 Pa
dq s
(0.622)
2.5 10 6 J/kg
=

dT (66860 Pa) 461 J/kg/K


# 2.5 10 6 J/kg #
&&
1
1
(611 Pa)exp %

%
((
$ 461 J/kg/K $ 273.16 K 281.1 K ''
(281.1 K)2

= 6.84 10 4 kg kg -1 K-1

The saturated adiabatic lapse rate is then given by:

135

S ECTION 3

Cloud Microphysics

While cooling via adiabatic uplift is generally necessary
for cloud formation, it is not sucient. Several microphysical
processes are responsible for the formation, growth, and
ultimate conversion of vapor to precipitation.

It turns out that while condensation is generally
observed when an air parcel reaches saturation, water vapor
alone is often insucient to form hydrometeors. Rather,
particulates that are pervasive in the atmosphere (aerosols,
ash, salt, etc.), which are often referred to as cloud
condensation nuclei (CCN), are at the heart of the initial
conversion of water vapor to hydrometeors. These CCNs are
typically on the order of 0.2 micrometers in diameter
compared to a water molecule which is on the order of 0.3
nanometers (i.e. three orders of magnitude smaller). The
process of water vapor molecules spontaneously joining
together via condensation to form a liquid droplet (or ice
crystal) is called homogeneous nucleation. While this is
theoretically possible, the small size of a water molecule,
combined with the surface tension eects as a function of
radius (i.e. higher surface tension for a smaller radius), makes
such droplet formation relatively rare and/or inecient.
Instead, CCNs provide a nucleation site for the water vapor
molecules to condense onto via what is referred to as
heterogeneous nucleation (Figure 5.2). The larger nuclei makes

F IGURE 5.2 Early in cloud formation large particles serve


as nuclei for water vapor to condense on (from

weathergamut.com/wp-content/uploads/2001/10/nuclei.jpg).

such condensation easier and such droplets more stable. In


cases where the number of CCNs is limiting, the cloud may
actually be super-saturated whereby the relative humidity in
the cloud is greater than 100% (generally by no more than a
few percent). This cloud droplet initiation process is the
theoretical basis for cloud seeding where artificial
particulates (often silver iodide) are introduced into the
atmosphere in an attempt to increase cloud formation with
the hope of increasing precipitation.

The dierence between a precipitating vs. nonprecipitating cloud is simply a reflection of the size of the
hydrometeors in the cloud. Figure 5.3 shows the typical drop
sizes in a cloud, ranging from the CCNs (radius of 0.1
microns) to a typical droplet of size 10 - 50 microns to a
typical rain drop of radius 1000 microns (1 mm). For
precipitation to occur, cloud hydrometeors must grow to
136

uplift mechanism is responsible for the cloud existence in the


first place and therefore an updraft (vertical velocity) is
generally present. Small drops will generally be suspended if
their terminal velocity (proportional to their mass) is
comparable to the updraft, but if too small may be blown
upward by the updraft, often causing evaporation of the
droplet and dissipation of the cloud. The magnitude of the
updraft largely depends on the mechanism for uplift and will
be described in more detail below.

F IGURE 5.3 Illustration of typical relative sizes of various hydrometeors in a warm cloud (from geog.ucbs.edu/~joel/g110_w08/
lecture_notes/precip_processes/agburt07_02.jpg).

sucient size to: i) achieve terminal velocities (Vt) that are


larger than the updraft velocity (Vup) in the cloud and ii)
survive evaporation between the cloud base and the surface.

In warm clouds (i.e. those above freezing and populated
solely with liquid water and vapor) there are two primary
growth mechanisms. The first is simply a diusional
condensation process. This is where individual water vapor
molecules diuse through the air (as a result of vapor
concentration gradients) and attach to a CCN or existing
droplet. In terms of speed, this process is slow since eectively
one vapor molecule is being added at a time to an existing
hydrometeor via diusion. Assuming sucient CCNs are
present, this condensation process is a limiting factor early in
a cloud life-cycle. It is important to keep in mind that an


Once droplets
of sucient size exist
in the cloud, some
start to fall relative
to the updraft.
Bigger droplets will
fall faster than
smaller ones,
overtaking them
(Figure 5.4). There
will generally be
some collisions
between droplets and
some coalescence.
This collision/
coalescence growth
mechanism yields a
much faster growth
rate and is generally
responsible for the

F IGURE 5.4 Illustration of collision


and coalescence process in a cloud
(from cmmap.org/images/learn/clouds/
cc.jpg).

137

unit volume of air per unit diameter bin. Figure 5.3 provides
some indication of this distribution, with small cloud droplets
being in large numbers (e.g., 106 droplets per liter of air in
this example) compared to much fewer large cloud droplets
(e.g., 103 per liter) and even fewer raindrops (e.g., 1 droplet
per liter). These values are just an example and the actual
distribution varies considerably from one cloud to another.
The most common DSD used to describe precipitating (warm)
clouds is the so-called Marshall-Palmer (1948) distribution:

N(D) = N 0 exp(cD)

F IGURE 5.5 Conceptual picture of cloud droplets being lifted

while too small to overcome updraft and ultimately falling once


large enough (from Ackerman and Knox, 2006).

development of large droplets. Ultimately, many of these will


leave the cloud base as precipitation (Figure 5.5).

The combination of these processes leads to a wide
spectrum of droplet sizes in a cloud. To quantify cloud
properties, a drop size distribution (DSD) function (N(D)) is
often used, which simply describes the number of droplets per

(5.3.1)

where N0 is generally considered a constant parameter (0.08


cm-4) and c is a parameter dependent on the cloud conditions.
One can think of the DSD as a probability density function
(pdf) describing the distribution of droplet sizes and the
Marshall-Palmer distribution is therefore equivalent to an
exponential pdf. In this context, one can show that the
parameter c is equivalent to the inverse of the average droplet
diameter, i.e.:

c=

1
;
D

D average drop diameter

(5.3.2)

Based on the above definition, the dimension of the DSD is


[-]/L3/L=L-4. Like most DSDs, the Marshall-Palmer
distribution implies a higher number of small droplets
compared to large droplets. It is however a simplification of
reality in that it predicts non-zero numbers of zero diameter
droplets (not physically possible). It is used however due to its
analytical tractability and because the erroneous specification
138

zero-sized droplets generally does not introduce significant


error in other characteristics. More realistic DSDs that are
often also used include Lognormal or Gamma distributions.

The primary use of the DSD is that it can then be used
to quantify other properties related to a given cloud. Some
simple examples include the total number of drops in the
cloud, i.e.:

Total # of drops = N(D)dD

(5.3.3)

and the total liquid water content (assuming spherical


droplets), i.e.:

D3
LWC = w
N(D)dD
6
0

E XAMPLE 5.3.1 ( CONTINUED )


distribution with an average drop diameter of 0.5
mm. a) What is the total number of drops in the
cloud (per cubic meter of air)? and b) What is
the liquid water content (in grams per cubic
meter of air)?
Based on the average drop diameter, the MarshallPalmer DSD parameter is equal to:

c=
(5.3.4)

which represents the mass of liquid water in the cloud per


unit volume. The first part of the integrand is the mass of a
droplet of diameter D, which is then multiplied by the number
of drops of that size bin and summed up via the integration
across all sizes. The integration limits used are to cover all
possible sizes although a physical upper limit would be drops
on the order of a few millimeters in diameter.
E XAMPLE 5.3.1
Suppose a cloud has a Marshall-Palmer drop size

1
1
=
= 2 mm -1
D 0.5 mm

a) The total number of drops can be obtained by


integrating the DSD:

Total # of drops =

N 0e cD dD =
0

N 0 cD
e
0
c

N0
N
[0 1] = 0
c
c
4
4 (100 cm)
(0.08 cm )
4
1
m
=
= 8000 drops m -3
1000 mm
2 mm 1
1m
=

b) The liquid water content is obtained assuming


spherical water droplets and integrating the product of a
single drop mass and the DSD over all drop sizes:

139

diameter corresponding to Vt(D = Dup) = Vup. Alternatively a


precipitation mass flux can be obtained by multiplying by the
density of water. The terminal velocity is usually
parameterized in terms of the drop size, with the simplest
(linear) form being given by:

E XAMPLE 5.3.1 ( CONTINUED )

LWC =

N0
D3
w
N 0e cD dD = w
6
6

w N 0 3! w N 0
=
6 c4
c4

De

3 cD

dD

Vt = D

(100 cm)4
(1000 kg m ) (0.08 cm )
1 m 4 1000 g
=
4
1 kg
-1 4 (1000 mm)
(2 mm )
1 m4
= 1.57 g m -3
-3

The updraft velocity is a complicated function of many


dynamical factors. The simplest models tend to be
proportional to the cloud thickness, e.g.:

Vup = K Z ;


The DSD can also be used in the calculation of the
precipitation rate leaving the cloud base by considering the
fact that the relative fall velocity of a drop is given by the
dierence between the terminal velocity and the updraft
velocity: (Vt -Vup) and that only drops that have a positive fall
velocity (i.e. that can overcome the updraft) will actually
contribute to the flux leaving the cloud base. Given this, the
precipitation leaving the cloud base can be expressed as:

D3
Pb =
N(D)[Vt (D) Vup ]dD
6
Dup

Z = cloud thickness

(5.3.7)

where the proportionality constants in Equations (5.3.6) and


(5.3.7) must be estimated or specified.

which shows that there is less than 2 g of water per


cubic meter of air in the cloud.

(5.3.6)

(5.3.5)

which will have dimensions of L T-1 and where Dup is the


To compute the precipitation that reaches the surface
one needs to account for the sub-cloud evaporation from
falling drops. This has two impacts: i) a critical diameter (Dc)
exists below which all drops will evaporate and ii) a fraction
of mass from the droplets with D > Dc will also be lost (due
to partial evaporation). To account for these, a modified
integral can be used to get the surface precipitation (P):

P=

Dmin

D3
X(D)
N(D)[Vt (D) Vup ]dD
6

(5.3.8)

where X(D) is a mass loss factor (Georgakakos and Bras,


1984) that varies between 0 and 1 (and is generally a function
of drop size diameter) and the lower integration limit is given
by:
140

Dmin = max[Dc ,Dup ]

(5.3.9)

which bounds the integral such that only droplets that


contribute to the precipitation flux reaching the surface are
included. Note that not all water leaving the cloud base will
reach the surface, which is confirmed by the fact that
Equations (5.3.5) and (5.3.8) imply that P is less than or
equal to Pb.

To summarize the sequence of events that produce
precipitation (for warm clouds):
1. Moist air is lifted (q = const.) and cooled at the dry
adiabatic lapse rate
2. Condensation occurs at the cloud base or LCL (q = qs)
3. Additional uplift by the updraft causes parcel to cool at the
moist adiabatic lapse rate (and maintains q = qs with qs
decreasing)
4. Droplet growth occurs via diusional condensation
5. Falling drops overcome updraft velocity and grow due to
collision/coalescence mechanism
6. Droplets leave cloud base where they enter a sub-saturated
environment that starts to evaporate water from droplet
surface (via diusional mechanism)
7. Those droplets that survive the path to the surface (i.e.
dont fully evaporate) contribute to the surface precipitation
flux

E XAMPLE 5.3.2
Estimate the cloud-base precipitation rate (in
mm/hr) for the cloud described in Example 5.3.1.
First assume the prevailing updraft velocity is
negligible and for simplicity that the terminal
velocity has a linear form as in Equation (5.3.6)
with a multiplicative parameter: 1 x 103 s-1.
Qualitatively, how would the answer change if the
prevailing updraft velocity were 0.5 m/s?
Qualitatively, how would the precipitation rate at
the surface dier from that at the cloud base?
The precipitation rate for negligible updraft velocity is
given by:
Pb =

D3

cD

N 0e [D]dD = N 0 e cDD 4 dD
6
6
0

4!
1
N 0 5 = 4 N 0 5
6
c
c
4
4 (100 cm)
= 4 (0.08 cm )
(1000 s 1 )
4
1m
1
5
1 5 (1000 mm)
(2 mm )
1 m5
1000 mm 3600 s
= 3.14 10 6 m s 1
= 11.3 mm h -1
1m
1h
=

For the case of a non-negligible updraft velocity, both

141

E XAMPLE 5.3.2 ( CONTINUED )


the integrand and lower integration limit will change.
The lower limit is increased in order to not include the
smaller droplets that will be blown upward by the
updraft, i.e.:

Vt (Dup ) Vup = Dup Vup = 0


Vup

0.5 m s -1
Dup =
=
= 0.5 mm

1000 s -1
The cloud base precipitation would then be obtained by:
D3
Pb =
N 0e cD [D Vup ]dD
6
Dup

The integrand will generally be smaller due to the


subtraction ofVup and the integration range will be less
as well. Both factors will reduce the amount of cloud
base precipitation. Note that this integral cannot be as
easily evaluated as the one where updraft is neglected. It
has to be done via integration by parts or done
numerically, i.e. using the trapezoidal rule.
If sub-cloud evaporation was also included, the integrand
would be further reduced (mass lost to evaporation) so
that the surface precipitation will be even less than the
answers obtained above. In the limit of saturated subcloud air, the evaporation would be negligible and the
surface precipitation would equal that at the cloud base.


It should be noted that the above description strictly
applies to warm cloud processes. In cold clouds (where
temperatures reach freezing) the basic concepts are the same,
with a few modifications. First, a cold cloud generally consists
of (at least initially) a mixture of ice crystals, super-cooled
liquid droplets, and vapor. The ice crystals can take on a
variety of shapes/sizes as discussed in more detail in the next
chapter. The presence of super-cooled liquid occurs as a result
of the ineciency of homogeneous freezing of existing liquid
droplets that are lifted into the cold region of the cloud. An
additional process occurs in cold clouds as a result of
dierences in the saturated vapor pressure over bulk water vs.
bulk ice as given by their respective Clausius-Clapeyron
equations (Equations (2.4.5) and (2.4.6) respectively). Droplet
or crystal growth/decay is driven by gradients in vapor
pressure between the surface of the hydrometeor and the
ambient cloud air. For sub-freezing temperatures the
saturated vapor pressure over liquid water is higher than that
over ice. So for the same ambient vapor pressure (and same
size liquid and ice hydrometeors), the vapor pressure gradient
will often be away from the liquid droplet and toward the ice
crystal. The end-result is that liquid droplets will decay via
diusion and ice crystals will grow. Depending on the time
scale, this will ultimately lead to growth of ice crystals at the
expense of supercooled liquid droplets. For cold clouds, falling
crystals will collide and form conglomerated crystals (e.g.
snowflakes), and will melt and/or sublimate in the subsaturated cloud layers.

142

S ECTION 4

Meteorology of Precipitation

The previous two sections highlighted the need for an
uplift mechanism to initiate cloud/precipitation processes.
These mechanisms are generally the result of large-scale
meteorological processes, some of which are directly connected
to the general circulation features discussed in Chapter 4.
Here we will focus on the qualitative description of the four
main mechanisms that provide vertical uplift in the
atmosphere.

The first mechanism involves large-scale horizontal
convergence (Figure 5.6). This consists of large-scale air
masses of similar thermal characteristics coming together at
the surface and being forced upward as a result of mass
conservation. The best example of this is the ITCZ which is
the coming together of the surface branches of the Hadley
cells. As mentioned in Chapter 5, the Hadley cells are a strong
and persistent feature of the atmospheric general circulation.
This uplift is the chief mechanism of tropical clouds near the
equator, is global in extent, and occurs throughout the year
(with seasonal shifts north/south of the equator). Of
particular note is the fact that water vapor is plentiful in the
tropics due to the large amount of net radiation near the
equator that drives evaporation. This combined with the

F IGURE 5.6 Schematic showing large-scale horizontal convergence leading to uplift. This mechanism is most likely to occur
at the ITCZ and/or in tropical low-pressure systems (copyright
Pearson Prentice Hall, 2005).

uplift results in persistent clouds and precipitation (Movie


4.4).

Another form of relatively large-scale convergence in the
tropics can occur in conjunction with a so-called tropical
depression. A tropical depression simply refers to an area of
relatively low atmospheric pressure. While there are persistent
highs and lows associated with the general circulation
described above (Figure 4.12), they also occur intermittently
in space and time simply due to semi-random motions and
waves in the atmospheric fluid. At latitudes of +/- 5-20
latitude the Coriolis force is strong enough that cyclonic flow
patterns can form around these lows (counterclockwise in the
143

F IGURE 5.7 Satellite image of a hurricane from space (from


hurricane-facts.com/hurricane-facts.jpg).

Northern hemisphere) with surface winds converging toward


the low pressure center. The convergence is similar to that at
the ITCZ in the sense that the air masses converging are of
similar thermal characteristics. Depending on the prevailing
conditions, these cyclonic storms may quickly dissipate or
grow in strength to ultimately form hurricanes (Figure 5.7).
Unlike the ITCZ, these cyclones generally migrate significantly
over the course of the life-cycle of the storm, often moving to
higher latitudes and gaining in strength over the relatively
warm ocean. Many of the largest storms and extreme events
that occur in such places as the Southeastern U.S. are due to

F IGURE 5.8 Traces of all Northern Atlantic hurricane tracks

from 1851-2005 (from upload.wikimedia.org/wikipedia/commons/3/31/


Atlantic_hurricane_tracks.jpg).

these storms that are initiated via convergence in the tropics.


For example, the tracks of all classified hurricanes in the
North Atlantic between 1851-2005 are shown in Figure 5.8.
The largest storms tend to be seasonal with hurricane season
in the Northern Atlantic occurring between July and
November, with a strong maximum in August/September
when oceanic temperatures are at their peak. Note that the
peak in sea surface temperature occurs after the peak in solar
radiation input due to the thermal inertia of the oceans
(resulting from the high heat capacity of water). Tropical
144

storms (or those that form hurricanes) tend to dissipate on


the order of several days to a week depending on the
prevailing conditions.

A second uplift mechanism, that occurs mostly in
midlatitude regions, is that associated with frontal
convergence. As the name suggests, it too involves the
convergence of air masses, but with the converging air masses
having dierent thermal characteristics (temperatures). Colder
air masses originate at high latitudes, while warmer air masses
originate at lower latitudes. They often come together in
midlatitudes as a result of the horizontal eddies described in

F IGURE 5.9 Schematic showing frontal convergence leading

to uplift. This mechanism is most likely to occur in midlatitude


regions and is associated with horizontal eddies (copyright Pearson Prentice Hall, 2005).

F IGURE 5.10 Schematic showing typical horizontal and vertical scales of a warm front (from sci.uidaho.edu/scripter/geog100/lect/

05-atmos-water-wx/05-part-7-atmos-lifting-fronts/05-30-front-warm-diag.j
pg).

relation to the general circulation in Chapter 4. Their


occurrence is highest in the winter season when the thermal
contrast between polar and tropical/midlatitude air is largest.
A front is simply an interface (actually a strong temperature
gradient) between the two converging air masses (Figure 5.9).
While both air masses are generally moving, one usually
overtakes the other (i.e. has a higher velocity). A front is
labeled a cold front when the cold air is overtaking the warm
or vice versa for a warm front. The uplift mechanism is driven
by buoyancy. Warm air is less dense than cold air. Hence
when they come together the warm air will rise over the cold.
It is this deflection of air upward that initiates cloud and
precipitation formation.

The dierences between cold and warm fronts are not
trivial as they can lead to significantly dierent cloud and
145

F IGURE 5.11 Schematic showing typical horizontal and verti-


Frontal systems are often shown schematically on surface
weather maps (Figures 5.12 and 5.13), where dierent colors/
symbols are used to denote the type of front. Note that the
systems tend to form via cyclogenesis where cyclonic/anticyclonic motions centered around highs or lows initiate
movement of air masses. In addition to warm and cold fronts,
stationary fronts (where there is no relative motion of one air
mass to the other) or occluded fronts (where one front is
overtaken by another) often occur before the frontal storm

cal scales of a cold front.

precipitation characteristics. Figures 5.10 and 5.11 show


schematics of a typical warm and cold front respectively. In a
warm front, the overtaking warm air tends to rise above the
cold air mass with a gentle slope in the vertical interface
between the two air masses (i.e. on the order of 1:150). This
tends to lead to relatively small uplift velocities and interfaces
that can stretch over more than 1000 km (Figure 5.10). To
first order, weaker uplift velocities lead to less droplet growth
and therefore weaker precipitation intensities. Hence, warm
fronts are expected to have potentially wide spread cloud
formation, often ahead of the location of the front at the
surface, but with lower intensity precipitation. In the case of
cold fronts, the overtaking cold air backs up as warm air is
forced to rise. This generally leads to a more abrupt interface
(i.e. a slope on the order of 1:75) with higher uplift velocities.
The result is generally more localized cloud formation of more
significant vertical extent, behind the location of the front at
the surface, and with higher intensity precipitation.

F IGURE 5.12 An example weather map showing frontal sys-

tems leading to cyclongenesis. Letters on map refer to different


types of air masses (c/m = continental/maritime; P/T = polar/
tropical (from sci.uidaho.edu/scripter/geog100/lect/05-atmos-water-wx
/05-part-8-midlat-cyclones/05-31-midlatitude-wave-cyclone.jpg).

146

F IGURE 5.14 Schematic showing orographic uplift (copyright


Pearson Prentice Hall, 2005).

F IGURE 5.13 Surface weather map showing cold/warm


fronts and high/low pressure zones.

dissipates (Figure 5.12). These frontal systems or storms


typically occur over the course of several days to a week and
provide the primary mechanism for winter-time precipitation
in many of the midlatitude regions of the globe.

A third mechanism of vertical uplift is the result of air
masses flowing over topography of significant relief, called
orographic uplift (Figure 5.14). The mechanism is simply
mechanical. Either due to local or large-scale atmospheric flow
patterns, air parcels near the surface are forced to rise due to
the solid lower boundary condition. Once lifted, the typical
thermodynamical processes occur, i.e. adiabatic cooling,
followed by potential condensation, cloud formation and

precipitation. Typically, the air parcels that are lifted up the


windward side of the mountain are forced down the
leeward side back to the surface. An important point is that
the adiabatic process is reversible, i.e. while it provides a
cooling mechanism on the way upward, it provides a warming
mechanism on the way downward. Also, the downward motion
(which warms the parcel) will not involve condensation and
therefore will occur at the dry adiabatic rate over the entire
descent. Whether or not clouds or rain will actually form
depends on three factors: i) the humidity of the surface air, ii)
the temperature of the surface air, and iii) the amount of
elevation change associated with the topography. If there is
little to no water vapor in the surface air, then no amount of
lifting will cool the parcel suciently to be able to squeeze out
enough water to generate clouds and/or precipitation.
Depending on the specific humidity and temperature of the air
and the Clausius-Clapeyron equation, one can easily
147

determine how much cooling would be required to cause


condensation. Given the surface air characteristics, the
magnitude of the elevation change (in combination with the
dry adiabatic lapse rate) will exactly determine whether
condensation will occur. In the case of small topographic
relief, the cooling may not be enough to cause condensation.

If conditions are favorable (i.e. humid air combined with
large topographic features), then condensation will occur at
some height above the surface (generally below the
topographic peak). Cooling up to that level (the LCL) will be
at the dry adiabatic lapse rate, while lifting to the top of the
mountain will cause cooling at the moist (saturated) adiabatic
lapse rate. In cases where the prevailing wind is in a given
direction for most of the year, a particular rain shadow eect
is typically seen where the windward side of the mountain
experiences a relatively humid climate, while the leeward side
experiences a much drier climate. The mechanism for this is
caused by two factors: i) water is left behind on the windward
side in the form of clouds/precipitation and ii) while the
ascending air cooled at the dry adiabatic rate for only a
portion of the ascent, the descending air will warm at the dry
adiabatic rate for the whole descent. In the special case of no
condensation, then one would expect the windward and
leeward surface air to have the exact same temperature and
specific humidity. But in the more general case where
condensation occurs, the leeward side surface air will have less
specific humidity (due to condensate left behind) and will be
warmer. Both of these factors cause a decrease in relative
humidity on the leeward side. Many mountainous regions of

the globe including Hawaii, the Sierra Nevada, Rocky


Mountains, Himalayas, Andes, etc. experience orographic
precipitation as a large portion of their annual totals.

An example of orographic eects and rain shadow can
easily be seen in Figure 5.15 which shows estimates of the
annual precipitation totals for the Hawaiian islands. The
latitude of Hawaii puts it in the trajectory of the trade winds
which blow consistently throughout the year in an easterly
direction. The Hawaiian islands are surrounded by moist

F IGURE 5.15 Mean annual precipitation over the Hawaiian

islands showing the clear signature of orographically-induced


precipitation.

148

tropical ocean air year-round and exhibit significant


topographic relief. The persistent trade winds move the moist
air up and over the islands showing a distinct rain shadow
eect. For example, in the big island of Hawaii, areas of the
windward (in this case eastern) side experience over 8 meters
of rainfall per year, while some of the coastal regions on the
leeward side experience less than 25 cm of rainfall (less than
the mean annual rainfall in Los Angeles). As one would
expect, this stark dierence in rainfall has significant
implications on other aspects of the climate and surface
hydrology. Most notably one can see the dierences in the

F IGURE 5.17 Photo of the Kohala coast on the leeward side of


Hawaii showing the arid climate as indicated by the lack of
vegetation. This region is in relatively close proximity to the location shown in Figure 5.16.

F IGURE 5.16 Photo of Waipio Valley on the windward side of


Hawaii showing the humid climate as indicated by the lush
vegetation.

vegetation patterns. Figure 5.16 shows a picture looking down


into a valley situated on the north eastern coast, while Figure
5.17 is looking down toward the north western coast. The
dierences in vegetation, which occur over a relatively short
geographic distance, are the direct consequence of orographic
precipitation patterns which are responsible for producing
significantly dierent climates.

The fourth and final mechanism responsible for uplift is
usually referred to as thermal convection. This mechanism
149

starts with dierential heating of the surface, where solar


radiation (during the day) heats the surface much more
quickly than the overlying air (Figure 5.18). This is because
the air is relatively transparent to solar radiation, while the
surface is essentially opaque with a high absorptivity. The air
in direct contact with the surface (which is heated via
conduction with the surface) is thereby generally warmer than
the overlying air, which sets up an instability. Warmer air has
a positive buoyancy that will cause it to rise as part of a socalled thermal. This phenomenon can often be seen on a hot
summer day where thermal bubbles of air can be seen in the
distance rising from the surface. Evaporation is often

F IGURE 5.18 Schematic showing thermal convection uplift


mechanism (copyright Pearson Prentice Hall, 2005).

associated with the heating of the surface so that the uplift


combined with the water vapor leads to cloud formation.
Thermal convection is often relatively localized in space and
has a strong diurnal cycle. The surface has thermal inertia so
that it takes several hours of heating to generate a large
instability which explains why convection often takes place in
the early or late afternoon. Additionally, the instability
generally causes convective events to have significantly higher
uplift velocities than the other mechanisms discussed above.
These strong updrafts often lead to tall cumulus clouds
(Figure 5.19; sometimes extending to the top of the
troposphere) that lead to relatively short-duration highintensity storms (often associated with thunder and lightning).

A final point that is important to remember is that the
mechanisms have been described here in isolation, but real
storms, especially those related to extreme events, may be
combinations of several mechanisms. For example it is not
uncommon for thermal convection events to occur in
mountainous areas during summer-time, where slope/aspect
may dictate spatial patterns with areas of increased solar
radiative heating of the surface (which will cause thermal
convection) combined with a background uplift due to
orographic eects. Another example would be winter-time
frontal events which then interact with surface terrain to
enhance cloud formation and precipitation. So while it is
convenient to conceptualize these mechanisms independently,
it is often the interaction between them that leads to
significant cloud formation and precipitation events.

150

E XAMPLE 5.4.1
Three islands have peak altitudes of 510 m, 1020
m, and 1530 m corresponding respectively to
pressures of 950, 900, and 850 mb. Each island
experiences a prevailing wind with surface air
having a temperature of 28 degrees Celsius and
specific humidity of 12 g/kg. The air is lifted up
and over the peak of the island before returning
to sea level on the leeward side. Which, if any, of
the three islands would be expected to have
orographic clouds?

F IGURE 5.19 Photo of a tall cumulus (convective) cloud.

The key point to examine is whether air when lifted to


the peak will have saturated yet as a result of adiabatic
cooling. If so, then clouds will have formed at some
altitude on the windward side. If not, then no clouds will
form anywhere on the windward side. The dry adiabatic
temperature changes from sea-level to the peak for each
island is easily obtained from the dry adiabatic lapse
rate and are: 4.98 K, 9.97 K, and 14.95 K for the
shortest to tallest island respectively. Hence the
respective temperatures at the top of each island are:
296.2 K, 291.2 K, and 286.2 K respectively. The
saturated specific humidity can be evaluated at each
peak as:

151

E XAMPLE 5.4.1 ( CONTINUED )


q s,peak1 =

es (Tpeak1 )
p peak1

# 2.5 10 6 J/kg #
&&
1
1
(611 Pa)exp %

%
((
461
J/kg/K
273.16
K
296.2
K
$
''
$
= 0.622
95000 Pa
= 18.7 g/kg
q s,peak 2 =

es (Tpeak 2 )
p peak 2

# 2.5 10 6 J/kg #
&&
1
1
(611 Pa)exp %

%
((
$ 461 J/kg/K $ 273.16 K 291.2 K ''
= 0.622
90000 Pa
= 14.5 g/kg
q s,peak 3 =

es (Tpeak 3 )
p peak 3

= 0.622

# 2.5 10 6 J/kg #
&&
1
1
(611 Pa)exp %

%
((
$ 461 J/kg/K $ 273.16 K 286.2 K ''

= 11.0 g/kg

85000 Pa

From these calculations it is clear that only the highest


peak has a saturated specific humidity lower than the
surface value. Hence as the surface air parcel is lifted
over either of the first two islands, no condensation will
occur. Only the tallest island would be expected to have
orographically generated clouds.

152

S ECTION 5

Precipitation Climatology
and Extremes

Of particular interest to hydrologists is the expected
(climatological) precipitation, which relates directly to water
supply issues, and extreme precipitation (both positive and
negative), which relates directly to flooding or drought. Here
we focus on a brief summary of climatology and extremes.

The climatology of precipitation is simply the long-term
average. As such it tends to simply be an expression of the
typical mechanisms described above that occur over a given
region. Movie 5.1 shows an animation of global monthly
average water vapor (left panel) and precipitation (right
panel) from 2002-2012 from various data sources. First and
foremost, precipitation requires the presence of water vapor as
a source of condensate. This is clearly shown as there is a high
correlation between the two maps. However, precipitation also
requires the presence of an uplift mechanism (described in the
previous section) and other microphysical processes. Some of
these large-scale mechanisms can clearly be seen, most
notably in the form of the ITCZ. It should be noted that the
animation and other climatologies described below, especially
at global scales, rely on models and remote sensing data for
many areas since the in-situ network is sparse.

Figure 5.20 shows a 30-year climatology for annual
average precipitation based on satellite-borne measurements

M OVIE 5.1 Monthly global water vapor and precipitation


(from earthobservatory.nasa.gov/GlobalMaps/).

(this will be discussed further in Section 6). According to


these results, the average annual global mean precipitation is
approximately 2.7 mm/day (975 mm/year), which is relatively
consistent with the global average of 1 meter/year described
in Chapter 1. Of particular note is the clear signature of the
ITCZ as a result of the Hadley cells (large-scale convergence)
centered around the equator, where annual mean rates of up
to 10 mm/day are seen. Localized areas of peak rainfall are
seen over India and the Indian Ocean (monsoon system), the
Amazon, and other places. The signature of the high-pressure
zones is seen in the map around +/- 30 latitude (i.e. see
correlation between Figure 4.12 and Figure 5.20), with a
secondary peak at higher latitudes.

These large-scale climatological features can also be seen
by taking the zonal average, which is show in Figure 5.21, and
153

F IGURE 5.21 Zonal average of satellite-derived global precipiF IGURE 5.20 Global precipitation climatology as diagnosed
from satellite data from 1979-2010 (from NASA).

illustrates the average precipitation as a function of latitude.


The peak zonal average is almost 6 mm/day centered around
the equator. In the areas of high-pressure zone, the
precipitation dips significantly below the global average. It
rises above the global average at higher latitudes before
dropping o considerably above +/- 60 latitude. The
latitudinal pattern is roughly symmetric, with asymmetry
primarily a result of the asymmetric distribution of land
masses across the globe (and between northern and southern
hemispheres). The pattern is essentially a composite of where
water is available, combined with where persistent uplift
mechanisms exist.

To zoom in from the global scale, Figure 5.22 shows a
30-year climatology for the annual average precipitation over

tation estimates shown in Figure 5.20 showing precipitation


distribution as a function of latitude (from NASA).

the continental U.S. Again, this climatology is a reflection of


the meteorological processes that occur with regularity in a
given region. For example, in the Western U.S. there is the
clear signature of orographic eects (Sierra Nevada, Cascade,
and Rocky Mountain ranges). Much of this precipitation falls
as snow as a result of frontal winter-time storms that result
from cold systems moving southwesterly from the northern
Pacific. In the non-mountainous regions, precipitation is
generally suppressed due to the persistent presence of a highpressure cell o the coast of the Western U.S. during the
summer. The midwestern region tends to have higher
precipitation due to a combination of winter-time frontal
systems along with summer-time thunderstorms as a result of
convective activity. The Eastern U.S. precipitation patterns
154

F IGURE 5.22 Mean annual precipitation climatology over the


continental U.S.

tend to be the result of frontal systems in the winter,


combined with convection and tropical storms (and sometimes
hurricanes) in the summer.

Extreme precipitation is of interest primarily with
respect to flooding and droughts. Flooding generally needs to
be mitigated via the the design of structures like levees,
detention basins or reservoirs, etc. Droughts are generally
mitigated by building supply reservoirs that can store and
supply water over multiple years to buer against the
possibility of dry years. In terms of extreme precipitation (i.e.

large precipitation events), the physics are the same, but often
involve some rare set of circumstances leading to above
average water vapor, uplift velocities, and hydrometeor
growth. In the case of droughts, the mechanisms are less well
known. In some cases they can be simply the result of the
natural variability in the system, but are also thought to be
related to feedbacks in the climate system, where a given state
(i.e. dry) is self-sustaining. A simple example of a proposed
feedback is that if there is initially a negative soil moisture
anomaly (i.e. below average conditions), this will lead to less
available water for evaporation. The reduced evaporation will
lead to reduced water vapor, which will lead to less
precipitation. Finally, the reduced precipitation will lead to
less soil moisture. Such a positive feedback mechanism is one
hypothesis for such sustained drought events. In general,
extreme events (both flood and drought) are much less
predictable than climatology.

In describing extreme precipitation events, three
parameters are generally of interest: precipitation intensity,
storm duration, and frequency of recurrence. It should also be
noted that extreme implies some relative comparison to
climatology (which varies in space/time), i.e. what might be
an extreme event at one location may not be at another.
Analysis of extreme events is therefore generally tied to a
particular region.

To put extreme events into some context, Figure 5.23
shows an example of extreme precipitation events from across
the globe in terms of varying duration and intensity. In
particular, the data points are those events with the
155

and Cherrapunji is associated with monsoon rainfall and


orography. What is interesting is that the data points fall
roughly along an empirical envelope of the form:

R = 425D 0.47

(5.5.1)

where R is in mm and D is the storm duration in hours.

F IGURE 5.23 Maximum amounts of recorded rainfall vs. duration (adapted from Dingman, 2008). Blue dots indicate actual data
points from locations across the globe and the red line is based
on Equation (5.5.1).

maximum recorded rainfall for a given duration. The duration


scale ranges from minutes to months (up to two years).
Examples include: 305 mm in 42 minutes (in Holt, Missouri
(USA) on June 22, 1947), 1.87 m in 24 hours (in Cilaos,
Reunion on March 15-16, 1952), and 22.5 m in six months (in
Cherrapunji, India in 1861). Because these are extreme events,
they are rare. Also, dierent storms are often associated with
dierent physical mechanisms. For example, the Holt,
Missouri storm was a supercell thunderstorm (i.e. thermal
convection) combined with a frontal system, the Reunion
storm was a hurricane combined with the islands orography,


For design purposes, most local codes specify a particular
hypothetical design storm with a characteristic duration
and frequency (return period). For a given region, these
specifications define a particular intensity for which to use in
the design. Example maps can be found in Dingman (2005)
which provides so-called intensity-duration-frequency (IDF)
maps of intensities and durations for dierent return period
(frequency) storms. The return period (i.e. T = 100-year
storm) indicates the average recurrence interval (i.e. once
every T years). Such design storms can then be used in
rainfall-runo models (Chapter 10) for predictions of design
runo hydrographs or peak flows resulting from these design
events.

156

S ECTION 6

Precipitation Measurement

Climatology maps and IDF curves used for design
mentioned in the previous section are ultimately determined
via measurements of precipitation. The traditional method of
measuring precipitation is via a rain gauge, which is simply a
vessel open to the atmosphere that periodically or
continuously records the quantity of water it collects. Figure
5.24 shows an example of a rain gauge. Gauges come in a
variety of forms including: i) non-recording gauges (i.e.
collects cumulative amount that must be measured manually),
ii) weighing-recording gauges (vessel sits on a scale that
measures the cumulative amount of water via its weight), iii)
tipping-bucket gauges (water is routed to a small vessel of
known volume that tips when filled and records each tipping),
iv) optical gauges (measures disturbance of an infrared beam
as a result of passing drops), as well as other varieties. These
gauges are of varying complexity (and therefore cost) and
provide varying information. The first two gauge types listed
above provide cumulative amounts (one manual and one
automated), but do not provide rainfall intensities (i.e. rain
per unit time), while the tipping bucket, optical, and other
types provide the actual rainfall intensities. In using an in-situ
gauge one does not want the gauge to interfere with the
variable it is trying to measure. For this reason, many gauges
have wind shields in order to minimize the prevention of
raindrop collection due to turbulent flow around the gage

F IGURE 5.24 Photo of an all-weather precipitation gauge.

(Figure 5.24).

A key aspect to remember is that a rain gauge is a pointscale measurement. What is generally of most interest is the
spatial field, i.e. the spatial distribution of precipitation, or
the spatial average (mean areal precipitation). If a network of
gauges are deployed with sucient spatial density, then it
would be expected to provide an accurate representation of
the spatial field, or at least its mean. Sucient density in this
context is not an independent variable, but is tightly coupled
to the variability of the underlying field. For example, if a
field has no spatial variability (i.e. is constant in space) then
157

one gauge measurement would be sucient to characterize it.


Precipitation however is a variable with a high degree of
spatial variability, thus requiring a high density network to
characterize it. However the number of gauges deployed is
limited due to both instrumentation and operating and
maintenance costs.

mountainous regions, where a significant amount of


precipitation falls as snow (i.e. in the Western U.S.), is much
more limited than in other areas. The source of rain gauge
data is most commonly government agencies, examples of
which in the U.S. include the National Weather Service
(NWS), National Climate Data Center (NCDC), etc.


Figure 5.25 shows the global rain gauge network. The
key point is that the vast majority of gauges are in developed
countries (U.S., Europe, Japan, Australia) due to costs. So
when large-scale hydrology problems are of interest, this lack
of in-situ gauges can be a factor in characterizing
precipitation. While the U.S. has far and away the largest
network, what is not obvious from the figure is that even in
the U.S. the gage network is insucient to fully characterize
precipitation. Note for example that the number of gauges in


A classical problem in hydrology is the calculation of
mean areal precipitation (MAP), i.e. over a watershed, from a
set of discrete gauge measurements (Figure 5.26). This can be
defined mathematically as:

F IGURE 5.25 Map showing global distribution of location


(yellow dots) of precipitation gauges.

F IGURE 5.26 Illustrative example of Thiessen polygon construction -- basin with gauge locations.

158

= 1 p(x,y,t)dx dy
P(t)
A
A

(5.6.1)

where A is the area of the region over which the averaging is


taking place (i.e. the watershed area), p(x,y) is the true
precipitation field and x and y are spatial coordinates. It
should be noted that the true precipitation field is never
known, instead it is only sampled at select (gauge) locations:
pg=p(xi, yi, t). The primary methodologies used to estimate
the MAP from the gauge data most commonly use a weighted
average. A general weighted average equation can be written
as:

While an arithmetic average is very easy to compute, it is


often inaccurate for representing MAP. One would expect it to
be accurate if the gauges are relatively evenly distributed over
the domain. But in many cases, often for practical reasons
(i.e. accessibility), gauges are unevenly distributed. In such
cases it is best to use a weighted average that takes this into
account.

In hydrology, one method often used to do this is a
graphical method called the Thiessen polygon method. The
graphical version aims to sketch out the area closest to each

= w p (t)
P(t)
g g
g =1

(5.6.2)

where respectively pg is the precipitation and wg is the weight


associated with gauge g and G is the total number of gages.
The weights must generally satisfy (assuming unbiased
inputs):
G

w
g=1

= 1 and wg 1

(5.6.3)

The simplest form of a weighted average is an arithmetic


average, in which case: wg=1/G (i.e. each gauge is weighted
equally). In this case:

= 1
P(t)
G

p (t)
g =1

(5.6.4)

F IGURE 5.27 Illustrative example of Thiessen polygon construction -- TIN construction.

159

gauge and use those sub-areas in the weighted average. In this


way the method can be thought of as a nearest-neighbor
method which can also be easily automated via a simple
program (as will be described in more detail below). The
Thiessen polygon method uses a weighting scheme with:
wg=ag/A, where ag is the subregion centered around each
gauge g that is closest to that gauge. The sub-areas can be
constructed via a simple algorithm:
1. Adjacent gauges are connected to form a network of
triangles (Figures 5.27 and 5.28). These triangles are often

F IGURE 5.28 Illustrative example of Thiessen polygon construction -- final TIN.

called a triangular irregular network (TIN). In general the


border connectors between adjacent gauges are
straightforward, but some interior connectors may seem
ambiguous at first. The shorter distanced connectors are
generally those that should be used in the construction of
the TIN (Figure 5.27, 5.28).
2. Perpendicular bisectors (i.e. at the midpoint) are drawn for
each connector in the TIN. If being done graphically, it is
important to make sure to carefully draw the perpendicular

F IGURE 5.29 Illustrative example of Thiessen polygon construction -- perpendicular bisector construction.

160

bisectors. When the TIN and perpendicular bisectors are


properly constructed, the bisectors will meet in the interior
of each TIN (Figure 5.29).
3. The bisectors need to be connected and cleaned-up to form
the irregular (Thiessen) polygons surrounding each gage
(Figure 5.30).
4. Check the constructed polygons to make sure they indeed
map out the nearest neighborhood areas. A simple check
can be done by randomly selecting a few points surrounding

F IGURE 5.30 Illustrative example of Thiessen polygon construction -- final Thiessen polygons.

each gage and convincing yourself that it is within the


appropriate area for the closest gauge.
5. Finally, determine the area surrounding each gage. When
done graphically, this is most commonly done using a
transparent graph paper overlay which allows for counting
squares within each polygon (Figure 5.31). The weighted
average is then given by:

ag

G
1
=
P(t)
A pg (t) = A ag pg (t)
g =1
g =1
G

(5.6.5)

F IGURE 5.31 Illustrative example of Thiessen polygon construction -- polygon area computation.

161

Rather than determining the total area and sub-areas in


Equation (5.6.5), they can simply be replaced by the total
number of squares and number of squares in each sub-area
respectively.

Note that it is common to only account for the sub-areas
within the overall domain and include gauges that fall outside
the domain (provided they map to some nearby portion of the
domain). For example, in Figure 5.31, two of the gauges are
outside the watershed, but map to sub-areas a1 and a3. It
should also be noted that the Thiessen method only need be
done once. Given a time series of precipitation at the gages
(pg(t)) the time series of MAP uses the same weights at each
time step. Hence the Thiessen polygon can be thought of as a
pre-processing step to get MAP from gauge data. The method
may then be used to get storm-scale MAP, annual average
MAP, etc. That said, one can imagine for large domains with
a large number of gauges, the graphical method leaves much
to be desired. Even if the TIN and bisectors are drawn
accurately, the counting squares method is highly tedious
and prone to error.

You should be able to convince yourself (as mentioned
earlier) when looking at Figure 5.31 that the Thiessen polygon
method is nothing more than assigning a nearest-neighbor
gage to each point in space. The weighting scheme is the same
as assigning the gauge value to each point within the
surrounding sub-area and then averaging all pixels. One can
take advantage of this by creating a simple algorithm that
takes as inputs: geographic coordinates of a set of gauges and
geographic coordinates of the domain of interest (i.e. a

F IGURE 5.32 Illustrative example of automated Thiessen


polygon construction over a large basin.

watershed mask) and computes the distance from each point


to each gauge and finally assigns the nearest gauge to each
pixel. This has been done in the MOD-WET code:
thiessen.m. A simple synthetic example is shown in Figure
5.32 for the Colorado River basin and a set of nearby gauges
(left panel). The above described algorithm is applied and
yields the set of Thiessen polygons shown in the right panel.
Beyond visualizing the polygons, the algorithm outputs the
number of pixels associated with each gauge and therefore can
easily be used to apply Equation (5.6.5) to compute the MAP.
Such an automated approach is most commonly used in
realistic problems, but the graphical method is useful to
conceptualize how the method works. It should be noted that
other methods are often used to estimate MAP, which can
162

include isohyetal, kriging, inverse distance weighting (IDW),


and other methods (Dingman, 2008).
E XAMPLE 5.6.1
For the basin and gauge distribution shown in
Figure 5.26, assume the measured rainfall for a
given storm at gauges 1-5 was: 15 mm, 10 mm,
14 mm, 11 mm, and 10 mm respectively.
Estimate the mean areal precipitation for the
storm using the Theissen polygon approach.
Compare the answer to that obtained from an
arithmetic average estimate.
The first step in the Thiessen polygon approach is the
construction of the polygons themselves. This has been
done in Figures 5.27-5.31. The relative areas can be
determined by counting the squares in Figure 5.31. A
graphical estimate generally will contain some error due
to partial squares. The number of squares associated
with each gage is approximately (rounded to nearest
square): 18, 63, 29, 61, and 52, for a total of 223 squares.
Hence the mean areal precipitation estimate is:

1 !
P =
(18)(15) + (63)(10) + (29)(14) + (61)(11) + (52)(10)#$
"
223
= 11.2 mm
The arithmetic average is instead:

E XAMPLE 5.6.1 ( CONTINUED )

1
P = !"(15) + (10) + (14) + (11) + (10)#$
5
= 12. mm
In this case there is about a 7% dierence between the
two estimates. One would expect the Thiessen polygon
method to be a better one as it accounts for the relative
contribution of rainfall measured at each gage. In the
case where the Thiessen polygons have the same areas,
the two estimates would be the same.


So far the discussion of the measurement of precipitation
has been limited to in-situ measurements via gauges and
estimation of MAP using the Thiessen polygon method. As
alluded to above, the high degree of spatial variability in
precipitation fields makes it dicult for a sparse gauge
network to accurately sample the fields (even using the
Thiessen polygon method). A lack of representativeness in
gauge measurements will end up propagating to the MAP
estimates via Thiessen polygon or other methods. This has led
to a significant amount of research into how to better estimate
precipitation and its spatial structure.

Chief among these newer methods are those under the
category of remote sensing. Remote sensing is becoming
pervasive in hydrology as an enabling technology beyond just
precipitation estimation. It refers to methods used to measure
163

(sense) a process without sampling it in-situ. This is most


commonly done via the measurement of electromagnetic
radiation (and its implicit interactions with media) from
ground-based or satellite-based platforms. The primary benefit
of these techniques is that they explicitly map variables, thus
avoiding the point-sampling issues associated with in-situ
measurement. That said, there are generally other tradeos,
chief among them that in-situ provides a direct measurement,

while remote sensing invariably provides an indirect


measurement that must be used to infer the process of
interest.

One commonly used remote sensing technique for
precipitation estimation is ground-based RADAR (radio
detection and ranging). This method uses pulses of microwave
radiation emitted by a transmitter to probe the atmosphere
for clouds and precipitation (Figure 5.33). The basic idea is
that relatively large particles (or in this case hydrometeors) in
the atmosphere will significantly scatter the radiation, with a
measurable component backscattered to a receiver on the
platform. In simple terms, clear sky conditions will have little
to no backscatter, while an atmosphere populated by a
significant amount of hydrometeors will generate backscatter.
From this signal the precipitation can be inferred. The

F IGURE 5.33 Schematic of a RADAR station sending out a


microwave pulse and sensing the return signal.

M OVIE 5.2 Animation showing the typical scanning behavior


of a single RADAR.

164

RADAR rotates, sending out pulses and receiving


backscattered radiation in all directions, thus mapping out
precipitation patterns of the region. Movie 5.2 provides an
animation of the RADAR scanning process.

The measured quantity by the RADAR is the reflectivity
(Z), which can theoretically be related to the cloud properties
(specifically the drop size distribution as):

Z = D 6N(D)dD

(5.6.6)

which has dimensions of L3, but is commonly expressed with


units of mm6 m-3. As shown in Equation (5.3.5), the
precipitation rate (leaving the cloud base) can also be related
directly to the DSD so that generally a power law
relationships of the following form can be derived:

Z(x,y) = aP(x,y)b

(5.6.7)

where the parameters a and b depend on the type of rain


cloud and other factors, but are such that reflectivity is an
increasing function of precipitation rate. In this context, the
reflectivity is often converted to a log-scale as decibels of Z
(dbZ) where dBZ=10 log10(Z), with Z in units of mm6 m-3.
The log-scale is useful since reflectivity can vary over many
orders of magnitude.

In the U.S., a network of RADAR (called NEXRAD)
stations have been installed to measure precipitation in realtime (Figure 5.34). Typically, a single NEXRAD station scans
the atmosphere within a couple hundred kilometers of the

F IGURE 5.34 NEXRAD RADAR network over the U.S. Dark


brown areas are those with missing coverage.

station. Together, reflectivity fields from many stations can be


mosaicked together. Coverage is pretty complete over the U.S.,
with the main holes being in mountain regions (Figure 5.34)
where the RADAR returns are not useful due to topographic
interference (referred to as ground clutter).

An example of a reflectivity snapshot from one of the
stations shown in Figure 5.34 is shown in Figure 5.35 (i.e.
Minot AFB in North Dakota). The station location is at the
center of the image and the scan provides a reflectivity map
centered on the station. In this particular case a storm system
is clearly seen. Areas of high reflectivity (up to 28 dBZ) occur
to the south east of the station where the storm is centered.
165

These high dBZ values are generally associated with high


rainfall rates. The reflectivity map also illustrates the
underlying heterogeneity in the precipitation field and the
potential problems associated with MAP estimation from a
low density rain gauge network. RADAR provides a relatively
high resolution map of reflectivity (approximately 4-8 km
resolution). In the U.S., the NEXRAD network is becoming
more commonly used for real-time weather and flooding
characterization. An example of a composite image over the
U.S. from the RADAR network is shown in Figure 5.36.
However, much like the global rain gauge network, RADAR is

F IGURE 5.36 Illustrative map of precipitation over the U.S.


from composite mosaic of NEXRAD data.

primarily limited to the U.S., Europe, Japan, and other


developed countries.

F IGURE 5.35 Illustration of measured reflectivity (in dBZ)


from a single RADAR station (from wunderground.com).


To overcome the issues associated with ground-based
sensors (either gauge or RADAR) another method becoming
more common to hydrology is the use of satellite-based remote
sensing. Satellite-based sensors measure electromagnetic
radiation (either emitted by or reflected by the atmosphere
and surface) in various wavelengths. In principle, the same
concept used in NEXRAD can be used by mounting a
RADAR on a satellite, but to date only one such research
satellite (the Tropical Rainfall Measurement Mission
[TRMM]) exists. Instead, more common sensors are those
measuring passive microwave or infrared radiation emitted by
166

the surface and/or atmosphere. These sensors are deployed on


a variety of platforms including polar-orbiting satellites (with
passive microwave sensors), which sample a given region on
average once or twice a day, and geostationary satellites (with
visible and infrared sensors) that look down at a given region
more or less continuously. Many algorithms have been
developed to attempt to relate these measurements to
precipitation. One such product (the Global Precipitation
Climatology Project; Adler et al., 2003) is shown for a given
month in Figure 5.37. Based on the sensors used, the spatial
resolution can be quite coarse (tens to hundreds of kilometers)
and the temporal resolution is at best hourly. While these
methods can have a significant amount of uncertainty, they
are capable of capturing some of the key modes of variability

F IGURE 5.37 Example of daily precipitation map from


satellite-based remote sensing.

seen in precipitation, and most importantly provide global


coverage. Figure 5.37 shows the expected patterns of high
precipitation near the equator, low precipitation regions in the
areas of high-pressure zones near +/- 30 degrees and
secondary precipitation peaks at higher latitudes, with
notably absent precipitation in areas where high-pressure
zones are located.

In summary it is important to keep in mind that while
we may be interested in getting the same variable (i.e. MAP),
dierent sensors are really seeing dierent things (often at
dierent spatial and temporal resolutions) as illustrated in

F IGURE 5.38 Illustration of differing scale and methods of


sensing precipitation.

167

Figure 5.38. Estimates from dierent sensors will also have


their own error characteristics. Future operational estimates of
precipitation will likely be a combination of all these sensors
to take advantage of their strengths while minimizing their
weaknesses.

E XAMPLE 5.6.2 ( CONTINUED )

Z=

D N e
6

cD

dD = N 0

6!
c7

Substituting the expression for c into the reflectivity


equation yields:

E XAMPLE 5.6.2
If a cloud is characterized by the MarshallPalmer (MP) drop size distribution, derive a
reflectivity vs. precipitation relationship in the
form of Equation (5.6.7) in terms of the
Marshall-Palmer parameters. In the derivation
use the cloud-base precipitation (which is what is
typically measured by the RADAR) and assume
updraft velocities are relatively small.
The cloud base precipitation under the assumptions
mentioned above is given by (from Example 5.3.2):
1
P = Pb = 4 N 0 5
c

Z = N0

6!

4 N 0

7 5

P 7 5

720N 0

(4 N )

7 5
P
7 5

i.e. where:

Z = N0

6!

(4 N )

7 5

P 7 5

;
7 5

b=

= aP b

with:

a=

720N 0

(4 N )
0

7
5

which can be rearranged in terms of the MP parameter:


15

!
1$
c = # 4 N 0 &
P%
"

= 4 N 0

15

P 1 5

The reflectivity for the MP distribution is given by (from


Equation (5.6.6)):
168

S ECTION 7

MOD-WET Codes

Relevant functions based on concepts introduced in
this chapter include:
Computation of mean areal precipitation using the Thiessen polygons:

compute_mean_precip_from_thiessen.m
Computation of saturated adiabatic lapse rate:

sat_adiabatic_lapse_rate.m
Automated Thiessen polygon construction:

thiessen.m

169

S ECTION 8

Conceptual Questions
1. If an air parcel is lifted upward by 1 km (without
condensation), by how much will its temperature change?
Will it be cooler or warmer? What is the rate at which it
changes temperature with height called?
2. Is the moist adiabatic lapse rate generally larger or smaller
than the dry adiabatic lapse rate?
3. What type of surface air parcel characteristics (in terms of
humidity and temperature) will be expected to have the
highest cloud base when lifted?

9. Describe what the drop size distribution of a cloud


represents and how it can be used to determine various
characteristics of the cloud.
10. Describe the four basic mechanisms responsible for vertical
uplift in the atmosphere related to cloud formation/
precipitation.
11. Name the vertical uplift mechanism primarily responsible
for the precipitation climatology in the Sierra Nevada.
12. What mechanism/s are primarily responsible for uplift
(and hence cloud formation and precipitation) in the
midwestern U.S. Be specific in terms of season.
13. Where on the globe (latitudinally) is precipitation
generally largest? Explain why.

4. What is a CCN and describe its relevance to cloud


formation?

14. Where on the globe (latitudinally) are the worlds major


desert bands located? Explain why.

5. What mechanism is responsible for cloud droplet growth


early in a cloud life-cycle?

15. Name two commonly used weighted average methods for


estimating mean areal precipitation from gages.

6. What mechanism is responsible for cloud droplet growth


later in a cloud life-cycle?

16. Describe the basic principle behind the remote sensing of


precipitation by RADAR.

7. What is a warm cloud vs. a cold cloud?


8. What are the typical sizes (order of magnitude) of a CCN,
a cloud droplet, and a rain drop?

170

S ECTION 9

Sample Problems
Problem 5.1. Clouds typically contain a spectrum of drops
of varying diameters as a result of condensation and droplet
growth processes. The distribution of the number of droplets
within clouds as a function of drop size (diameter) is often
described by the Marshall-Palmer drop size distribution
(DSD).
a) Sketch the Marshall-Palmer DSD. According to this DSD,
what are the implied most probable and least probable drop
sizes? Is the location of the actual peak physically realistic?
b) Derive an expression for the total (integrated) number of
drops in a cloud as a function of the Marshall-Palmer DSD
parameter c. Based on your answer, what are the implications
on the total number of drops in a cloud as the average drop
diameter in the cloud increases?
c) Derive an expression for the total liquid water content in a
cloud (i.e. mass of liquid water per volume of air) assuming
spherical cloud droplets. This requires writing down the total
mass of drops of a given size, which is the product of the drop
size distribution and the mass of a single drop and then
integrating across all possible drop sizes. Hint: A useful
integral is:

n x
x
e dx =
0

n!
n+1

d) Compute the total liquid water content (g water m-3 air)


for a particular cloud with an average drop diameter of 0.4
mm.
e) Assuming the cloud is 3.5 km thick (and that the droplet
spectrum is uniform throughout the height of the cloud), how
much total liquid water (in kg m-2) is in the cloud?
f ) Based on your answer in part (e), what is the amount of
latent heat energy (in J m-2) associated with the vapor that
was condensed to form this cloud? Is this energy released into
or taken out of the atmosphere?
g) Assuming all of the cloud water ultimately fell and reached
the ground as precipitation, what would be the total amount
of precipitation in mm?
Problem 5.2. Assuming spherical droplets, the total amount
of precipitation (volume of water per unit ground area per
unit time) reaching the surface from a cloud is:

P=

Dmin

D3
X(D)N(D)
[VT (D) Vup ]dD
6

where N(D) is the drop size distribution, the second term is


the volume of a drop of diameter D, VT(D) is the terminal
velocity (which is generally a function of drop size), and Vup is
the updraft velocity.
171

a) Write the full expression for the precipitation flux for the
simplified case of no sub-cloud evaporation, negligible updraft
velocity and a Marshall-Palmer DSD.

d) For the most realistic case where sub-cloud evaporation


occurs, explain (qualitatively only) how you would expect the
precipitation rate to compare to the values above.

b) Using your expression from part a) compute the rainfall (in


both kg m-2 s-1 and mm hr-1) with a Marshall-Palmer DSD at
the cloud base with a parameter value: c = 22 cm-1. For the
terminal velocity, assume the proportionality constant is 3050
s-1.

Problem 5.3. The figure shown below illustrates a basin with


an established set of precipitation gages in the region. The
annual mean observed precipitation for each gage is listed in
the following table.

c) For the case of non-negligible updraft velocity (but still


negligible sub-cloud evaporation), what does the precipitation
rate equal? Assume the cloud is 1000 m thick and the updraft
velocity proportionality coecient is equal to 6.8 km1/2 hr-1.
Qualitatively explain your result and how it compares to your
answer above. [Note: First, you will need to determine the
lower limit of the integral for this case. You can think about it
by analyzing the equation (it will be a function of the model
parameters) or by plotting the integrand and identifying
which diameter corresponds to the case where the integrand
goes from negative (updraft bigger than terminal velocity) to
positive (terminal velocity bigger than updraft velocity).
Second, while the equation with the non-zero lower integration
limit can be solved analytically (via integration by parts), you
can (much) more easily evaluate it numerically, i.e. using
numerical integration, e.g. with the trapezoidal method
being one approach. In MATLAB, you can use the trapz
function to perform this numerical integration].

a) Compute the areal average annual precipitation using an


arithmetic average approach.
b) Compute the areal average annual precipitation using a
Thiessen polygon approach by constructing the polygons by
172

STATION

ANNUAL MEAN (CM)

1.1

1.7

1.3

1.8

1.9

2.3

hand and counting squares for each polygon area. Make sure
to be very precise in the construction of the polygons.
Consider only the area inside the basin for the weights.
c) Discuss the dierence in the two annual averages you find
for the watershed. Why are they dierent?
d) Which method do you think provides the most
representative mean areal precipitation estimate over the
basin? Why? So which method should generally be employed
for estimating mean areal precipitation?
Problem 5.4. Ground-based RADAR is now becoming more
commonly used in the U.S. to estimate precipitation. The
basis for this remote sensing technology is that microwave
pulses can be sent out into the atmosphere and are then
backscattered based on how many cloud/rain drops there are
in the field of view. From this backscattered power the rain

rate can be estimated. In this problem you will explore how to


estimate precipitation from RADAR backscatter.
For a RADAR system, the reflectivity Z (expressed in units L3
(i.e. often as mm6 m-3)) is a measurable quantity that if
related to precipitation can be used to estimate it remotely.
a) Using the Marshall-Palmer distribution, integrate the
above equation to express Z in terms of c and N0.
b) To relate the reflectivity to precipitation, you need an
expression for precipitation in terms of the same drop size
distribution parameters. For simplicity, assume the
precipitation is reasonably approximated by:

D3
P = N(D)
[VT (D)]dD
6
0
which corresponds to the case where updraft velocity is small
relative to terminal velocity and sub-cloud evaporation is
negligible.
Using the Marshall-Palmer distribution and assuming the
terminal velocity is directly proportional to diameter, i.e.:

VT (D) = D
along with the simplifying assumption that terminal velocity
is much larger than updraft velocity (VT>> Vup) integrate the
above precipitation equation to express precipitation (P) in
terms of all other parameters. Rewrite the equation to express
c as a function of all other variables.
173

c) Substitute your expression for c derived above into your


answer to part a) to express Z in terms of P and other
parameters in the following form:

Z = aP b
d) What is the theoretical value of the coecient b? What
terms make up the coecient a? For a cloud droplet terminal
velocity parameter of 3000 s-1, determine the constant a in
units of mm8/5 hr7/5. You should use the standard value for
the Marshall-Palmer distribution parameter N0.
e) Demonstrate that a=325 and b=7/5, if the units for Z are
mm6 m-3 and the input units of P are mm hr-1. In other words,
show that the following is true:

Z[mm 6 m -3 ] = 325(P[mm hr -1 ])7 5

Problem 5.5. Reflectivity measurements can be obtained


using Doppler radar to estimate the precipitation rate based
on the concepts described in the previous problem. The figure
below is a reflectivity map (dBZ) that was created using
Doppler radar for North Carolina on October 14, 2011 at 6:00
UTC. The circle on the map refers to the radars coverage
area.
a) Rewrite the equation derived in the previous problem so
that precipitation (in units of mm/hr) is a function of the
reflectivity in terms of dBZ.

b) Estimate the maximum measured reflectivity shown in this


map. Use the model you developed in the previous problem to
compute the estimated precipitation (in mm/hr)
corresponding to the observed reflectivity at this location.
c) The hourly mean areal precipitation, rate over a particular
domain for a given hour is often found to be much lower than
the maximum precipitation rate. What does that tell you
about the potential spatial variability of precipitation?
174

d) Based on the assumptions made in the development of the


Z vs. P equation above, would you expect your estimates to
be an underestimate or overestimate of the amount of
precipitation measured on the ground? Explain.

windward side of the island. For your calculation you may


reasonably assume that the weak trade winds lead to a very
small uplift velocity and that the high humidity tropical air
leads to negligible sub-cloud evaporation from the rain drops.

Problem 5.6. A tall tropical island receives persistent trade


winds that lead to orographically induced clouds and
precipitation. The wind is such that air parcels at the surface
are forced up the eastern (windward) side of the island and
back down to the surface on the western (leeward) side of the
island. The incoming surface air on the eastern side of the
island is generally 32C with a relative humidity of 90%.

d) Describe qualitatively what the expected climate on the


leeward (western side) of the island will be. Specifically, how
will the surface air temperature and relative humidity on the
leeward side compare to the surface air temperature and
relative humidity on the windward side? Justify how you
arrive at your answer.

a) Determine the dew point temperature of the surface air on


the windward side of the island.
b) Under the simplifying assumption that the dew point
temperature is constant with altitude and that clouds will
form when the surface air parcels are cooled to that dew
point, at what height will clouds form on the windward side of
the island?
c) The persistent clouds that form typically have a drop size
distribution consistent with the Marshall-Palmer distribution
with an average drop diameter of 0.3 mm. Assume that the
terminal velocity of a raindrop is given by:

VT (D) = D;

= 3000 s 1

Estimate the precipitation rate in mm/hr that occurs on the


175

Chapter 6

Snow
Processes

S ECTION 1

Learning Objectives
By the time you finish this chapter you should be able to:
1. Define the surface energy balance and name and define the
various fluxes involved
2. Define and be able to compute snow porosity, liquid water
content, snow density, and snow water equivalent

10. Estimate the phase and snow water equivalent state


based on the peak snow water equivalent and the amount
of cumulative energy that has been absorbed by the
snowpack
11. Understand and describe how variations in energy inputs
(especially due to topography) impact patterns of snow
and snowmelt over a basin
12. Understand and describe how vegetation impacts
snowpack evolution

3. Recall the order of magnitude density of new fallen snow


4. Describe and sketch the key components of the water and
energy balance of a snowpack (at a point in space) over
the accumulation and melt seasons
5. Name and describe the primary snowpack metamorphism
processes
6. Name and describe the three phases of snowmelt and
why/when they occur
7. Estimate the energy input needed to complete the
warming phase
8. Estimate the energy input needed to complete the
ripening phase
9. Estimate the energy input needed to complete the output
phase
177

S ECTION 2

Surface Energy & Mass


Balance

In moving from the previous chapters to the remaining
ones, we are shifting from the atmosphere to the surface and
sub-surface. One can think of the focus of the previous
chapters being largely on global hydrometeorology and the
primary forcing variables of surface hydrologic processes.
These are the net radiation, which drives the surface energy
budget, and precipitation, which drives the surface water
budget. The surface water and energy budgets are not only
relevant to snow processes, they are simply introduced here as
this is the first chapter dealing with the land surface. We will
see these again when discussing evaporation.

Net radiation at the surface is a continuous input in time
that can be either positive (i.e. a net energy input) or
negative (i.e. a net energy output). When it is positive, this
means the surface (either snow or vegetation/soil) is gaining
heat. The surface responds by dissipating this heat via several
mechanisms. If we consider an idealized surface (i.e flat/
homogeneous/infinitesimally thin with no heat storage; Figure
6.1) with net radiation input (Rn), the dissipation of energy is
accomplished via three primary mechanisms: a latent heat flux
(LE), which represents the energy associated with
vaporization of water/snow into vapor that is returned to the
atmosphere, sensible heat flux (H), which represents the

F IGURE 6.1 Schematic of surface energy balance (SEB) for an


idealized surface. Direction of arrows shown represent a positive flux. The opposite direction would be indicated by a negative flux.

energy conducted/convected between the surface to the


atmosphere, and ground (or surface) heat flux (G), which
represents energy conducted into the surface of the soil or
snow. For these fluxes to be in balance at the surface, we can
write:

Rn = LE + H + G

(6.2.1)


It is often said that the net radiation gets partitioned
into the other three fluxes on the right-hand-side. How it gets
partitioned is very much a function of the surface conditions
178

(i.e. wet/dry, warm/cool, etc.) The sign convention used in


this equation is that LE and H are positive when away from
the surface (into the atmosphere) and G is positive when away
from the surface (into the ground/snow). The sensible heat
flux is proportional to the dierence in temperature between
the surface and overlying air (will be discussed in more detail
in Chapter 8) and is positive when the surface is warmer than
the overlying air. For snow, especially in the melt season, the
overlying air temperatures can be greater than the snow
temperature, meaning that H is negative (provides a source of
energy that ultimately partly contributes to melting of the
snowpack). Recall that the net radiation is itself made up of
shortwave and longwave fluxes (Equation (3.8.3)), where the
outgoing longwave flux depends on the temperature of the
surface. In this way, one can think of the net solar input and
incoming longwave as the real energy inputs to the system
and the outgoing longwave, latent, sensible, and ground heat
fluxes as the surface response to this input. One should also
keep in mind that implicit in the surface energy balance (SEB;
Equation (6.2.1)) is a time and space dependence of each
term. We saw previously that the incoming radiation can be
greatly impacted by terrain, albedo, and other factors that
vary in space and vary diurnally and seasonally in time. These
variabilities will propagate to the other terms in the SEB.

The latent heat flux notation (LE) is actually shorthand
for the latent heat of vaporization (or sublimation in the case
of ice/snow) multiplying the evaporation/sublimation flux, i.e.

LE = LvE or LE = LsE

(6.2.2)

where E is the surface evaporation/sublimation in mass flux


density units (kg m-2 s-1). When multiplied by the units of the
latent heat of vaporization/fusion/sublimation (J kg-1), this
yields the appropriate units of W m-2. One can think of the
latent and sensible heat fluxes (as well as the outgoing
longwave flux) as cooling terms, as all three will act to reduce
the temperature of the surface. It turns out that evaporative
cooling (LE) is in fact the most ecient of these cooling
mechanisms so that if liquid water is prevalent at the surface,
a large fraction of the net radiation will go into vaporizing
water. The ground/surface heat flux (G), will conversely act
as a warming term that will conduct energy into the
underlying media (i.e. snow or soil) and warm it. It is
generally driven by temperature gradients between the surface
and sub-surface in the media. The temperature change of the
surface will depend directly on the magnitude of G and the
thermal heat capacity of the media. The ground heat flux is
generally into the surface media during the day (and summer)
when net radiation is high, and toward the surface during the
night (and in winter) when the net radiation is low.

In contrast to radiation, precipitation inputs are
inherently intermittent, i.e. they are positive for relatively
short durations (storms) and zero most of the time (interstorm periods). If one again considers an idealized surface, the
precipitation input gets partitioned into infiltration into the
soil and/or accumulation in the form of snow, evaporation/
sublimation from the surface, and runo. Writing equations
for mass balance can be done using consideration of the
relevant fluxes and Equation (1.5.2) as will be shown below.
179

In general, a precipitation event will directly lead to snow


accumulation (assuming solid precipitation), runo, and/or
infiltration into the soil depending on the state of the surface.
Evaporation/sublimation tends to be small during the
precipitation event (due to reduced net radiation), but may
play a large role during inter-storm periods. Mass balance will
be discussed in more detail in the context of snow
accumulation and unsaturated soil processes.

180

S ECTION 3

Snowpack Characteristics

In areas with sub-freezing temperatures, the processes
described in Chapter 5 can lead to solid-phase precipitation.
Such snowfall events generally require cold cloud processes
aloft along with sub-freezing air that extends all the way
down to the surface. Exceptions to this include hail storms
where there are cold cloud processes aloft (often due to
towering cumulus clouds), but above-freezing air in the subcloud layer. In such cases, the hail stones are generally large
enough (and fall fast enough) that the above-freezing subcloud temperatures are not sucient to fully melt them before
they reach the surface. In the context of this chapter we will
mostly be discussing the former case of snowfall leading to a
variety of snowpack processes. For a more detailed discussion
of snow processes, the reader is referred to Male and Gray
(1981) and Armstrong and Brun (2010).

Snowfall generally consists of snow grains (or
snowflakes), rather than droplets, that are often made up of a
large collection of tiny ice crystals (Figure 6.2). A snowpack is
simply an accumulation of snow grains on the ground. These
grains come in various sizes and morphologies depending on
the atmospheric conditions in which they were generated.
Figure 6.3 shows some typical types (habits) of snow grain
crystals as a function of atmospheric temperature. Unlike rain
droplets which are essentially the density of liquid water, the

F IGURE 6.2 Examples of different shapes of solid precipitation hydrometeors.

crystal nature of snow makes it such that the density of new


fallen snow is not that of solid ice, but considerably less (this
is most obvious in a dendritic snowflake where there are many
open spaces that will form pores in the snow volume). The
relative density of new fallen snow compared to liquid water
can range from less than 5% to upward of 30% and varies
based on the meteorological conditions (Figure 6.4). A
common assumption is that new fallen snow has a density of
approximately 100 kg m-3 (i.e. approximately 10% that of
liquid water). This is useful rule-of-thumb, but is most
certainly a generalization that can be erroneous.

As a result of its constituents, snowpack is a granular
porous media (i.e. one that contains pores that can be filled
with air/liquid water). Its porous nature, leads to the
181

that we are considering a unit area on the ground. The


snowpack volume (Vs) can be considered to be simply the sum
of the volumes of solid ice (Vi), liquid water (Vw), and air
(Va):

Vs = Vi +Vw +Va = dsA

(6.3.1)

where ds is the snow depth above ground and A is the unit


area. When the snowpack is cold (i.e. T < 0C) it is by
definition dry (Vw = 0) so that the snowpack is made up of ice
and air only. When the snowpack is melting (T = 0C), then
the pores will contain some liquid water. Based on the porous

F IGURE 6.3 Typical crystal shapes as a function of cloud temperature.

development of several definitions used to characterize


snowpacks. It should be noted that soil (which will be
discussed in Chapter 7) is also a porous media and hence
shares some of the same definitions. The key dierence is that
a snowpack is a highly dynamic media that evolves seasonally,
whereas soil is a relatively static porous media.

It is often convenient to think of a snowpack on a per
unit ground area basis when defining various characteristics.
So at a given moment in time, the volume of a snowpack is
often expressed in terms of its depth with the understanding

F IGURE 6.4 Illustration of varying snowfall (i.e. density/


grain size) as a function of meteorological differences.

182

structure, several snowpack characteristics are used and are


defined below. It should be kept in mind that for a snowpack
these properties change throughout the accumulation and
melt season. Moreover, they are usually a function of depth in
the snowpack (i.e. the properties vary with height).

The snow porosity is a measure of the volume of pore
space (some of which may be filled with liquid water) to the
total snow volume and is given by:

s =

Va +Vw
Vs

(6.3.2)

which is a dimensionless quantity. A snowpack made up of


new fallen snow may have a porosity as high as 90%, while a
solid icepack would have a porosity approaching 0%. The
liquid water content of a snowpack is given by:

Vw
Vs

(6.3.3)

where the liquid water content is zero for a dry snowpack and
positive for a wet snowpack.

The snow density is given by the ratio of mass of solid
ice and liquid water (where air mass is neglected) to the total
volume of snow, i.e.:

s =

M i + M w iVi + wVw
=
= i (1 s ) + w
Vs
Vs

i density of ice = 917 kg m -3


w density of liquid water = 1000 kg m -3
which can simply be thought of as a weighted average of ice
and liquid water densities by their respective volume fractions.
While mathematically the liquid water content could vary up
to the porosity, typically a snowpack cannot hold more than
5-10% of its pore space filled with liquid water. This upper
limit (i.e. liquid water retention capacity) is actually a
complicated function of the individual snowpack pore
morphology. Simple empirical expressions have been developed
in terms of snowpack density, e.g. (Eagleson, 1970):

ret

s2
s
4
= 0.0735 + 2.67 10 ;
w
w

(6.3.5)

[ ] = kg m -3

Finally, and perhaps most importantly, we can define the
variable that is typically of most interest to hydrologists,
namely the snow water equivalent (SWE). This is simply
defined as the equivalent amount of liquid water mass that
would result if the entire snowpack were melted and is given
by:

SWE =

( )Vs + (1 s )Vs i w
A

(6.3.6)

(6.3.4)
where the numerator represents the liquid water in the
snowpack (first term) plus the equivalent liquid water bound
183

up in ice (second term) and is normalized per unit area to


give dimensions of depth of equivalent water. The above can
be rewritten using the previous definitions as:

SWE = + (1 s ) i w ds = ds
w

(6.3.7)

which shows that the SWE is the snow depth multiplied by


the ratio of snow density to liquid water density. By construct,
this shows that in general SWE is less than or equal to the
snow depth.

As mentioned above, it is important to keep in mind that

snow tends to be a stratified media, especially when deep,


such that all of the above properties may vary with depth in
the snowpack. An example is shown for illustrative purposes
in Figure 6.5, where snow at three dierent sites (at two
dierent times: March and April) is characterized in terms of
density and temperature profiles. Mechanisms for some of the
trends (i.e. densification and warming between March and
April) will be discussed in more detail in the next section. In
such layered snowpack cases, the SWE may instead be
determined via the summation of the water equivalent of each
layer in the snowpack, i.e.:

SWE = ( s
n

w )dz n

(6.3.8)

where the n index simply corresponds to the property of a


given layer n, and dzn corresponds to the thickness of layer n.
E XAMPLE 6.3.1

F IGURE 6.5 Examples of variation in snow profiles as a function of location and time of season. Each panel corresponds to
a different site. Bars correspond to density on the top axis and
dotted lines correspond to temperature on the bottom axis.
Black and grey correspond to March and April conditions respectively (from Filippa et al, 2010).

The following snow measurements from a


homogeneous snowpack were taken: average snow
temperature of -15C, snow depth of 2 meters,
and an average snow density of 200 kg m-3. What
is the snow water equivalent (in units of cm),
liquid water content, and snow porosity of the
snowpack at the measurement time?
The snow water equivalent is given by:
184

E XAMPLE 6.3.1 ( CONTINUED )

! 200 kg m -3 $
SWE = #
&(2 m) = 0.4 m = 40 cm
-3
" 1000 kg m %
The liquid water content is known to be equal to 0 since
the snowpack is sub-freezing. Liquid water only exists in
a snowpack when it is at freezing. The snow porosity is
given by (rearranging Equation (6.3.4)):

s
200 kg m -3
s = 1 = 1
= 0.78
i
917 kg m -3
Hence the snowpack at the measurement time is 78%
pore space and 22% solid ice.

185

S ECTION 4

Snowpack Accumulation &


Metamorphism

Many seasonal snowpacks have two relatively distinct
seasons: the accumulation season (where SWE is increasing
with time) and the melt (or ablation) season (where SWE is
decreasing with time). While processes that occur in one
season can occur in the other, here we will separate the two
for conceptual purposes, highlighting the key processes that
occur primarily in each. Here we also first ignore the impacts
of vegetation, focusing on snowpack processes only and then
discuss how vegetation impacts these processes.

During the (winter) accumulation season, precipitation is
generally high, air temperatures are cold and decreasing, and
net radiation is small or even negative. The bulk mass balance
for a snowpack (i.e. with the entire snowpack as the control
volume) can be written as follows (i.e. from Equation (1.5.4)):

d(SWE)
= P E M
dt

(6.4.1)

F IGURE 6.6 Typical SWE time series during accumulation

season for a high-elevation location with seasonal snowpack (in


this case in the Sierra Nevada).

storage, where in this case the water is stored in the


snowpack. During the accumulation season E and M are
generally small and P is intermittent, so that SWE is mostly a
monotonically increasing function, which is relatively constant
between storms, and has sharp increases during a storm
(Figure 6.6). So to the extent that precipitation can be
accurately characterized (not very easy in mountainous
terrain), the accumulation season SWE is closely coupled to
the cumulative precipitation.
E XAMPLE 6.4.1

where P, E, and M are the precipitation (in this case mostly


snowfall but can also include wind-blown snow redistribution),
evaporation/sublimation, and melt output/runo respectively.
Simply put, as is always the case with mass balance, a surplus
of mass input (i.e. P > E + M) will lead to an increase in

Suppose a region has a clear accumulation season


with negligible melt and sublimation and three
storms each of which had cumulative precipita186

E XAMPLE 6.4.1 ( CONTINUED )


tion amounts of 10 cm, 15 cm, and 25 cm. What
is the SWE at the end of the accumulation
season?
The mass balance (assuming negligible sublimation and
melt) can be written as (from Equation (6.4.1)):

d(SWE)
= P SWE =
dt

P dt

2. Destructive metamorphism: This mechanism is the result of


relative instability of water at very high curvatures (in
grains). The saturated (equilibrium) vapor pressure for a
bulk surface (either liquid or ice) is given by the
appropriate Clausius-Clapeyron equation, but for curved
(ice) surfaces (of radius r) is given by:

2
ei (r) = ei ()exp

rRv wT

(6.4.2)

surface tension of water ( 7 10 2 Nm -1 )

Hence the SWE is simply the integral of the


precipitation events. During inter-storm periods the
precipitation is by definition zero, so the integral is
simply the summation of the three storm events. The
SWE is thus equal to 50 cm.


Density however is not as straightforward as SWE. New
fallen snow generally has a relatively low density (100-200 kg
m-3). As soon as snow begins to accumulate, gravity and other
metamorphism factors begin to take hold with the end result
being a densification (i.e. increase in density) over time
(Figure 6.7). The primary mechanisms at work include:
1. Gravitational settling: This mechanism causes a
densification of the snowpack as a result of the weight of
overlying snow, which compacts the snowpack, reducing the
porosity.

F IGURE 6.7 Typical snow depth (upper panel) and snow density (lower panel) time series during accumulation season for a
high-elevation location with seasonal snowpack (in this case in
the Sierra Nevada).

187

where the first term on the right-hand-side is just short-hand


notation for the saturated vapor pressure over a bulk surface
(i.e. the Clausius-Clapeyron Equation (2.4.6)). The second
term can be thought of as a correction for curvature (surface
tension) eects, with a small temperature (T) eect. The
result is a very high saturated vapor pressure for small radius
particles. This eect is the same one that prevents
homogeneous nucleation from occurring in clouds. A high
saturated vapor pressure drives vaporization, so that water
vapor molecules evaporate from these high curvature locations
and end up depositing on areas of less curvature. The end
result is a loss of small pointy grains and an increase in
larger more spherical grains, both of which tend to increase
the density of the snowpack. Because the mechanism occurs as
a result of small crystals, it typically only occurs in relatively
new snowpacks (i.e. with densities less than 250 kg m-3) or
near the top of snowpacks after new snow falls.
3.

Constructive metamorphism: This mechanism is also the


result of vapor movement and deposition, but is instead
associated with temperature gradients. As shown in Figure
6.8, temperature profiles are seldom uniform in the
snowpack during the accumulation season, where cold air
above the snowpack causes large fluctuations near the
surface, while the insulating capacity of the snow often
keeps the snow in contact with the ground near freezing. If
one neglects radius eects (i.e. after destructive
metamorphism has taken place) the presence of solid ice
leads to a vapor pressure which is saturated and given by
the Clausius-Clapeyron equation. Hence at each depth in

F IGURE 6.8 Schematic illustration of diurnal variation in


snowpack temperature profile.

the snowpack the actual vapor pressure is to a good


approximation given by:

L
e(z) = ei (T(z)) = es 0 exp s
Rv

1
1

T0 T(z)

(6.4.3)

188

where since ei is solely dependent on temperature, the


existence of a temperature gradient implies a gradient in
vapor pressure. As we will see in Chapter 8, a gradient in
vapor pressure implies an evaporative (vapor) flux, where the
flux goes from areas of higher vapor pressure to lower vapor
pressure. For a typical temperature profile (Figure 6.8), there
is generally a decreasing trend in temperature from the base
to the surface of the snowpack (with superimposed variability
due to diurnal forcings) that drive vapor from lower in the
pack to higher in the pack, or in general toward colder regions
of the pack. This process tends to drive grain growth and is
often the dominant densification mechanism in the

F IGURE 6.9 Example photographs of snow grains that have

undergone dry metamorphism. Most of the pointy ends have


disappeared due to vapor flux.

accumulation season. An example of grain metamorphism as a


result of constructive and destructive mechanisms is shown in
Figure 6.9. The net result is a rounding of grains over time.
4. Melt metamorphism: This mechanism does not take place
in the accumulation season (aside from intermittent events),
but typically during melt. Liquid water can be introduced
into the snowpack via melt near the surface or rain on
snow. This liquid water will often migrate downward into
the snowpack, re-freezing at a given depth where the
snowpack is sub-freezing. The water moves through the
open pores and so when it freezes it generally binds existing
grains together into a single larger aggregated grain. Small
grains are generally quickly merged into larger ones as a
result of this mechanism (Figure 6.10). Upon a re-freeze,
latent heat is released which goes into warming of the pack,
which can then accelerate constructive metamorphism.

In cases where a snowpack has distinct accumulation and
melt seasons, dierent mechanisms play more or less
important roles in dierent parts of the season. Gravitational
settling is ever-present, while destructive metamorphism plays
a role early in the accumulation season and near the snowpack
surface after snow events. Constructive metamorphism is most
important during the bulk of the accumulation season (when
temperature gradients are largest). During melt, the
temperature gradients tend to disappear as the snowpack
becomes isothermal, leaving melt metamorphism as the
dominant mechanism.

189

E XAMPLE 6.4.2
Suppose that during the accumulation season a 2
meter snowpack has a linear temperature profile
of: T(z) = mz; z 0
where m is the temperature lapse rate in the
snowpack, with a value of -2 K/m. Derive an
expression for the vapor pressure profile in the
snowpack and describe the implied vapor flux.
The vapor pressure is, to a good approximation, equal to
the saturated vapor pressure (with respect to ice) since
the air is in equilibrium with the solid ice matrix. The
vapor pressure gradient can thus be written using the
chain rule in terms of the temperature gradient:

de dei dei dT
=
=
dz dz dT dz
The first term on the right-hand-side is simply equal to
the Clausius-Clapeyron equation (Equation (2.4.4), but
for bulk ice) and the second term is equal to the
temperature gradient from the temperature profile, so
that the vapor pressure gradient is equal to:

F IGURE 6.10 Example of photograph of snow grain that has

undergone wet metamorphism. The larger grain is clearly made


up of what had been several smaller grains.

#L
Ls
de Ls ei T
=
=
e exp %% s
2
2 s0
dz Rv T z RvT(z)
$ Rv

#1
1 && T
%%
((((
$T0 T(z) '' z

190

E XAMPLE 6.4.2 ( CONTINUED )


which when substituting the specific profile is given by:

"L
mLs
de Ls ei
=
m=
e exp $$ s
2
2 s0
dz Rv T
Rv (mz)
# Rv

"1
1 %%
$$
''''
#T0 mz &&

Given the negative temperature gradient (i.e. negative


value for m) the vapor pressure is largest at the soilsnow interface and smallest at the snow surface. The
vapor flux will generally be from areas of high vapor
concentration to low vapor concentration. In this case,
the vapor flux will be from lower in the snowpack to
higher in the snowpack. This vapor flux will drive the
constructive growth mechanism of grains due to vapor
flux. Note: Once the snowpack becomes isothermal
during the melt season (i.e. m = 0) the vapor gradient
will also be zero.


Another key snow property related to metamorphism is
snow albedo. In general, snow has a very high reflectivity (in
the visible part of the spectrum) compared to almost all other
media, with albedos as high as 90% in some cases. Since
downwelling shortwave is the largest energy flux in many
cases, the albedo can strongly regulate the overall heat
absorption by the snowpack. For snow however, the albedo is
not a constant, but evolves, primarily due to grain size
evolution in the snowpack and addition of lower albedo
contaminants (i.e. dirt). Figure 6.11 shows an illustrative

example of the evolution of albedo as a function of age of


snow. Typically new fallen snow has the highest albedo as
small grains increase volume scattering. As the grains increase
in size, the reflectivity tends to decay. As such, albedo is often
modeled as a decaying function of time, where the time
variable is number of days since the last snowfall (Figure
6.11). Alternatively, physically-based models predict albedo
directly from grain size growth equations. When a new
snowfall occurs, the albedo generally jumps back up to the
maximum value before decaying again. The decay rate can

F IGURE 6.11 Typical models for snow surface albedo as a

function of time since last snowfall. General decay corresponds


to grain growth mechanisms.

191

also be a function of liquid water content of the snowpack.


Another factor that is sometimes accounted for in albedo
models (albeit often in a simplistic way) is the eect of
contamination of the snow. Things like dust, soot, and other
aerosols, tend to all add low albedo constituents into the snow
as the snow ages and tend to produce a decrease in the
albedo.

where ki is the thermal conductivity (in this case for ice). The
energy balance can then be written as:

(iciT )
= (Heat flux) + q
t

(6.4.4)


As described above, the metamorphism in the snowpack
is to a large extent driven by temperature profiles (and
specifically gradients). Since most of the transport of energy
(and vapor) occur in the vertical, the energy budget for the
snowpack can be written schematically (i.e. for a given layer)
as:

where the second term on the right-hand-side is simply using


the divergence operator (in general in three dimensions) and q
represents a source/sink (which also varies with z and t) and
is associated with melt (sink) and re-freeze (source) events
that consume or release energy respectively. If one only
considers the vertical flux convergence then this equation
simplifies to the well-known one-dimensional (1D) heat flow
equation:

Change in energy storage = Convergence of energy flux +


source/sink

(iciT ) 2(kiT )
=
+q
2
t
z

which can also be formally derived from Reynolds Transport


Theorem (Equation (1.5.1)). The energy storage in a
snowpack is related primarily to the specific heat capacity of
the medium (in this case ice, ci):

Stored Energy = iciT


where T is the temperature of the snowpack (a function of
depth z and time t). The conductive energy flux in many
media is given by a Fickian (diusion) type of behavior, where
the flux is proportional to the gradient in the state variable,
which in the vertical direction is:

Heat flux = ki (T z)

(6.4.5)

which is a second-order partial dierential equation for T(z,t)


that, provided initial and boundary conditions, can be solved
for the evolution of snowpack temperature. Of particular
importance are the boundary conditions, which at the surface
is given by the SEB, i.e. flux of energy being conducted into
the surface is given by:

G = Rn LE H = Rs(1 ) + Rl Rl LE H

(6.4.6)

while the boundary condition at the bottom of the snowpack


is the conduction of heat between the underlying soil and
snow.
192


From Equation (6.4.6), it becomes clear that the primary
heating terms of the snowpack are the net shortwave radiation
(where albedo plays a significant role), the downwelling
longwave radiation, and in cases where the air temperature is
warmer than the snowpack surface, the sensible heat flux. The
primary cooling mechanisms are the outgoing longwave
radiation, the latent heat flux, and in cases where the
snowpack temperature is warmer than the overlying air, the
sensible heat flux. These fluxes drive the energy balance of the
snowpack, which in turn drives the various metamorphism
processes described above (Figure 6.12). During the
accumulation season the net radiation is often negative (i.e.

the snowpack surface is emitting more radiation than it is


receiving). The net result is a sink of energy at the surface
and a general cooling of the snowpack (i.e. sometimes
significantly below freezing). Once spring arrives and the net
radiation becomes positive, there is a source of energy at the
surface, and the snowpack begins to warm. In the case of no
melt and no introduction of rain into the snowpack, q = 0 in
Equation (6.4.5). During the melt season, however the source/
sink term becomes important. The melt processes are
discussed in more detail in the next section.

F IGURE 6.12 Schematic showing some of the key inter-storm


processes in a snowpack.

193

S ECTION 5

Snowmelt

In an idealized snowpack, the snowmelt season is that
period generally following the accumulation season. In truth,
the two seasons are generally not completely distinct and
depend on the climatology of the region. At certain elevations,
snowpack is quite transient, where many accumulation and
melt events can occur throughout the winter and spring. In
any case, the melt period generally begins when the surface
net radiation transitions from negative to positive (Rn > 0).
Melt itself will not occur immediately when this transition in
net radiation occurs due to the thermal inertia of the
snowpack. At the time net radiation becomes positive, the
snowpack will generally be at sub-freezing temperatures. The
positive net radiation implies an energy input to the
snowpack. The specific impact this energy input has defines
three commonly described phases of the snowmelt season.

The first phase is the so-called warming phase. This
phase consists of the period of time when the positive net
radiation goes into increasing the temperature of the
snowpack toward the melting point of water. At any given
time, the amount of energy required to raise the average snow
temperature to the melting point is generally referred to as
the cold content of the snowpack, which can be defined as:

Qcc = ci wSWE(Tsnow Tm )

where ci is the specific heat capacity of ice (2102 J kg-1 K-1),


Tsnow is the average snow temperature and Tm is the melting
point of water (273.15 K or 0C). Keep in mind that by
definition during this period the snowpack is generally dry (no
liquid water content) except for transient rain-on-snow or
surface melt (which then percolates and re-freezes lower in the
snowpack).

Note that both SWE and Tsnow are time varying
quantities that describe the mass and energy state of the
snowpack at a given time. The temperature and SWE of the
snowpack are generally at a minimum and maximum at the
end of the accumulation season/beginning of the melt season
(i.e. warming phase) respectively. Keep in mind that in this
context the beginning of the melt season may be long before
the snow actually begins to change phase. The amount of
energy required to complete the warming phase is simply the
cold content at the beginning of the melt season which is
notationally referred to as: Qm1 for the energy input required
for melt Phase 1 (i.e. the warming phase). It is primarily
dictated by the specific heat capacity of the snowpack, which
by definition tells how much energy is required to raise the
temperature of the snowpack by 1 degree per unit mass.
Multiplying by mass and the total temperature change
required to reach the melting point provides the required
cumulative warming phase energy. In cases where the
snowpack properties (i.e. SWE and temperature) vary with
depth, the cold content can be computed for each layer and
summed up to get the total cold content. The length of Phase

(6.5.1)
194

1 is dictated by Qm1 and the amount of energy input to the


snowpack (driven primarily by net radiation).

additional water can be retained in the pores. The net energy


required to complete the output phase is given by:


The second and third phases of melt have to do with the
actual conversion of solid ice to liquid water. Keep in mind
that phase transition of a bulk media occurs at an isothermal
temperature (0C for water) so that during the last two
phases the temperature of the snowpack is constant
(isothermal) and at the melting point. By definition, the
snowpack is wet during these phases. The second phase is
referred to as the ripening phase and begins when the
snowpack becomes isothermal. Any cumulative energy input
beyond Qm1 will start to melt the snow. However early on in
the process, water is retained in the pores (i.e. none leaves the
snowpack as runo). The amount of water retained is that
given by Equation (6.3.5). The net energy input required to
complete the ripening phase (Phase 2) is simply:

Qm 3 = (SWE retds )wLf

Qm 2 = retds wLf

(6.5.2)

which is simply the mass of (retained) melted liquid water


multiplied by the latent heat of fusion (Lf = 3.34105 J kg-1).
Since the liquid water retention capacity is generally relatively
small (5% maximum), the length of Phase 2 is often relatively
short, depending on the net radiation being input to the
snowpack.

The third and final phase is the melt output phase. It is
simply the extension of actual melt once the snowpack is
ripe, and corresponds to the period where any additional
melt will actually flow through and leave the pack since no

(6.5.3)

where the first term on the right-hand-side is simply the


remaining amount of water that needs to be melted (after
ripening). Note that as expected, the total amount of energy
required for the second and third phases where melt is
actually occurring is given by:

Qm 2 + Qm 3 = SWE wLf

(6.5.4)

where SWE corresponds to the amount at the beginning of


melt (neglecting sublimation losses to SWE). The above
expressions are most easily understood in the context of a
single-layer snowpack. However when modeling a multi-layer
snowpack energy balance (i.e. Equation (6.4.5)) the picture
becomes more complicated. The snowpack can have melt in
the upper surface layers that then flows downward in the
snowpack where it re-freezes. The re-freezing releases latent
heat energy that warms the snowpack in that location. The
process of surface melt, percolation, re-freezing with latent
heating, will continue until the entire pack is isothermal at the
freezing point. At that point ripening and melt output will
occur with the addition of any more energy.

All of the above has been in reference to processes at a
given point in space. When considering basin-scale processes,
it is important to keep in mind how things vary. In particular,
SWE accumulation will vary depending on orographic and
other eects so that the distribution of SWE at the beginning
195

of the melt season will vary considerably across the basin.


Moreover, as discussed in Chapter 3, radiation inputs can vary
widely in complex terrain due to slope, aspect, sky view
factor, etc. Hence the length of the melt phase at dierent
locations in the basin will vary considerably. The end result of
these factors is that snowpack melt-out itself will vary
considerably across a basin.

are clearly shown to be a strong function of both month and


latitude, with snow accumulating first and melting out last at
higher latitudes. Elevation also is shown to play a key role in
both accumulation and melt.


It is important to keep in mind that the above described
accumulation and melt of snow is highly seasonal over much of
the globe. Movie 6.1 shows a monthly time series of images of
snow cover over the northern hemisphere over North America
(from January 2004-December 2004). Accumulation and melt

Suppose the snowpack properties described in


Example 6.3.1 are at the end of the accumulation
season. Compute the energy input requirements
for the warming, ripening, and melt output
phases. For simplicity, assume that the amount of
volumetric (liquid) water content that can be
held in the pores without drainage (i.e.
associated with ripening) is 0.02.

E XAMPLE 6.5.1

The required energy input for the warming phase is


equal to the cold content of the snowpack at the end of
the accumulation season, i.e.:

Qm1 = (2102 J/kg/K)(1000 kg/m 3 )(0.4 m)(15 K)


= 12.6 10 6 J/m 2
The required energy input for the ripening phase is given
by:

Qm 2 = (0.02)(2 m)(1000 kg/m 3 )(3.34 10 5 J/kg)

M OVIE 6.1 Illustration of seasonal (monthly) snow cover

over the northern hemisphere based on satellite measurements


over the period January 2004-December 2004 (from
sos.noaa.gov/videos/seasonal_blue_marble_400.mov).

= 13.4 10 6 J/m 2
The required energy input for the melting phase is given
196

E XAMPLE 6.5.1 ( CONTINUED )


by:

Qm 3 = (0.4 m (0.02)(2 m))(1000 kg/m 3 )(3.34 10 5 J/kg)


= 1.2 10 8 J/m 2
From these calculations we can see that the bulk of the
energy input requirements are associated with the actual
melt and less so for the warming/ripening. It is
important to note that these are the required energy
inputs. The actual source of the energy is the cumulative
energy inputs into the snowpack over the melt season.
The melt energy fluxes are typically expressed in units of
W m-2, which when integrated over time yields the same
units of J/m2. The energy inputs are mainly driven by
net radiation and sensible heating of the snowpack,
which are both typically increasing functions of time
moving forward into the melt season.

197

S ECTION 6

Impact of Vegetation

The impacts of vegetation on snow processes (and land
surface processes in general) are, among others, those related
to energy balance at the surface and those related to mass
balance. Each is discussed briefly below.

In terms of radiative energy balance at the snow surface,
the primary impact of vegetation is the attenuation and/or
augmentation of radiative energy fluxes. The clear- and
cloudy-sky shortwave fluxes discussed in Chapter 3 dealt with
those fluxes reaching the surface. In the case of bare soil/snow
these fluxes would directly contribute to the SEB. When
vegetation is present, those fluxes would correspond to those
reaching the surface just above the canopy. The downwelling
above-canopy shortwave radiation will interact with the
vegetation canopy, with the vegetation having media
properties as defined in Chapter 3 including transmissivity,
absorptivity, and reflectivity. The vegetation reflectivity
(albedo) is generally much lower than for snow. In the case of
a dense vegetation canopy (i.e. forest), the transmissivity will
also be relatively low. Together this means that much of the
above-canopy radiation will be attenuated (absorbed) by the
vegetation.

The interaction of solar radiation with canopy is
relatively complicated. To avoid complex radiative transfer
calculations, simple empirical expressions have been developed

to model the attenuation of downwelling solar radiation due to


vegetation canopy. In such models, the attenuation is modeled
as an additional correction factor in the downwelling
shortwave flux equation shown in Equation (3.6.9) so that the
downwelling flux reaching the snow surface would be:

Rs = fsv fsc "#ts + + ts + 2 $% Rs 0

(6.6.1)

where fsv is an attenuation function for shortwave radiation by


vegetation that depends on some vegetation characteristic. A
simple example of such a function is (Dunne and Leopold,
1978):

fsv = exp(3.91F)

(6.6.2)

where F is the forest cover fraction, which was specifically fit


to a lodgepole pine forest canopy a function of vegetation.
Note that this type of equation makes no attempt to account
for dierences in density/type of vegetation canopy, only
fractional cover. Other examples of such attenuation factors
use metrics such as the so-called leaf area index (discussed in
more detail in Chapter 8) to better capture the dierences in
attenuation by dierent kinds of canopies. Note that Equation
(6.6.1) is general in that if there is no forest cover (F = 0), it
reduces to the cloudy-sky case (which itself reduces to the
clear-sky case if there is no cloud cover).

The longwave radiation reaching the surface can also be
impacted by vegetation. Namely the longwave flux can
typically be augmented (increased) as a result of: i) a larger
198

emissivity relative to the eective atmospheric emissivity and


ii) a larger vegetation temperature as a result of absorption of
the atmospheric shortwave radiation. For the latter case, a
separate energy balance for the vegetation canopy must be
employed. Alternatively, an often-used simplifying assumption
is that the canopy has a negligible heat capacity which would
result in the canopy temperature being in approximate
equilibrium with the air temperature. In this case, only the
emissivity impacts would play a role in the longwave
augmentation. In such cases, the downwelling longwave flux
reaching the surface is often modified by using an emissivity
that is a weighted average of the atmospheric emissivity and
the vegetation canopy emissivity, i.e.:

d = (1 F)[flc (cloud)a ] + (F)[v ]


The impact of vegetation on mass balance of the
snowpack on the ground (or the surface in general) primarily
has to do with the so-called interception of precipitation by
the canopy. The leaves of a dense canopy can hold a nonnegligible amount of snow (or rain) that is at least initially
removed from the snow/soil surface mass balance. An example
of this in the case of snow in an evergreen forest is shown in
Figure 6.13. The interception is most often treated as a small
reservoir that captures any precipitation coming in contact
with the canopy and can hold up to a specified volume (or

(6.6.3)

where the d subscript refers to downwelling and the v


subscript refers to vegetation and the weighting is simply done
by the fraction of forest cover. Note that a common
simplifying assumption is that the emissivity of a forest
canopy is equal to 1.0 so that Equation (6.6.3) becomes:

d = (1 F)flc (cloud)a + F

(6.6.4)

Using such an approach, the downwelling longwave at the


surface is then given by:

Rl = dTa4

(6.6.5)

which is a generalized equation that reduces to the original


equations in Chapter 3 for the special case of no vegetation
cover and/or no cloud cover.

F IGURE 6.13 Photograph of intercepted snow by an ever-

green forest canopy (from twdef.usu.edu/photos/TWDEF%20Choice/


images/canopy_interception.jpg).

199

depth, i.e. volume per unit area) of water before the water
begins to drain or slough o (in the case of snow) the canopy
onto the surface. The amount of precipitation that falls
through the spaces in the canopy combined with the drainage
from the canopy is usually referred to as throughfall. The
primary impact of the intercepted water is that it can then
evaporate/sublimate back into the atmosphere before ever
reaching the surface. Hence, some of the water that would
have contributed to the snowpack and/or soil mass balance in
a bare surface case, will be removed from the surface mass
balance when vegetation is present. The evaporation (or
interception) loss from the canopy is discussed in more detail
in Chapter 8.

Both of the energy and mass balance implications
associated with vegetation cover discussed above play a
significant role in the spatial and temporal variability seen in
snowpacks. Moreover, they extend to non-snow covered
regimes, where the impact on energy/mass budgets are
equally important. This includes a significant impact on
evapotranspiration at the surface, which is discussed in more
detail in Chapter 8.

200

S ECTION 7

Snow Climatology

The climatology of snow is largely just a climatology of
precipitation, which drives snow accumulation, and a
climatology of radiation, which drives snow melt. Together
these will determine various climatological factors related to
snow including: average SWE, duration of accumulation and
melt seasons, types of snow, etc. Snowfall occurrence is largely
a product of latitudinal and elevational dependence. Areas at
high latitudes and/or at high elevation will experience air
temperatures cold enough to turn any appreciable
precipitation into snowfall.

Figure 6.14 shows the major mountain ranges across the
globe which experience significant seasonal snowfall and
accumulation as a result of elevational lapse rate eects
(despite being at midlatitudes). Much of the remaining
portions of snow covered areas of the globe are a result of less
radiation at higher latitudes. The combination of varying
elevation, latitude, and other factors leads to varying
climatological classification of snow types. A common
classification involves six dierent types of snow (Sturm et al.,
1995):
Tundra: Thin, cold, wind-blown snow usually found above or
poleward of tree line

F IGURE 6.14 Maps showing locations of major mountain


ranges across the globe.

Taiga: Thin to moderately deep, low-density snowpack found


in forests in cold climates
Alpine: Intermediate to cold, deep snowpacks, typically low
density
Maritime: Warm, deep snowpack with coarse-grained snow
due to wetting
Prairie: Thin (except in drifts) moderately cold snow with
substantial wind drifting
201

Ephemeral (or no snow): Thin, extremely warm snow that


melts soon after deposition

Figure 6.15 illustrate the spatial distribution and defined


types of snow across the Northern Hemisphere. The
classifications include factors like snow depth, density,
temperature, grain size, and vegetation presence/absence. For
example, in the continental U.S., Europe, and central Asia the
snowpacks are generally maritime, alpine, and prairie, while at
higher latitudes they transition to taiga and tundra. Based on
the snow processes described above, these varying snowpack
characteristics will impact the climatology of albedo and other
surface energy and mass balance processes.

The climatology of snow accumulation is tied mostly to
snowfall patterns. These climatologies are quite dicult to
determine quantitatively based on the relatively sparse

F IGURE 6.15 Map of Sturm snow classification of snow types


over the Northern Hemisphere.

F IGURE 6.16 Map of mean annual total snowfall over the continental U.S. (in inches).

202

patterns in this map are highly correlated with the snowfall


map shown in Figure 6.16 with both latitudinal and
elevational dependence. The map shows general areas in the
mountainous regions with snow cover presence greater than 60
days, up to values higher than 150 days. In localized areas
with north-facing aspects and/or with significant vegetation
cover snow may persist even longer. Dierences are also due to
radiation (and other energy) inputs, as those at lower
latitudes and nearer coastlines may be subject to less heating
during the melt season or more abrupt transitions between
snow and rain season, where rain-on-snow events can often
lead to rapid melt-out.

F IGURE 6.17 Contour map showing the average annual number of days with snow cover.

availability of in situ data in many regions of the globe.


Figure 6.16 shows the annual snowfall climatology estimate
from NOAA over the continental U.S. The patterns seen are
essentially a superposition of the two eects described above:
a general increase in amounts of snowfall at higher latitudes
(most recognizable in the eastern U.S.) combined with areas
of high snowfall localized in mountainous regions (most
recognizable in the western U.S.).

The snowfall amounts combined with radiation inputs
during the melt season will, to first-order, determine how long
seasonal snowpacks last on the ground. Figure 6.17 shows a
map of snow cover duration over the continental U.S. The
203

S ECTION 8

Snow Measurement

In describing snow measurement one must distinguish
between snowfall measurement and snowpack measurement.
The former attempts to measure precipitation in the form of
snow. Hence, the same techniques discussed in Chapter 5 can
theoretically be used to measure snowfall, including snowfall
gauges, RADAR, and satellite-based techniques. The primary
complication with snowfall gauges have to do with the fact
that wind eects can be more amplified for low density
snowflakes (compared to raindrops) in the sense that the
turbulence caused by the presence of the gauge itself can
much more easily deflect snow such that it is not captured by
the gauge. Hence gauge under-catch, which is prevalent in the
case of rain, is generally exaggerated in the case of snowfall
measurement. It is not uncommon for a gauge to catch only
50% of the snowfall that might actually be falling, which can
lead to large (negative) biases in snowfall estimates.
Additionally, snow can freeze or block the gauge opening, so
often gauges designed to measure snowfall are equipped with a
heating unit to melt the snow which can then flow into the
tipping bucket or water reservoir.

The RADAR-based techniques used for snowfall
estimation also have diculty, mainly caused by the much
more complicated relationship between snow hydrometeors
and reflectivity. While liquid raindrops are usually assumed to

be spherical hydrometeors, such an assumption in the case of


solid precipitation can be quite erroneous. Moreover, the
disparity of the habits of snow crystals (Figure 6.3), means
that the uncertainty in the relationships between hydrometeor
size/shape and reflectivity is much larger. Similarly high
uncertainties plague relationships in satellite-based estimates
of snowfall.

The other type of measurement of snow is the amount of
snow on the ground at a given location or time. In some sense,
this is generally what is of most interest and is essentially, at
least in the accumulation season, an integrated measure of

F IGURE 6.18 Example of snow depth measurement (from National Weather Service, Wilmington Ohio).

204

how much snowfall occurred. Several techniques are used to


measure snow in situ, some of which are manual. These
manual measurements include snow courses, where snow
depth is measured using a snow stake (Figure 6.18) or SWE is
measured using a cylindrical probe that removes a column of
snow that is then weighed (Figure 6.19). Such snow courses
often require trekking into remote mountainous regions. As
such, they are often performed only monthly near the peak of
the snow accumulation season and during the melt season.
Such networks of long-term sampling often form the basis for
spring snowmelt and runo planning in many snow-dominated
regions, including California which has had the California

F IGURE 6.19 Example of SWE measurement via weighing


snow cores (from nationalatlas.gov/articles/climate/a_snow.html).

Cooperative Snow Survey (http://cdec.water.ca.gov/snow/) in


place dating back to the early 1900s.

In order to increase the temporal resolution of SWE
measurements and avoid the manual labor required for snow
surveys, more recently snow pillows have become popular for
essentially continuous (daily) measurement of SWE. A
schematic of such a snow pillow station is shown in Figure
6.20 with a picture from a real site shown in Figure 6.21. A
snow pillow is a sealed envelope filled with a non-freezing
liquid that rests on the ground and is connected to pressure
transducers. The basic idea is that accumulated snow on top
of the pillow will impart a compressive force on the snow
pillow that will produce a measurable change in pressure in
the fluid. These sites often have some meteorological sensors
at the site to measure air temperature, precipitation and other
variables. A standardized network of snow pillow sites referred
to as snow telemetry (SNOTEL) sites is coordinated across
the western U.S. by the Natural Resources Conservation
Service (of the USDA).

Figure 6.22 shows a map of the SNOTEL sensor network.
The telemetry aspect of a SNOTEL site has to do with the
sending of the logged data to a remote receiver, which
alleviates the need to go to the site for data collection. Other
state agencies have additional snow pillow sites that are also
often telemetered, but not part of the ocial SNOTEL
network. The end result is a daily time series of SWE during
the snow season. Because of their cost, snow pillow sensors are
generally less abundant that the number of snow course
measurements, but they nevertheless provide a useful
205

F IGURE 6.21 Photo of a snow pillow site prior during the dry
season.

variability in space. Given this level of heterogeneity, the


current networks generally under-sample the SWE fields,
especially in mountainous terrain.

F IGURE 6.20 Schematic illustration of a snow pillow site


(from nationalatlas.gov/articles/climate/a_snow.html).

complement to the other measurements. It is very important


to keep in mind that these and the other standard
measurements described above are all point-scale
measurements of variables that have a high degree of


Another in situ point-scale measurement that is
impractical for operational implementation, but often used for
more detailed measurement of snowpack stratigraphy is the
digging of snow pits. This technique involves digging a pit to
ground level to expose a face of the snowpack (Figure 6.23).
Once exposed, various profile measurements including
temperature, density, SWE, hardness, grain size, etc. at a
206

F IGURE 6.23 Example of snowpack characterization via a


F IGURE 6.22 Locations of SNOTEL sites in the Western U.S.

snow pit where measurements of temperature, density, hardness, grain size profiles are made.

207

regular interval across the entire profile are taken from the
face of the snow pit.

While the remote sensing of snowfall is dicult as
described above, other aspects of snow are routinely measured
via satellite-based remote sensing. Satellite-based remote
sensing generally measures electromagnetic radiation in a
particular and relevant part of the spectrum. Snow generally
has two relevant wavelength bands that provide useful
information: visible/near-infrared (Vis/NIR) and microwave.

essentially measuring the reflected solar radiation from the


surface and can therefore identify the presence or absence of
snow relatively easily (except for the case of snow under dense
vegetation canopies). Figure 6.24 shows a map of snow cover
over the globe for a particular day during Northern
Hemisphere winter obtained from polar orbiting satellites. The
data on sensors measuring in visible/near-infrared is often


In the Vis/NIR (i.e. shortwave) part of the spectrum
snow is highly reflective compared to other media (Table 3.2).
Satellite sensors measuring in this part of the spectrum are

F IGURE 6.25 Example of satellite-derived SWE estimate


F IGURE 6.24 Satellite-derived snow covered area.

over the northern hemisphere from passive microwave measurements.

208

available at spatial resolutions ranging from tens to hundreds


of meters with varying temporal resolution (daily to every
couple weeks).

In the microwave spectrum, satellite sensors measure the
emitted (passive) microwave radiation from the surface, which
is sensitive to the amount of snow mass on the ground. So
microwave measurements are more directly connected to SWE
than Vis/NIR and retrieval algorithms have been developed to
estimate SWE from space-borne microwave sensors (Figure
6.25). However, the spatial resolution of microwave
measurements are on the order of tens of kilometers (i.e. much
coarser than Vis/NIR). At these coarse scales, there can be
significant sub-grid heterogeneity within a single remote
sensing footprint. This can often pose problems for the
retrieval of SWE in mountainous terrain.

209

S ECTION 9

MOD-WET Codes

Relevant functions based on concepts introduced in
this chapter include:
Disaggregation of meteorological station forcings:

distribute_met_forcings.m
Computation of surface pressure over complex terrain:

disaggregate_press.m
Computation of near-surface specific humidity over complex terrain:

disaggregate_qair.m
Computation of air temperature over complex terrain:

disaggregate_Tair.m
A simple 1-layer snow model with mass/energy balance:

snow_model.m
A snow albedo model based on the USACE formulation:

albedo_usace.m

210

S ECTION 10

Conceptual Questions
1. Name the four fluxes in the surface energy balance
equation. What are the units on the fluxes?
2. Describe what the porosity of the snowpack represent.
3. What is the liquid water content of a dry snowpack. What
(order of magnitude) is the maximum liquid water content
of snowpack?
4. Describe the meaning of the liquid water retention capacity
of a snowpack.

10. Describe the key ways in which vegetation impacts


snowpack processes (and surface processes in general).
11. Name and describe the types of snowpack used in the
Sturm classification system. What are the primary types of
snowpack in the Western U.S.
12. Name the commonly used in situ techniques for
characterizing snowpack properties and clearly note what
properties are provided by each.
13. Describe the basic principle behind a snow pillow for
characterizing SWE.
14. In what ways can snowpacks be characterized from remote
sensing?

5. What is the order of magnitude density of new fallen snow


to solid ice?
6. Write the basic mass balance equation of a snowpack. Be
careful to define each term.
7. Name and describe the primary snowpack metamorphism
mechanisms. What impact do they generally have on
snowpack density?
8. How does snowpack albedo generally change with time?
What is responsible for this change?
9. Name and describe the three main phases of the snowmelt
process. What snowpack properties are chiefly responsible
for the amount of energy required for each phase?
211

S ECTION 11

Sample Problems
Problem 6.1. Measurements of snowpack in three locations
in a small basin yield the following measurements:
LOCATION

DEPTH (CM)

95

102

105

SWE (CM)

27

27

29

TEMPERATURE
(C)

-3

-5

a) Determine the snow density and porosity at each location.


b) What can you say, if anything, about the liquid water
content in the snowpack at each location. Using the empirical
Equation (6.3.5), estimate the liquid water retention capacity
in the snowpack at each location.
c) Compute the cold content of the snowpack at each
location. What do your answers imply about the phase in
which the three dierent snowpacks are in?
Problem 6.2. The temperature profile in a snowpack is
measured as follows:

DEPTH ABOVE SOIL


(CM)

TEMPERATURE
(C)

-1

50

-3

100

-2

a) Compute the saturated vapor pressure in the pore space at


each level in the snowpack. You can reasonably assume
curvature eects are negligible so that you can use the bulk
Clausius-Clapeyron equation for ice.
b) Snow grain size evolution is largely driven by vertical
temperature gradients which are responsible for vertical vapor
gradients. Compute the vapor gradients between the first and
second and second and third levels in the snowpack. The
gradient is just the dierence in vapor divided by the distance
between the two locations (i.e. dierence in height).
c) Assuming vapor transport will be from areas of high vapor
concentration to areas of low vapor concentration (something
that will be discussed in more detail in Chapter 8 in the
context of evaporation), describe the expected transport of
vapor in the snowpack. In other words, which level will vapor
be transported to and what levels will vapor be transported
away from? What is the expected impact on grain size in the
three layers? Explain.
Problem 6.3. The end of the snow accumulation period in a
small watershed for a particular year is April 1st. Suppose the
212

snow accumulation is uniform over the basin, and the


following April 1st measurements were taken: a snow
temperature of -5C, snow depth of 1.2 m, and a snow density
of 450 kg m-3. Assume that April 1st is the very beginning of
the melt season and that the amount of volumetric (liquid)
water content that can be held in the pores without drainage
(i.e. associated with ripening) is 0.02.
a) What is the snow water equivalent, volumetric (liquid)
water content in the snowpack, and snow porosity on April
1st?
b) Compute the energy input required to respectively
complete the warming, ripening, and melt output phases.
What fraction of the total energy required to complete melt
goes into each phase?
c) Incoming radiation can vary greatly throughout a basin
due to topography (slope/aspect/shading). Due to such
heterogeneity, for the first thirty days after April 1st, the
snowpack at two locations in the basin absorb an average of
15 W m-2 (location #1) and 75 W m-2 (location #2).
Determine the state of the snowpack (snow water equivalent,
phase (dry/wet), snow temperature) at each location after the
thirty days. For simplicity, assume the amount of liquid water
retained in the snowpack is constant from the end of the
ripening phase until all water is melted.

temperature and relative humidity are: 5C and 75%


respectively. Suppose the only dierence between the two
locations in the basin is the vegetation cover. One location is
bare and the other is covered by 50% forest cover.
a) Estimate the dierence in downwelling shortwave radiation
reaching the snow surface for the two dierent locations.
What impact is the dierence expected to have on the
snowpack? Explain.
b) Estimate the dierence in downwelling longwave radiation
reaching the snow surface for the two locations. What impact
is the dierence expected to have on the snowpack? Explain.
c) Do the dierences in shortwave and longwave reinforce or
counteract each other?
Problem 6.5. A snowpack will generally begin to first melt
at the surface where the net radiative forcing is largest. This
initial melt may lead to percolation in the snowpack that then
refreezes lower in the snowpack. Suppose conditions are such
that a snowpack experiences 2 cm of SWE melt at the surface.
The melt water percolates downward to a layer with a
temperature of -1.5C where it re-freezes. After re-freezing,
what is the temperature of the snow at that location (i.e. as a
result of latent heating). Will all of the water re-freeze?

Problem 6.4. Suppose the regional downwelling shortwave


(during clear-sky conditions) at a given time over a given
location is: 500 W m-2 and that the reference-level air
213

Chapter 7

Unsaturated
Flow and
Infiltration

S ECTION 1

Learning Objectives
By the time you finish this chapter you should be able to:
1. Define the meaning of a porous media
2. Use the soil triangle to define a particular soil type/
texture
3. Define the volumetric water content and relative soil
saturation and how they are related to each other
4. Describe what factors (i.e. pressure head, elevation head,
velocity head) control flow in porous media
5. Define matric head and hydraulic conductivity and how
they depend on soil moisture and soil type
6. Define and explain Darcys Law for flow in unsaturated
conditions
7. Understand and describe the two terms that in particular
control vertical flow in unsaturated soils

10. Sketch a picture of how the soil moisture profile in the


unsaturated zone evolves during an infiltration event
11. Sketch a picture of how the infiltration rate evolves in
time during a storm for the general case of switching
boundary conditions
12. Sketch a picture of how the infiltration rate evolves in
time for the limiting cases of very intense storms (i.e.
precipitation rate very high relative to saturated
conductivity) and low intensity storms (i.e. precipitation
rate lower than saturated hydraulic conductivity)
13. Understand the primary assumptions used in models for
infiltration capacity (e.g. Philip, Green-Ampt)
14. Understand and describe the dierence between actual
infiltration and infiltration capacity (or potential
infiltration)
15. Convert between infiltration rate and cumulative
infiltration
16. Apply the time-compression approximation to compute
the actual infiltration under non-ponded conditions (i.e.
the general case)

8. Write down Richards equation for one-dimensional


(vertical) flow in unsaturated soils
9. Understand and describe what drives infiltration rate

215

S ECTION 2

Unsaturated Zone Characteristics



Soil is a porous media (like snow) consisting of a matrix
of individual solid grains and interconnected pores that can be
filled with water and/or air. The unsaturated (or vadose) zone
is the region of soil found between the land surface and the
saturated zone (water table) that lies at some depth beneath
the surface (Figure 7.1). The main distinguishing feature
between the saturated and unsaturated zones simply have to
do with whether the pore space in the soil is partially or fully
filled with water. The unsaturated zone is actually more
complex in that the pores are filled with both water and air.
The thickness of the vadose zone is variable throughout a
basin, where generally speaking the water table depth is
deepest at higher elevation regions and shallowest along
streams in the basin. If a particular basin is vegetated, the
rootzone of the soil is primarily contained within the vadose
zone. For a more detailed description of vadose zone processes
the reader is referred to Parlange and Hopmans (1999).

In terms of fluxes, the primary source of water to the
vadose zone is infiltration, which by definition is the flux at
the surface (associated with precipitation events), while water
can also move upward from the water table below via
capillary rise. In general, percolation is the flux of water
moving down through the vadose zone and recharge is the flux

F IGURE 7.1 Schematic illustration of porous media including


unsaturated and saturated zones (from quizlet.com/4706106).

leaving the vadose zone and entering the groundwater system.


Another distinguishing feature of the vadose zone is that the
water movements are mostly vertical, which is in contrast to
the groundwater system where fluxes are mostly horizontal.
As such, flow dynamics in the unsaturated zone are often
treated as simply one-dimensional (1D) in the vertical.
Unsaturated flow dynamics consist not only of water moving
downward through the vadose zone, but also upward via
216

redistribution and evaporation from the soil surface.


Additionally, root uptake (and subsequent transpiration from
the leaves of the vegetation) constitutes a sink of moisture
that can be distributed throughout the rootzone.

or relative saturation:


In terms of characterizing porous soil media, many of the
same or analogous definitions that were used for snow are
used for soil. The mineral particle density is given by:

both of which are dimensionless. The former is a


representation of the fraction of total soil volume filled with
water and therefore can theoretically vary between zero and
the soil porosity. The s subscript is used to denote saturated
(i.e. all pores are filled with water). The relative saturation is
a normalized metric that represents the fraction of pores filled
with water and therefore varies between 0 and 1. When
discussing soil moisture it is crucial to know which of these
two metrics are being used as they are approximately a factor
of two dierent due to the porosity. In reality, the lower limit
of soil moisture is not generally zero, but a residual amount
that is bound to the soil grains, which is dependent on the
type of soil.

m =

Mass of minerals
Volume of minerals

(7.2.1)

and is relatively invariant for soils with a value around 2650


kg m-3. The bulk density is given by:

b =

Mass of minerals
Volume of soil

(7.2.2)

and varies depending on the soil type and porosity. The soil
porosity has the analogous definition used in snow (in this
case volume of pore space per unit volume of soil) and can be
related to the above densities:

s = 1 b m

(7.2.3)

The water content in the unsaturated zone is usually


characterized by the volumetric water content:

Volume of water
Volume of soil

(7.2.4)

s=

(7.2.5)


Most of the above defined properties depend on soil
type. However this has not yet been rigorously defined. The
common way of defining soil type (or texture) is via a grain
size distribution. This involves taking a soil sample and
sorting it with progressively smaller sieve openings to
characterize the distribution of soil grains in various size bins.
For the types of soil commonly found in the vadose zone,
three end-member types are defined: sand (size range between
0.06 and 2 mm), silt (size range between 0.002 and 0.06 mm)
and clay (sizes less than 0.002 mm). For a given soil sample
one can determine the fractional weight in each of these end217

silty clay loam. The important aspect of this is that many


experiments have been done to correlate these defined soil
types with specific properties (including those defined above)
that are useful for characterizing the storage and/or flow
properties of a given soil.

Table 7.1 shows the typical porosity for the various soil
types as determined via analysis of over 1000 soil samples by
Clapp and Hornberger (1978). The general trend is that sandy
soils have the lowest porosities, while clay soils have the
highest. This may be counterintuitive at first, as larger

T ABLE 7.1. A VERAGE POROSITY AS A FUNCTION OF


SOIL TYPE (C LAPP & H ORNBERGER , 1978). V ALUE
SHOWN IN PARENTHESES IS THE STANDARD
DEVIATION .
SOIL TYPE

F IGURE 7.2 USDA Soil Conservation Service (SCS) soil tex-

ture triangle for determining soil type from fraction of clay, silt,
and sand.

members. Based on this classification, one then typically uses


the so-called soil texture triangle (Figure 7.2) to identify the
particular soil type, which are arbitrarily defined names
indicating the combination of constituents (silt loam, sandy
loam, sandy clay, silty clay, etc.). For example a soil sample
with 30% clay, 10% sand, and 60% silt would be classified as a

POROSITY

Sand

0.395 (0.056)

Loamy sand

0.410 (0.068)

Sandy Loam

0.435 (0.086)

Silt Loam

0.485 (0.059)

Loam

0.451 (0.078)

Sandy clay loam

0.420 (0.059)

Silty clay loam

0.477 (0.057)

Clay loam

0.476 (0.053)

Sandy clay

0.426 (0.057)

Silty clay

0.492 (0.064)

Clay

0.482 (0.050)
218

grained soils generally have larger pores. However, clay tends


to have smaller but many more pores due to the configuration
of the particles (Figure 7.3). The implication of this is that
while clays theoretically have more water storage capacity, the
ability for water to actually flow is much higher in sandy soils.
This will be seen more explicitly in following sections.

In terms of connections between porosity and soil
moisture, Figure 7.4 shows the range of soil moisture
conditions one may expect to find in a soil. During and after a
storm, the soil may be near-saturated, in which case it is easy
for water to drain from the soil due to gravity. After the
immediate gravitational drainage, the state where most of the

F IGURE 7.4 Illustration of soil moisture content in a porous


media.

F IGURE 7.3 Illustration of porosity for sand vs. clay soil.

pores are still filled is called the field capacity. The typical
situation is shown where a mix of air and water is stored in
the pores, where water may be held up against gravity due to
capillary forces (discussed more below). Finally, if additional
drying is experienced (i.e. by evaporation or root uptake), the
soil moisture condition will eventually reach the wilting point.
In this case, only immobile water that is essentially adhered to
grains is left behind. In this condition, a plants system will
not be able to uptake moisture and may wilt. The amount of
water corresponding to each of these states is dependent upon
the soil type as shown in Figure 7.5. The reason for this has
219

E XAMPLE 7.2.1
A soil sample taken from the field with a 10 cm
long by 5 cm diameter cylindrical tube has a field
weight of 306 g and an oven-dried weight of 245
g. For the soil sample, calculate the: a) bulk
density and porosity and b) volumetric soil
moisture and relative saturation.
a) The volume of the (cylindrical) soil sample is given
by:

Vsoil =

(5 cm)2(10 cm) = 196.4 cm 3


4

The bulk density is given by the ratio of the dry soil


mass to the soil volume:

F IGURE 7.5 Estimates of field capacity and wilting point soil


moisture values as a function of soil type.

b =

245 g
= 1.25 g cm 3
3
196.4 cm

The soil porosity is given by:


to do mostly with flow through soils rather than storage and
is discussed in the next section.

s = 1 (1.25 g cm 3 ) (2.65 g cm 3 ) = 0.53


b) The volumetric soil moisture and relative saturation
depend on the amount of water in the soil. The mass of
water in the sample is 306 g - 254 g = 61 g. Based on
the density of liquid water this amounts to 61 cm3. The
volumetric soil moisture is then:

220

E XAMPLE 7.2.1 ( CONTINUED )

61 cm 3
=
= 0.31
196.4 cm 3
The relative saturation is then given by:

s=

0.31
= 0.59
0.53

221

used here is gauge pressure, which means pressure above


atmospheric pressure. In general this energy (per unit weight)
has dimensions of length, can vary with space and time, and
will dictate flow direction and magnitudes. In general for
subsurface soil flow, the velocity (V) is small making the
kinetic energy term negligible:

S ECTION 3

Flow in Unsaturated Porous Media



Here we introduce the basics of flow in porous media.
The principles apply to both unsaturated and saturated flow,
but will be applied for unsaturated flow first. The key point
with respect to flow is that water flows down the prevailing
energy gradient. The total energy at a given point in a flow
field (neglecting internal energy) consists of the sum of
potential energy, kinetic energy, and work done by the
pressure force, and can be written as:

1 WV 2 pW
Total Energy = Wz +
+
2 g
w g

(7.3.1)

In the case of water flow, the density corresponds to that of


liquid water. It is then common to divide the above by weight
(W) and define the hydraulic head, which is simply the total
energy per unit weight:

V2
p
H =z+
+
2g w g

(7.3.2)

where the terms on the right-hand-side are referred to as the


elevation head, velocity head, and pressure head. The
elevation head is defined relative to a datum. It should be
noted that in the context of flow problems, the pressure p

H z+

p
= h(x,y,z,t)
w g

(7.3.3)

where h is the piezometric head and the spatial and temporal


dependence is shown for emphasis. As will be discussed below,
spatial gradients in h are what drive flow, where in particular
water will flow from areas of higher h to areas of lower h. It
should be noted that implicitly each term is representative of
the average state of the system over some small representative
elemental volume (REV). The REV can be thought of as a
small dierential volume of soil (containing soil particles, air,
and water). So for example, the pressure is the average water
pressure over the REV.

The above principles apply to either saturated or
unsaturated flow, however the pressure head term in the case
of unsaturated flow has unique connotations. In saturated
soils one can envision water being pushed by pressure
gradient forces, where the pressure is higher than atmospheric
surface pressure. In unsaturated soils the water is being held
up against gravity, or pulled by pressure gradient forces that
are less than atmospheric surface pressure. Because pressure
in this context is defined as gauge pressure, this means that
222

pressure in the unsaturated zone is negative. Such negative


pressures are often referred to as suction or tension. The
reason for the suction forces have to do with capillarity and
surface tension in the soil.

One can envision soil as containing several dierent pore
sizes that could be conceptualized as capillary tubes of
varying diameter. If dierent capillary tubes are placed in a
beaker of water that is open to the air (Figure 7.6), the water
level inside the capillary tubes will rise above the water level
outside the tubes. Moreover, the smaller the tubes, the higher
the water level will rise. The dierence in height between the
water inside and outside of the tubes is due to suction
(negative pressure associated with surface tension) where the
pressure at the water surface outside the tubes is by construct

atmospheric (zero gauge pressure). When a soil is


unsaturated, water will be held more tightly against gravity in
the smaller pores in the same way that water is pulled higher
in the tubes with smaller diameters (Figure 7.6). The result is
that water in smaller pores will have higher (more negative)
suction pressures. Since smaller pores have higher suction,
these pores will tend to hold water longest. Furthermore, since
the pressure p in Equation (7.3.3) refers to the average over
an REV of soil, high (i.e. more negative) suction pressures are
generally associated with dry soils and vice versa. In the limit
of a soil becoming completely dry, the suction pressure will
approach minus infinity. In the limit of a soil becoming
saturated, the suction pressure will approach zero. This is
generally the case at the water table interface between the
unsaturated and saturated zones.

Since the density of liquid water is essentially constant,
the pressure head term is often redefined as the matric head
(or suction head):
p
=
(< 0)
(7.3.4)
w g

F IGURE 7.6 Illustration of capillary rise as a function of capillary tube diameter. Capillary rise is proportional to suction in
water. Soil suction in soil pores act analogously with smaller
pores holding water at higher suction pressures. The capillary
rise is marked by a curved meniscus due to surface tension.

so that the piezometric head for unsaturated flow can be


written as:

h = +z

(7.3.5)


The flow (flux of water) in porous media is often
described by the volumetric flow (Q; [Q]=L3/T) per unit
cross-sectional area (A; [A]=L2) and is given by Darcys
223

(1856) Law:

q=

Q
= Kh
A

(7.3.6)

() () ()
+
+
x y z

h
=
+1
z
z

and represents a gradient when operating on a scalar field or a


divergence when operating on a vector field. The Darcy flux
(or specific discharge) q has dimensions of velocity [L T-1], K
is the hydraulic conductivity [L T-1], and the last term is the
gradient in piezometric head h(x, y, z). The head is a scalar
field so that the del operator is the gradient and represents a
vector quantity. As such q is itself a vector quantity (i.e. has
directionality). A gradient is a vector in the steepest ascent
direction (i.e. slope is defined as positive in the increasing
direction). Since flow is always in the steepest descent
direction, there is a negative sign in Darcys Law (Equation
(7.3.6)). In general q can be a three-dimensional quantity (i.e.
with three components in the x-, y-, and z-directions). Since
unsaturated flow is mainly in the z-direction only, the relevant
Darcy flux is given by:

q z = K z

h
z

h(z,t) = (z,t) + z

(7.3.8)

The key point is that the flow is proportional to both the


hydraulic conductivity and to the head gradient:

where the del operator is given by:

() =

that h is only a function of z (and time) i.e. h(z, t) and


therefore:

(7.3.7)

If this is the only component of the Darcy flux, then it implies

(7.3.9)


Based on Equations (7.3.7) and (7.3.9), flow in the
unsaturated zone depends on the hydraulic conductivity of the
soil and matric head of the soil. In particular these two
variables depend on two factors: i) Soil type (which is fixed in
time, but varies in space) and ii) soil moisture (which varies in
both space and time). To quantify these relationships,
experiments have been performed to develop constitutive
relationships between the soil type, moisture and conductivity
and matric head.

Several models exist, but a commonly used one is the
Brooks-Corey model (1966). The water retention curve (i.e.
soil matric head vs. soil moisture) is given by:

( ) = s
s

(7.3.10)

s saturated matric head

b Brooks-Corey parameter

and the hydraulic conductivity curve is given by:


224


K( ) = K s
s

K s saturated hydraulic conductivity

(7.3.11)

c = 2b + 3

where the parameters in the model are functions of soil type


(Table 7.2).

It is important to keep in mind that the saturated matric
head (sometimes referred to as the air-entry bubbling
pressure) is negative. The Brooks-Corey b parameter is a
positive number that is generally greater than 1, meaning that
the retention curve is nonlinear. For a given soil type at
saturation, the matric head is equal to the saturated matric
head and the hydraulic conductivity is equal to the saturated
hydraulic conductivity. As soil moisture becomes smaller, the
matric head becomes larger in magnitude (more negative),
while the hydraulic conductivity becomes smaller
(approaching zero). Figure 7.7 shows a schematic of the water
retention and hydraulic conductivity curves. Table 7.2 shows
that as a soil goes from sand to clay the saturated hydraulic
conductivity becomes less. Note that this is in contrast to the
trend in porosity (Table 7.1) indicating that while fine-grained
soils can actually store more water, the flow in such soils is
generally lower. Additionally, the relationships become more
nonlinear as soils become more fine-grained. It should be
clearly noted that these parameters are in general highly
uncertain and the tabulated values shown in Table 7.2 are
estimates from one dataset.

T ABLE 7.2. A VERAGE B ROOKS -C OREY P ARAMETERS


AS A FUNCTION OF SOIL TYPE (C LAPP &
H ORNBERGER , 1978). V ALUE SHOWN IN
PARENTHESES IS THE STANDARD DEVIATION .
SAT.
SOIL
TYPE

HYDRAULIC
CONDUCTIVITY
[cm/hr]

SAT.
MATRIC
HEAD
[cm]

BROOKSCOREY
PARAMETER

Sand

63.36

-12.1 (14.3)

4.05 (1.78)

Loamy sand

56.16

-9.0 (12.4)

4.38 (1.47)

Sandy Loam

12.49

-21.8 (31.0)

4.90 (1.75)

Silt Loam

2.59

-78.6 (51.2)

5.30 (1.96)

Loam

2.50

-47.8 (51.2)

5.39 (1.87)

Sandy clay
loam

2.27

-29.9 (37.8)

7.12 (2.43)

0.61

-35.6 (37.8)

7.75 (2.77)

Clay loam

0.88

-63.0 (51.0)

8.52 (3.44)

Sandy clay

0.78

-15.3 (17.3)

10.4 (1.64)

Silty clay

0.37

-49.0 (62.1)

10.4 (4.45)

Clay

0.46

-40.5 (39.7)

11.4 (3.7)

Silty clay
loam


Given these relationships, they can be used in Darcys
Law to express flux in terms of the soil moisture, i.e.:
225

(b+1)

" %
d
b
= s $$ ''
d
s # s &

(7.3.13)

where while the magnitude can vary with soil moisture, the
sign of Equation (7.3.13) is always positive (Figure 7.7).

F IGURE 7.7 Plot of matric head (blue curve/left axis) and hydraulic conductivity (green curve/right axis) as a function of
volumetric soil moisture content.

d
q z = K( )
+ 1 = K( )
K( )
z
d

(7.3.12)

where the hydraulic conductivity is given by Equation (7.3.11)


and the derivative term can be determined via the retention
curve model (Equation (7.3.10)):


The key point illustrated by Equation (7.3.12) is that the
flux at any depth in the soil is driven by two factors: i)
capillary forces driven by soil moisture gradients (first term)
and ii) gravity drainage (second term). The gravity term is
always acting downward (negative). The matric head term,
however, can act either upward or downward, depending on
the sign of the soil moisture gradient. If the soil moisture
gradient is positive (i.e. increasing soil moisture as you go up
in the soil column), then the matric head (gradient) term is
negative and reinforces the gravity term. If the soil moisture
gradient is negative (i.e. decreasing soil moisture as z increases
upward in the soil profile), then the matric head (gradient)
term is negative and acts against gravity. In both cases, the
matric head gradient term acts to drive flow from wetter soil
locations to drier soil locations. Depending on the conditions,
if the gradient is negative and large in magnitude, the first
term may be larger than the second so that flow will actually
be upward in the soil. The relative magnitude of the terms
and the sign of the matric head term will vary (significantly)
in time. During infiltration events, the soil moisture in the
surface layer will become moister than the soil below, driving
flow downward in conjunction with gravity. During a drydown period, where evaporation removes moisture from the
soil, the gradient in soil moisture may reverse and become
226

large enough to counteract gravity. The specific governing


equation for unsaturated flow dynamics will be discussed in
the next section.

Finally, given the retention curve model, one can also
then more precisely define the values for other parameters like
the field capacity and wilting point that were mentioned in
Section 2. The field capacity is defined (albeit arbitrarily) as
the level of moisture reached after drainage for a few days or
more precisely the soil moisture corresponding to a matric
head of -340 cm (Dingman, 2008), which can be used in the
Brook-Corey model to invert for soil moisture:

340 cm
fc = s

1
b

(7.3.14)

where this can be determined solely from soil type. Similarly


the permanent wilting point is defined (again somewhat
arbitrarily) as the soil moisture at which the matric head is
equal to -15000 cm (i.e. very dry):

wp

15000 cm
= s

1
b

(7.3.15)

which again only depends on soil type. These two parameters


provide the bounds within which soil moisture is typically
found (except immediately after storms when the soil moisture
will be above field capacity) as shown in Figure 7.5.

E XAMPLE 7.3.1
Soil moisture sensors in the field show that the
volumetric soil moisture profile follows the
following linear relationship: (z) = surf mz
where z is the depth below the surface (i.e. is
zero at the surface and negative below the
surface) and m is a positive constant. Determine
an expression for the Darcy flux in the soil for
this soil moisture profile. Estimate the magnitude
and direction of the flux at a depth of 0.5 m if
the soil is a sandy loam, the slope m is equal to
0.05 m-1 and the surface volumetric soil moisture
is equal to 0.4.
If the soil moisture varies with depth, then so will the
matric head and hydraulic conductivity. The matric
(pressure) head in the soil is given by:
b

b
! mz $
! (z) $
&
(z) = s ##
&& = s ## surf
&
s
" s %
"
%

and the hydraulic conductivity is given by:


c

c
! mz $
! (z) $
&
K(z) = K s ##
&& = K s ## surf
&
s
" s %
"
%

where the soil hydraulic parameters are typically either


227

E XAMPLE 7.3.1 ( CONTINUED )

E XAMPLE 7.3.1 ( CONTINUED )

measured or estimated based on soil texture. The Darcy


flux is given by:

# &
#d &
q z = K( )% + 1( = K( )%
+ 1(
$ z
'
$ d z
'

(b+1)

= 1.87 cm/hr

The matric head gradient is given by:


(b+1)

bm # surf mz &
(
=
%
(
z
s s %$
s
'

where

" (z) %
d
b
= s $$
'
d
s # s '&

(2(4.9)+3)

" 0.4 0.05(0.5) %


= (12.49 cm /hr) $
'
(0.435)
#
&

(b+1)

" mz %
b
'
= s $$ surf
'
s #
s
&

and

(4.9)(.0005 cm 1 )
=
(21.8 cm)
(0.435)
(4.9+1)

# 0.4 (0.05 m 1 )(0.5 m) &


%
(
(0.435)
$
'
= 0.295 cm cm -1

= m
z

Substituting all of the terms into the flux equation


yields:
c

" mz %
'
q z = K s $$ surf
'

#
&
s

(b+1)
(
+
"
%

mz
*bm $ surf
'
+
1
s
$
'
*
s
#
&
s
)
,

For a sandy loam soil, the soil parameters can be


estimated from Tables 7.1 and 7.2. The hydraulic
conductivity at a depth of 0.5 m is given by:

where note that this is opposite in sign to the elevation


gradient (equal to 1) and smaller in magnitude. This
tells us that the flow will be downward since the matric
head term is in the opposite direction but smaller than
the gravity term. The flux is given by:

q z = (1.87 cm/hr)"#(0.295) + 1$% = 1.3 cm/hr


which is negative and therefore downward.

" mz %
'
K(z = 0.5 m) = K s $$ surf
'

#
&
s

228


= q
t

S ECTION 4

Modeling Unsaturated
Zone Flow Dynamics

For 1D (z-direction) flow this becomes:


= (q z ) = K( )
+ 1
t
z
z
z


The modeling of unsaturated flow (or the unsaturated
flow equation) does not consist solely of Darcys Law, but
requires consideration of water going into or out of storage,
which changes the flow properties (matric head, hydraulic
conductivity). This can be done via application of the
dierential form of mass balance:

M
= ( wq dx dy dz)
t

(7.4.1)

where this is simply derived from the mass balance Equation


(1.5.2) where the right-hand-side has been transformed from
the surface integral to one over the dierential volume via the
divergence theorem. The left-hand-side is the change in mass
storage and the right-hand-side represents the convergence of
flow into the dierential volume of soil. For an unsaturated
soil we can write that the mass of water in the dierential
volume is given by:

M = w dx dy dz

(7.4.3)

(7.4.2)

Substituting this into the above equation, assuming density is


constant, and canceling like-terms yields:

(7.4.4)

which can be rewritten in terms of the dependent volumetric


soil moisture variable:


=
K(

)
+
1

t z
d

(7.4.5)

The collection of terms multiplying the soil moisture gradient


is often referred to as the diusivity (D), i.e.:

D( ) = K( )

d
d

(7.4.6)

In the case where the soil has vegetation roots within the
unsaturated zone, then a sink term S can be added:

"

= $D( ) + K( )' + S(z,t)


t z #
z
&

(7.4.7)

The root uptake sink term (S) can vary with depth and time
depending on the root density structure and available soil
moisture. In the case of no vegetation or root uptake, S = 0
and Equation 7.4.7 reduces to 7.4.5. This becomes the
229

governing equation for unsaturated flow and is commonly


referred to as the Richards (1931) Equation. Note that this is
a second order nonlinear partial dierential equation (PDE)
where the dependent variable is soil moisture and the
independent variables are depth in the soil and time. To solve
this PDE, one needs 2 boundary conditions and an initial
condition. The domain this is applied over is the unsaturated
soil column and the boundary conditions correspond to those
at the surface (infiltration/evaporation) and at the lower
boundary (recharge flux into water table or saturated
conditions). The nonlinearity comes from the dependence of
the hydraulic conductivity and matric head on soil moisture.
Due to the nonlinearity and the complicated surface boundary
conditions, Richards Equation is dicult to solve analytically.
Therefore in most applications it must be solved numerically.
This involves discretizing the soil column into many layers and
using a finite dierence or finite element solution of the
equation.

Bras (1990) discusses some special case solutions to the
Richards Equation under various simplifying assumptions. For
example, if one assumes that the gravity term is negligible and
the the diusivity is constant, Equation (7.4.5) reduces to:

2
=D 2
t
z

(7.4.8)

which corresponds to a uniform initial soil condition followed


by an abrupt shift to a dierent value that is held at the
surface (i.e. what might be seen in a high-intensity storm), the
solution to Equation (7.4.8) is given by (Eagleson, 1970):

" z
%
'
(z,t) = 0 + (i 0 )erf $
1/2
$# 2(Dt) '&

(7.4.10)

where erf is the so-called error function:

erf(y) =

e x dx

(7.4.11)

which is tabulated or available in most numerical software


packages (i.e. erf in MATLAB). With such knowledge of the
moisture profile, one could then determine the Darcy flux
anywhere in the soil column. For the above solution this
would (again assuming the gravity term is negligible to be
consistent with the original assumption) be:
1/2

#D &

q z (z,t) = D
= (i 0 ) % (
z
$ t '

# z
&
(
exp %
% 2(Dt)1/2 (
$
'

(7.4.12)

However, outside of the somewhat restrictive conditions


mentioned above, analytical solutions to Richards Equation
are not possible and it must be solved numerically.

Under the following initial and boundary conditions:

" ,
$
=# i
$% 0 ,

z 0,t = 0

(7.4.9)

z = 0,t > 0
230

E XAMPLE 7.4.1

E XAMPLE 7.4.1

For the soil moisture profile snapshot in Example


7.3.1, determine the expression for the
instantaneous volumetric soil moisture storage
change.

which is a complicated (nonlinear) function of the soil


profile and soil parameters. It can be further rearranged
to yield:

As determined in Example 7.3.1, the Darcy flux is given


by:
c

" mz %
'
q z = K s $$ surf
'

#
&
s

(b+1)
(
+
"
%

mz
bm
surf
* $
'
+ 1'
* s$
s
#
&
s
)
,

Richards equation provides the storage change, which in


the z direction is given by the following expression for
this case:

(b+2)+c

# mz &

bcm
(
= K s 2 s %% surf
(
t

s
$
'
s
2

K sb(b + 1)m 2

s2

c1

cm # surf mz &
%
(
+ Ks
(
s %$
s
'

(b+2)+c

# mz &
(
s %% surf
(

$
'
s

(b+2)+c

# mz &
bm
(
= (c + b + 1)K s 2 s %% surf
(

s
$
'
s
2

c1

cm # surf mz &
%
(
Ks
(
s %$
s
'

#
&,

)
= (q z ) = +K(z)% (z) + 1(.
t
z
z *
$ z
'#
&

2
=
K(z) % (z) + 1( + K(z) 2 (z)
z
z
$ z
'
c1 #
(b+1)
&
)
,
)
,

mz

mz
cm
bm
. % s + surf
.
=
K s ++ surf
+ 1( +
.
+
.
%
(
s
s
s
*
- $ s
*
'

) mz ,
.
K s ++ surf
.

*
s

(b+2) &
#
2
)
,

mz
%b(b + 1)m + surf
(
.
s+
2
.
%
(
s
s
*
$
'

231

d
f (t) = K( [z = 0,t ])
d z

S ECTION 5

Infiltration

A complete modeling of unsaturated zone dynamics
requires the solution of Richards equation. If the soil moisture
profile evolution is known, then everything about storage and
flux in the entire unsaturated zone is known (or can be
determined). The one aspect of unsaturated flow that is of
most interest to hydrologists is infiltration, which is the flux
at the surface resulting from a precipitation event. The prime
motivation for modeling infiltration is that, if infiltration and
precipitation are known, then the surface runo is known. The
infiltration rate can be defined as:

Infiltration rate = f (t) = q z

z =0

(7.5.1)

+ K( [z = 0,t ])

(7.5.3)

z =0,t

which implicitly requires knowledge of the soil moisture and


its gradient at the surface. The above expression provides the
infiltration rate ([L/T]). In some cases, the cumulative
infiltration ([L]) is used which is defined by:

F(t) =

f (t )dt f (t) =
0

dF
dt

(7.5.4)

To model infiltration, there are generally two options:


1. Solve Richards Equation which provides q everywhere
including at the surface, but is computationally expensive.
2. Develop models based on some simplifying assumptions.

which is simply the Darcy flux at the surface (z = 0). The


negative sign is used because the Darcy flux is defined as
positive upwards, but the infiltration is typically defined as
positive downward. Note that the infiltration rate is only a
function of time since it is the Darcy flux evaluated at the
surface which removes the spatial dependence. Substituting
Darcys Law into the above equation yields:

f (t) = K
+ 1
z
z =0
which can be written more explicitly as:

(7.5.2)

M OVIE 7.1 Animation of infiltration as a function of soil type.


232

In this Chapter we focus primarily on Option #2. To start we


will focus on the qualitative nature of the infiltration process.
Movie 7.1 illustrates the process of infiltration for a sand, silt,
and clay all experiencing the same rainfall rate. It should be
noted that the flux of water could also be from snowmelt.
Hereafter, a storm is used to refer to either case.

Figure 7.8 illustrates the soil moisture profile at dierent
times during a typical infiltration event. The limiting cases
will be discussed afterward. The initial soil moisture profile
(just before the start of infiltration) is uniform with depth
with the available storage being equal to the dierence
between the porosity and the initial volumetric soil moisture.
When infiltration begins, a so-called wetting front develops.
Early on in the storm, i.e. before the ponding time (tp), all of
the water can be infiltrated, i.e.:

f = P;

t tp

Water from infiltration quickly fills some of the pores of the


soil at the surface faster than it can be drained from below. So
the soil moisture becomes larger over time near the surface,
ultimately filling the pores so that the soil becomes saturated
at the surface. The time at which the soil surface becomes
saturated is the ponding time, referring to the fact that at
that time, water will begin to pond (build up) at the surface
and will have the condition:

F IGURE 7.8 Soil moisture profiles at various times during

steady infiltration starting from a uniform initial condition.


Note the general downward movement of the infiltration front
and the diffusion of the front over time.

At t = t p :

(z = 0,t p ) = s ;

K( [z = 0,t p ]) = K s ;

0<


=
z z

z =0,t =t p

After the ponding time, the wetting front will continue to


233

move downward, with the prime result being that the


moisture gradient at the surface gets smaller over time:
At t > t p :

(z = 0,t > t p ) = s ;

K( [z = 0,t > t p ]) = K s ;

0<


=
z z

<
t >t p

t =t p


Recall from Equation (7.5.3) that the first term in the
infiltration is proportional to the soil moisture gradient (at the
surface), while the second term is equal to the hydraulic
conductivity (at the surface). Since the gravity term at the
surface is constant after ponding, the implication is that as
the soil moisture gradient decreases over time the infiltration
rate will also decrease over time. Figure 7.9 shows a schematic
of the actual infiltration rate as a function of time, which can
be written mathematically as:

P, t t t

0
p
f (t) = *
f (t), t p t tr

(7.5.5)

where f*(t) is a (yet undefined) monotonically decreasing


function and tr is the duration of the storm (or snowmelt).
Note that in the context of solving Richards Equation, this is
equivalent to a switching boundary condition: a flux
(Neumann) boundary condition before tp and a moisture
(Dirichlet) boundary condition after tp. Note that the
infiltration is not expected to decay to zero. Rather, at the
time when the soil moisture gradient at the surface reaches
zero (i.e. first term in the Darcy flux is zero) the infiltration
rate will simply be driven by gravity and equal to the

F IGURE 7.9 Conceptual picture of actual infiltration rate

(shown in blue) involving a threshold at the ponding time.


Prior to ponding all rainfall infiltrates, while after ponding the
infiltration capacity rate limits the rate at which water can enter the soil (adapted from Mays, 2005).

saturated hydraulic conductivity (Ks). Whether this limit is


reached during a given storm depends on factors like the
initial condition, storm intensity/duration, and soil properties.
Beyond t=tr, the infiltration will cease and redistribution of
soil moisture will ensue.

Some limiting conditions to infiltration are worth
mentioning. For very intense storms or highly impervious soils
(P >> Ks), one would expect immediate ponding (tp = 0). In
this case there would not be a constant infiltration rate early
234

in the storm, but rather a monotonically decaying function


over the entire course of the storm, i.e.:

f (t) = f *(t),

t 0 = t p t tr

(7.5.6)

This is actually a special case called infiltration capacity (or


potential infiltration) that will be used in the next section to
determine the actual infiltration. These situations will
generally yield large amounts of runo. In the other limit
where a storm intensity is weak and/or the soil is extremely
pervious (P < Ks), one would expect no ponding to occur at
all. In this case there would simply be a constant amount of
infiltration rate equal to the storm intensity (P), but less than
the soil conductivity (Ks), i.e.:

f (t) = P,

t 0 t tr

(7.5.7)

M OVIE 7.3 Animation of infiltration under heavy rainfall


conditions relative to soil properties (from COMET program).

This will result in complete infiltration and no runo. Movie


7.2 illustrates this special case, whereas Movie 7.3 illustrates
the more general case where infiltration is insucient to
remove the precipitation from the surface, thus resulting in
surface runo.

M OVIE 7.2 Animation of infiltration under light rainfall


conditions relative to soil properties (from COMET program).

235

S ECTION 6

Infiltration Capacity
Models

In the discussion of infiltration in the previous section it
was mentioned that after ponding the infiltration rate should
be described by a monotonically decaying function f*(t). To
gain some insight into that function, and ultimately derive it,
we can start by examining models for infiltration
capacity (or potential infiltration): fc(t), which is the
infiltration rate that would occur under immediately ponded
conditions. Analytical models for the infiltration capacity can
be developed under the following simplifying conditions:
1. Ponded upper boundary condition, i.e.:

(z = 0,t) = s
2. The unsaturated zone is infinitely deep, i.e. the
groundwater table is not near the surface
3. The initial condition is uniform with depth, i.e.:

(z,t 0 ) = 0
Under these conditions, two commonly used models for
infiltration capacity can be developed: the Philip solution and
the Green-Ampt model. Note that the analytical model for
flow shown in Equation (7.4.12) was derived under similar

assumptions (if the surface is assumed saturated) with the


additional assumption that the gravity term was negligible.

An infinite series solution to Richards Equation for the
infiltration rate under these conditions is given by:

fc (t) = q z

z =0

= A1t 1/2 + A2 + A3t 1/2 + A4t + ...

(7.6.1)

The Philip (1960) solution keeps the first two terms, which is
expected to be valid under short times:

fc (t) =

Sp
2

(7.6.2)

t 1/2 + K s

where the soil sorptivity (Sp) is given by:

2b + 3
S p = (s 0 )K s s

b
+
3

1/2

(7.6.3)

This solution predicts that the decay occurs to the order t -1/2,
and asymptotes to Ks as expected. Interestingly, the constant
diusivity case shown in Equation (7.4.12) (when evaluated at
the surface: z = 0) shows the same time dependence. In some
versions of the model, Ks is replaced by a fitting parameter.
Also note the sorptivity is only a function of soil properties
and the (uniform) initial condition (before infiltration begins).
Based on the above model, the cumulative infiltration
capacity can be derived directly:

Fc (t) =

f (t )dt = S t

t0

1/2

+ K st

(7.6.4)

236


The second model often used for infiltration capacity is
the so-called Green-Ampt (1911) model. This model uses the
same assumptions as those above, with one additional
simplification: the wetting front is a discontinuous front at
depth z = -Lf (Figure 7.10). The Green-Ampt model applies
Darcys Law across the wetting front:

qz = K s

h hsurf
dh
K s f
dz
z f z surf

(7.6.5)

where the head at the wetting front is given by:

h f = f + z f = f Lf

(7.6.6)

the eective matric head at the wetting front is given by:

2b + 3
f

b + 3 s

(7.6.7)

and the head at the surface is given by:

hsurf = surf + z surf 0 + 0 = 0

(7.6.8)

which assumes the depth of the ponded water is negligible.


Substituting these into Equation (7.6.5) yields:

fc (t) = q z = K s

Lf + f
Lf

= K s 1 +
Lf

(7.6.9)

This is not the final solution however since the length of the
wetting front (Lf) itself depends on the amount of cumulative
infiltration, i.e.:

Fc (t) = Lf (s 0 )

F IGURE 7.10 Conceptualization used to represent the wetting front in the Green-Ampt model for infiltration capacity
(adapted from Mays, 2005).

(7.6.10)

which assumes a sharp wetting front and yields:

Lf =

Fc (t)
(s 0 )

(7.6.11)

237

Substituting into Equation (7.6.9) then yields:

f (s 0 ) dF
= c
fc (t) = K s 1 +
Fc

dt

(7.6.12)

Note that this is an ordinary dierential equation for Fc. It


can be integrated via separation of variables to yield the
following (implicit) solution:

f (s 0 )
1

t=
Fc (t) + f (s 0 )ln

Ks
Fc (t) + f (s 0 )

(7.6.13)

The key point is that the Green-Ampt model is an implicit


solution (i.e. known in terms of the dependent variable instead
of the independent variable) rather than the generally
preferred explicit solution. To determine fc(t) one can use the
following method:
1. Choose values of Fc over a range of increasing values: [0
Fc1, Fc2, ... ]
2. Solve for t using Equation (7.6.13) to get: [0, t1, t2, ...].
Keep those solutions in the range between 0 and tr.
3. Estimate fc(t) using Equation (7.6.12) corresponding to
each time.

238

S ECTION 7

Modeling Actual Infiltration with the Timecompression Approximation



Both the Philip and Green-Ampt solutions assume
ponded conditions to obtain analytical solutions for the
infiltration capacity: fc(t). In addition to the ponded surface
condition, they assume that the initial condition is uniform
with depth and the unsaturated zone is very deep. These
equations would only give us models for the actual infiltration
rate if the actual conditions corresponded to those described
above. In general that is not the case. When the precipitation
rate (P) is not intense enough to cause immediate ponding,
then initially the infiltration rate will equal the precipitation
rate. If ponding does occur during the course of the storm
(depends on soil type and precipitation intensity), then from
that point forward we would expect the infiltration rate to
decrease in a way similar to that described by the infiltration
capacity curves mentioned above.

The basic idea behind the time-compression
approximation (TCA) is to avoid having to solve Richards
Equation numerically for a general set of conditions and
instead use a time-shifted analytical infiltration capacity curve
(shifted by a compression time tc) to determine the
infiltration after the time to ponding (tp). Note that the time

shift is not simply equal to the time to ponding as will be


described below. Given this general model for actual
infiltration, for a storm starting at time t0 = 0, with intensity
(P), and duration tr, we can write the actual infiltration rate
mathematically as:

P, t 0 t t p

f (t) =
fc (t tc ), t p t tr

(7.7.1)

where f* in Equation (7.5.5) has been replaced by fc. Here the


function is kept general in terms of notation, but could be
either the Philip or Green-Ampt models where the time

F IGURE 7.11 Illustration of the model for actual infiltration

using the time-compression approximation. The dotted line represents an infiltration capacity model (i.e. under immediately
ponded conditions) which is shifted by the compression time.

239

argument (t - tc) is used instead of t. For a positive value of tc,


this just represents a time-shift to the right. This is in
acknowledgment of the fact that the decay behavior after
ponding should be comparable to infiltration capacity. In the
above model for actual infiltration there are two unknowns: tp
and tc that need to be determined. A conceptual picture of
this model is shown in Figure 7.11.

To determine the two unknowns we need to construct
two independent equations that constrain the model. These
conditions should provide physically-based constraints on the
actual infiltration model. The first, and perhaps easiest to
understand, is simply that the infiltration rate as predicted by
the shifted analytical model should equal the precipitation
rate at the ponding time:

fc (t tc )

t =t p

= fc (t p tc ) = P(t p )

F IGURE 7.12 The first condition that needs to be met by the


TCA (i.e. matching infiltration rates at ponding time).

(7.7.2)

This matching condition ensures that the infiltration rate is


continuous in time (as it should be) and is shown graphically
in Figure 7.12. The second condition may be a bit less
intuitive, but has to do with making sure that at the ponding
time, the amount of cumulative infiltration predicted by the
analytical model is consistent with how much has actually
occurred. This is just another way of saying that the starting
point in the infiltration capacity curve is that which would be
expected given the actual cumulative infiltration. This can be
written mathematically as:

tp

t p tc

P(t)dt =

fc (t)dt

(7.7.3)

which for a constant P can be written as:

Pt p =

t p tc

fc (t)dt

(7.7.4)

This condition is shown graphically in Figure 7.13. Together,


Equations (7.7.2) and (7.7.3) represent two equations for the
two unknowns tp and tc. Keep in mind that the expressions are
in terms of a general infiltration capacity model. So these two
240


If applying the Philip solution (and assuming the storm
intensity P is constant), using Equations (7.7.2) and (7.7.4), it
can be shown that:

tp =

S p2

Ks
1
+

2P(P K s )
2(P K s )

Sp
tc = t p

2(P

K
)

(7.7.5)

(7.7.6)

where the actual infiltration rate over the entire storm is then
given by:

P, t 0 t t p

f (t) = S
p
(t tc )1/2 + K s , t p t tr

(7.7.7)

where the specific values for tp and tc can be computed from


Equations (7.7.5) and (7.7.6). Note that the ponding and
compression times depend on both the storm intensity and
soil properties.

F IGURE 7.13 The second condition that needs to be met by


the TCA (i.e. matching the cumulative infiltration up to the
ponding time).

equations can be solved for the two unknowns for any


infiltration capacity model (e.g. Philip or Green-Ampt).


Similarly, if using the Green-Ampt model, it can be
shown that:

tp =

Ks
( 0 )
P(P K s ) f s

( )
1

f
s
0
tc = t p
Pt p + f (s 0 )ln

Ks
Pt
+

p
f
s
0

(7.7.8)

(7.7.9)
241

where the actual infiltration rate over the entire storm is then
given by:

) P,
+
+
# ( )&
f (t) = *
f
s
0
(,
+ K s %1 +
%
Fc (t tc ) ('
+,
$

t0 t t p
t p t tr

(7.7.10)

where Fc(t - tc) vs. t needs to be determined via the implicit


Green-Ampt model equation as described above.

Given these models for actual infiltration, it is worth
reiterating the special cases mentioned above:
1. In the case of a high intensity precipitation event, i.e.

P
it can be seen that for either model, the time to ponding and
compression time both go to zero. This simply reduces the
actual infiltration rate to the infiltration capacity rate (as
would be expected since it is consistent with the original
assumption of instantaneous ponding).
2. In the case where the precipitation rate is less than the
saturated hydraulic conductivity of the soil, then no ponding
will ever occur (since both infiltration capacity curves
asymptote to the hydraulic conductivity). In this case the
actual infiltration rate is simply given by: f(t) = P and the
above computation of ponding and compression times can be
skipped altogether. So it is important to first check whether P
> Ks when computing actual infiltration. Otherwise, using the

above equations will yield nonsensical values for the ponding


and compression times. Once known, the actual infiltration
function can be used to estimate cumulative infiltration and
the so-called infiltration excess runo during the storm. The
discussion of runo is saved for Chapter 10.
E XAMPLE 7.7.1
A rainstorm of duration tr = 8 hr and constant
rainfall intensity P= 30 mm hr-1 occurs over an
area which has a soil type characterized by the
following parameters:

s = 0.35; b = 1; s = 50 cm; K s = 11 mm/h


The soil has an initial soil relative saturation of
50%. Compute the actual cumulative infiltration
from this storm.
To compute the actual infiltration we should use the
time-compression approximation which requires the
choice of an infiltration capacity model. Here we will use
the Philip solution for infiltration capacity. As a first
step it needs to be determined whether ponding is
possible. Since P > Ks in this case, there is at least the
potential for ponding so that the ponding time and
compression time equations can be applied. If this were
not the case there would be no runo and all water
would infiltrate.

242

E XAMPLE 7.7.1 ( CONTINUED )

E XAMPLE 7.7.1 ( CONTINUED )

Given the choice of the Philip solution, first compute the


sorptivity (Equation (7.6.3)):

Note this is less than 240 mm, which would be the


cumulative infiltration if all water infiltrated. Ponding
generates runo and reduces infiltration.

1/2

(
" 2(1) + 3 %+
S p = *(0.35 0.5(0.35))(11 mm/h) 500 mm $
'(1)
+
3
#
&,
)
34.7 mm h 1 2

We could also choose to apply the Green-Ampt model,


where the relevant parameters (using Equations (7.7.8)
and (7.7.9)) are:

Next, Equations (7.7.5) and (7.7.6) can be used to


compute the ponding and compression times:

! 2(1) + 3 $
f = #
&(500 mm) = 625 mm
1
+
3
"
%

"
%
(34.7 mm h 1/2 )2
11 mm/h
tp =
$1 +
'
2(30 mm/h)(30 11) mm/h # 2(30 11) mm/h &

(s 0 )=0.175;

= 1.36 h
2

An expression for the cumulative infiltration can be


obtained by integrating Equation (7.7.7):

tr
0

f (t)dt = Pt p + [S p (t tc )1/2 + K st ]tr

= Pt p + S p [(tr tc )1/2 (t p tc )1/2 ] + K s (tr t p )


mm
mm
(1.36 h) + (34.7 1/2 )[(7.47 h)1/2
h
h
1/2
(0.83 h) ] + K s (6.64 h) = 177 mm

tc = 2.1 hr

1
"(30 mm/h)(2.1 h) + (109.375 mm)
11 mm/h #

ln {109.375 mm / ((30 mm/h)(2.1 h) + 109.375 mm)}


= 0.9 h

= Pt p + S p (tr tc )1/2 + K str S p (t p tc )1/2 K st p


= 30

11 mm/h
(109.375 mm)
(30 mm/h)(30 11) mm/h
= 2.1 hr

tp =

" 34.7 mm h 1/2 %


tc = 1.36 h $
' = 0.53 h
2(30

11)
mm/h
#
&

F(t = tr ) =

f (s 0 ) = 109.375 mm

The cumulative infiltration (Fc) as a function of time (t)


can be obtained via Equation (7.6.13). By construct, at
time t = tp, the cumulative infiltration should equal Ptp.
The cumulative infiltration at the end of the storm will
be the cumulative infiltration corresponding to the time:
t = tr - tc. Doing so, yields a cumulative infiltration:
243

E XAMPLE 7.7.1 ( CONTINUED )

Fc (t = tr tc = 7.1 h) = 187 mm
which is about 6% dierent than predicted by the Philip
solution.

244

S ECTION 8

MOD-WET Codes

Relevant functions based on concepts introduced in
this chapter include:
Brook-Corey retention curve model:

brooks_corey_PSI_and_K.m
Field capacity (using Brooks-Corey model):

field_capacity.m
Green-Ampt infiltration capacity model:

green_ampt.m
Philip infiltration capacity model:

philip.m
Soil sorptivity:

sorptivity.m
Time-compression approximation for actual infiltration:

TCA_infiltration.m
Permanent wilting point (using Brooks-Corey model):

wilting_point.m

245

S ECTION 9

Conceptual Questions
1. Define the meaning of porosity. How is it used to transform
between volumetric water content and relative saturation in
a soil?
2. What are the three end-member soil types in the SCS soil
triangle?
3. Which has a higher porosity, sand or clay? Which has a
higher field capacity? Which has a higher wilting point?
4. What are the two terms in the piezometric head? Describe
what each term physically represents.
5. What is the sign of matric head. Explain why.
6. How does the hydraulic conductivity and matric head
change as soil moisture decreases.

9. In unsaturated zone problems, name the variable that, if


known, allows for the determination of everything else
about the system. Hint: It is the dependent variable in
Richards equation.
10. Mathematically, how are the infiltration rate and
cumulative infiltration related?
11. Concisely explain the dierence between an infiltration
capacity rate and an actual infiltration rate. Under what
conditions are they the same?
12. Name the two commonly used models for infiltration
capacity.
13. Concisely describe the conceptual meaning of the ponding
time and compression time that are solved for in the timecompression approximation for infiltration.
14. If for a particular storm the precipitation intensity is less
than the saturated hydraulic conductivity of the soil, do
you need to apply the time compression approximation?
Justify your answer.

7. The Darcy flux dictates that the flow (per unit area) is
proportional to the piezometric head gradient. What soil
property is the proportionality constant in the Darcy
equation? What two factors generally cause that parameter
to vary in unsaturated soil (i.e. what does it depend on)?
8. When fully expanded, what are the two terms in the Darcy
flux in an unsaturated soil? Explain the physical meaning of
each term.
246

S ECTION 10

Sample Problems
Problem 7.1. After collecting multiple soil samples in the
field, you return to the laboratory and conduct soils tests to
classify the soil types present in your study basin. Sieve
analysis indicates that soil #1 has a composition of 34% sand
and 46% silt, while soil #2 has a composition of 81% sand
and 15% silt.
a) Classify both soils (i.e. determine the soil texture) using
the soil texture triangle. Based on the soil parameters listed in
Tables 7.1 and 7.2, plot the matric head and conductivity vs.
volumetric soil moisture curves for these two soils. Discuss the
relationship between soil moisture and matric head and
hydraulic conductivity. How do these relationships vary with
soil type?
Use the soil parameters from part a) for soil #1 to answer the
following:
b) Soil moisture sensors installed at location #1 (i.e. the
sampling site of soil #1) indicate that the volumetric soil
moisture profile at a given snapshot in time is:

(z) = 0 + az;

2 m z 0

where z is the depth below the surface and is negative, a =


0.05 m-1 , and the soil moisture at the surface (i.e. z = 0 m) is
0.28. Compute and plot the soil moisture profile.
c) For this soil and soil moisture profile, write the
corresponding analytical expression for the profile of matric
(suction) head. Compute and plot the matric head profile. Be
careful with the signs.
d) For this soil and soil moisture profile, write the
corresponding analytical expression for the profile of
piezometric head (matric head plus elevation head). Compute
and plot the head profile. What can you say qualitatively
about the flow direction based on the head profile? Justify
your answer.
e) Write the corresponding equation for the Darcy flux
(vertical flow) as a function of depth based on the above
profile. What is the magnitude and direction of the flux (in
units of cm/day) at a depth of 0.65 meters? What is the
relative saturation at this depth?
f ) The field capacity and permanent wilting point provide
bounds on typically observed soil moisture. Compute these
parameters and comment on how the profile compares to
these values.
g) Richards equation (mass balance) indicates how soil
moisture storage will increase or decrease (i.e. fill up or drain
from the pores) over the profile, i.e. for one-dimension:

247


q(z)
=
t
z

characterized by the following parameters: soil mineral density


2.65 g cm-3, bulk density 1.72 g cm-3, and Brooks-Corey
parameters:

In other words, the convergence of flux based on the profile at


a given time dictates exactly how storage is changing over the
profile. Based on your expression for the Darcy flux derived
above, derive an expression for the storage change and plot
the profile of storage change. Discuss where water is going
into/out of storage in the profile.

s = 50 cm; K s = 25 mm hr -1; b = 1

h) A special case in the unsaturated zone is hydrostatic


conditions (i.e. the case where the Darcy flux is equal to zero).
For this special case, derive an expression for the value of the
matric head gradient.
Problem 7.2.
a) The Philip and Green-Ampt equations provide models for
infiltration capacity (or potential infiltration). What
assumptions are used in the development of the Philip and
Green-Ampt infiltration capacity models? Clearly explain the
dierence between actual infiltration rate and infiltration
capacity (potential infiltration). Under what specific
conditions are the two the same?
b) What factors aect the infiltration (capacity) rate?
Explain.
c) Compute and plot (on the same figure) the infiltration
capacity (in mm/hour) curves vs. time for both models for a
two-hour storm. Assume the initial relative soil saturation
(not volumetric soil moisture) is equal to 0.35. The soil is

d) What is the predicted infiltration capacity for the two


models at the end of the storm? How similar/dierent are the
two model predictions? Explain.
Problem 7.3. A rainstorm of duration tr = 18 hr and rainfall
intensity P = 30 mm hr-1 occurs over an area which is
classified as a Sandy Clay Loam with an initial soil saturation
of 50%. These conditions will form the baseline case for the
sensitivity tests described below. Note: Any computations
related to the actual infiltration rate implies you should use
the time compression approximation (TCA). Here you can use
the TCA with the Philip solution.
a) What is the physical meaning of the ponding time in
infiltration? Under what storm intensity condition will
ponding never occur? Compute the ponding time for the
baseline case.
b) Compute and plot the ponding time for a range of storm
intensities consisting of the baseline case, 50% of the baseline
case and 200% of the baseline case. Explain the trend seen in
ponding time as a function of precipitation rate.
c) Compute and plot the ponding time for a range of initial
conditions consisting of the baseline case, 5%, 25%, 75%, and
248

95% relative saturation. Explain the trend seen in ponding


time as a function of initial condition.

c) What is the cumulative infiltration as a result of the


storm?

d) Compute the expected actual cumulative infiltration and


infiltration excess runo (in mm) for the baseline storm case.
e) What fraction of the total rainfall is converted to runo for
the baseline case?
f ) Qualitatively, how would your answers to parts d) and e)
change if the precipitation intensity was higher than the
baseline case?
g) Qualitatively, how would your answers to parts d) and e)
change if the initial condition in soil moisture was lower than
the baseline case?
h) Qualitatively, how would your answers to parts d) and e)
change if the soil type were a sand?
Problem 7.4. After a prolonged drought, a sandy soil is
completely desiccated (s = 0). Based on soil properties, the
soil is determined to have a sorptivity of Sp = 21.8 cm hr-1/2.
A storm of duration tr = 5 hours and of constant intensity P
= 6 cm hr-1 occurs. Using the Philip solution for infiltration
capacity and the time-compression approximation:
a) Write the analytical equation for the time-to-ponding for
these dry conditions.
b) Using your expression from part a), what is the time-toponding for this storm?
249

Chapter 8

Evaporation

S ECTION 1

Learning Objectives
By the time you finish this chapter you should be able to:

10. Define, describe, and compute the aerodynamic


resistance to turbulent heat transfer
11. Define the Bowen ratio
12. Compute the Bowen ratio from dual-level measurements
of air temperature and humidity

1. List the three requirements needed for evaporation to


occur

13. Estimate evaporation from the Energy Balance Bowen


Ratio (EBBR) method

2. Convert back and forth between evaporation mass flux,


latent heat flux, and depth flux (i.e. flux density)

14. Compute the latent and sensible heat flux using the
mass-transfer approach

3. Write down the units of evaporation mass flux and latent


heat flux

15. Apply the Penman or Penman-Monteith models to


estimate potential evaporation and/or evapotranspiration
respectively

4. Define potential evaporation


5. Write down the surface energy balance
6. Define the individual terms of the surface energy balance
and what they represent

16. Explain how vegetation tend to control water loss via


their stomata and what environmental factors regulate
transpiration
17. Explain and compute the vegetation canopy resistance

7. Estimate evaporation as a residual in the energy balance


equation
8. Describe the functional form of the average horizontal
wind speed profile in the atmospheric surface layer
9. Determine the aerodynamic roughness length and zeroplane displacement height in terms of the height of surface
roughness elements
251

S ECTION 2

Basics of Evapotranspiration

Evapotranspiration (ET) refers to all the processes by
which water in liquid phase at or near the Earths surface
becomes atmospheric water vapor. In particular it generally
refers to the flux of water from the surface into the
atmosphere. Evaporation usually describes the direct
vaporization of water from either open water surfaces or bare
soil. Transpiration refers to water loss from within the leaves
of plants. Together the two fluxes are referred to as
evapotranspiration. Sublimation is a special case whereby
vaporization occurs directly from ice. A more comprehensive
treatment of evaporation is given in Brutsaert (1982).

3. Evapotranspiration rates during inter-storm periods control


the antecedent soil moisture before the subsequent storm,
which impacts runo and flooding.
4. Evapotranspiration is a key flux that couples the surface
water and energy budgets (i.e. recall SEB in Chapter 6:
Rn - G = LE + H).

The direct measurement of evapotranspiration however is
extremely dicult, especially compared to other variables like
precipitation and streamflow. The closest readily-available
measurement networks are so-called evaporation pans (Figure


Evapotranspiration is important for several reasons
including:
1. The long-term water balance, which depends on
evapotranspiration (as well as precipitation and runo)
determines the available water for human use.
2. Much of the food supply is grown via irrigated agriculture
and ecient irrigation practices requires knowledge of
evapotranspiration.

F IGURE 8.1 Photo of a standard evaporation pan which provides an estimate of potential evaporation.

252

8.1). Such measurements are easy to make, but are very


sparsely distributed and do not measure actual
evapotranspiration, but rather the potential evaporation
(discussed in more detail below). As a result, estimates of ET
are most often made via the use of models that require more
readily-available measurements.

Long-term estimates of ET can be determined via
developments already discussed earlier, namely mass balance
and/or energy balance. For example for a watershed control
volume, the long-term average mass balance can be written as:

dS
= P E Q 0
dt

(8.2.1)

where ET can be estimated as a residual provided the other


terms are known:

E = P Q

(8.2.2)

Similarly, the long-term surface energy balance can be written


as:

G = Rn LE H 0

(8.2.3)

where the long-term ground heat flux can be reasonably


assumed small for similar reasons to the long-term water
storage. A non-zero value would imply a continual warming/
cooling of the surface. Again, solving for the ET (specifically
in terms of latent heat flux) as a residual yields:

LE = Rn H

So Equations (8.2.2) or (8.2.4) provide a mechanism for


estimating the long-term values of ET for a given control
volume. In many applications we need to estimate it over
much shorter time-scales (e.g. hourly or daily). The remainder
of the chapter focuses on models useful for that purpose.

In terms of the basic physics of ET, one can think of
three necessary requirements:
1. A water source (generally an open water surface, soil
moisture, or snowpack),
2. An energy input for vaporization (generally in the form of
net radiation Rn),
3. A transport mechanism to remove vapor-rich air from the
near-surface (otherwise air will saturate; need gradients in
moisture to maintain a net flux).
Note that when water is not a limiting factor in the surface
ET (e.g. open water surfaces or very moist soil), then only the
second and third requirements control the ET flux. Under
such conditions, the evapotranspiration flux is referred to as
the potential evapotranspiration (Ep), which may be referred
to as potential evaporation or potential evapotranspiration in
the case of a bare soil or vegetated surface respectively. By
definition the actual evapotranspiration must be less than or
equal to the potential evapotranspiration, i.e.:

E Ep

(8.2.4)
253

S ECTION 3

Mass-transfer Model for


Evaporation

Evaporation is the net flux of vapor from the surface to
the overlying air. In truth, e.g. over a water surface, individual
molecules are continually moving (somewhat randomly) from
the air to the liquid water phase and vice versa (Figure 8.2).
If the rate of exchange is balanced, then there is no net flux
and hence the evaporation is zero. If there are more water
molecules leaving the liquid phase than returning to it, then
evaporation is occurring. Conversely, if more water molecules
are entering the liquid surface than leaving it, the net flux is a
condensation flux.


Evaporation can be thought of as a diusive process,
whereby a flux is driven by concentration gradients. In this
case, the concentration refers to the concentration of water
vapor molecules, which can be expressed in terms of the vapor
pressure. Over a liquid water surface, the air in immediate
contact with the liquid has a saturated vapor pressure. So
when the air above the surface layer is sub-saturated, there is
a gradient in vapor concentration, which would be expected to
correspond to a vapor flux away from the surface. It is
important to remember that the saturated vapor pressure is a
function solely of temperature, hence if there is a temperature
gradient between the surface (Ts) and the overlying air (Ta),
this may lead to a vapor pressure gradient even if both are
saturated. In Figure 8.3 the air in contact with the liquid has
a vapor pressure equal to es(Ts), while that above is at
ea<es(Ts). Under these conditions, there would be expected to
be an evaporative flux away from the surface. In general,
neither the surface or overlying air may be saturated, however
the gradients will still drive the net flux.

Speaking more generally of a diusive flux, the rate of
transfer (flux) of constituent X in direction z can be written
(using Ficks 1st Law):

Fz (X ) = DX

F IGURE 8.2 Conceptual picture of water vapor molecules in a


bulk water surface and the overlying air (adapted from Dingman,
2008).

dC(X )
dz

(8.3.1)

where C(X) is the concentration of constituent X and DX is


the diusivity of constituent X in the fluid (in this case air).
This simply states that the flux is expected to be driven from
254

areas of high concentration to low concentration (down the


concentration gradient) where the diusivity is a constant of
proportionality (to be determined). Note this is analogous in
form to Darcys Law and can be used to describe dierent
fluxes in many systems.

If applied to water vapor in the atmosphere (i.e.
evaporation) we can write:

E = Fz (vapor) = Dv

d v
dq
de
= Dv
= Dv
dz
dz
p dz

(8.3.2)

where in this form: [E ]= kg m-2 s-1. To get the latent heat


flux, the above could simply be multiplied by the latent heat
of vaporization (Lv). The sensible heat flux in the SEB can
also be conceptualized as a diusive flux (of heat) and is
therefore presented here in parallel to evaporation:

H = Fz (heat) = DH

d( c pTa )
dz

= c pDH

dTa
dz

(8.3.3)

where in this form: [H ]= W m-2. Note, as can be seen from a


unit analysis, the density in Equations (8.3.2) and (8.3.3) is
air density. The specific heat of air is: cp = 1004 J kg-1 K-1.

The next question is what are the diusivities related to?
When dealing with molecular diusion (e.g. envision dye
diusing in a water-filled beaker), the diusivities are simply
related to the constituent and the fluid. However in landatmosphere interaction, molecular diusion is not the driving
process. Instead, it is turbulent fluid motions in the

atmosphere that transport constituents from near the surface


to higher in the atmosphere (and vice versa). These turbulent
eddies, which are caused by frictional wind shear and thermal
stratification in the surface layer (i.e. lower few hundred
meters) of the atmosphere, are much more ecient than
molecular diusion in transporting vapor, heat, and
momentum between the land and atmosphere. Figure 8.3
shows a schematic of the mean horizontal wind profile as a
function of height above the surface as well as turbulent
eddies (of varying sizes) superimposed on the mean flow. At
the rough surface the mean velocity must be zero (no-slip
boundary condition), and much higher in the troposphere it is
driven by the large-scale circulation of the atmosphere. The
surface layer is the lower portion of the atmosphere where
there is a strong gradient in the horizontal velocity profile.

The profile of the mean horizontal wind for a neutral
surface layer (i.e. with a temperature gradient equal to zero)
has a well-known logarithmic form:

V (z) =

u* " z d %
ln $
';
# z0 &

z > z0 + d

(8.3.4)

von Karman constant = 0.4


u* friction velocity
The parameters z0 and d are called the momentum roughness
height and zero-plane displacement height respectively and are
representative of the impact of the roughness of the surface on
the mean flow. They are often parameterized in terms of a
characteristic roughness height of the surface (h) as:
255

z 0 0.1h

d 0.7h

(8.3.5)

The logarithmic profile implies a positive wind shear, i.e.:

V
>0
z

(8.3.6)

This wind shear is one of the mechanisms that generates an


overturning of the fluid leading to turbulence. The other
mechanism is if the surface layer also exhibits a temperature
gradient (i.e. non-neutral). The surface layer generally has a
positive temperature gradient (unstable condition) during the
daytime and a negative temperature (stable condition)
gradient at night. The unstable condition generally enhances
turbulence while the stable condition generally suppresses
turbulence. The positive wind shear also implies a negative
(i.e. toward the surface) momentum flux.

F IGURE 8.3 Plot of mean horizontal wind speed profile over-

lying a rough surface with turbulent eddies superimposed on


the mean flow. Turbulence is generated mechanically by wind
shear (forced convection) and can be enhanced or suppressed
by thermal gradients in the surface layer (free convection). The
turbulence is the primary transport mechanism for energy, water vapor and momentum fluxes between the surface and atmosphere (adapted from Dingman, 2008).


So the diusivities in the flux equations for E and H are
related to turbulence. Here we will focus on the case over bare
soil/open water and will extend to vegetated surfaces in the
next section. If it is assumed that DH = DV, it can be shown
that the impact of turbulence can be embedded in E and H
models (using a finite dierence approximation):

2V22

(q 2 q1 )
2
( " z d %+ (V2 V1 )
** ln $ 2
'-z
) # 0 &,

(8.3.7)

256

(8.3.8)

2V22

(T2 T1 )
H c p
2
( " z d %+ (V2 V1 )
** ln $ 2
'-z
) # 0 &,

m momentum stability correction factor

h sensible/latent heat stability correction factor

where the subscripts on T, q, and V represent values of those


variables at two measurement levels: z1 and z2. The above
models are for a neutral surface layer. Note that, strictly
speaking, we would expect the sensible heat flux to approach
zero in neutral conditions since the temperature gradient is
zero. Modifications to account for thermal stability eects can
be included (shown below). In applying these models we often
choose one level to be at the surface (i.e. z1 = z0 + d) where
by definition V1 = 0 and re-define the other variables as:

T1 = Tsurf ,

q1 = q surf , V2 = V ,

T2 = Ta ,

raN

The stability correction factors scale the resistance in a way


that recognizes enhancement or suppression of resistance to
turbulent transport in unstable or stable conditions
respectively. A commonly used empirical form for these
functions are (Businger et al., 1971):

h = m2 = (1 15RiB )1/2 ;
h = m = (1 5RiB )1;

RiB 0 (unstable)
0 RiB < 0.2 (stable)

(8.3.11)

where RiB is the bulk Richardson number, which is a nondimensional ratio of the thermal stability to wind shear:

q 2 = qa

From this we can then define the aerodynamic resistance to


turbulent transport (under neutral conditions):

( " z d %+
** ln $
'-) # z 0 &,
=
2V

(8.3.10)

ra = raN mh

RiB =

(g /Tsurf )T z
(V z)

(8.3.12)

(8.3.9)

where the N subscript is meant to denote neutral conditions.


This resistance has dimensions of T/L (i.e. s m-1). The height
z in Equation (8.3.9) is the reference-level height at which the
meteorological measurements are taken (e.g. often 2 meters
above the surface). For the more general (non-neutral surface
layer) case, the resistance can be generalized using the
approximation:

Note that for neutral conditions (i.e. where the temperature


gradient in the above equation is zero):

RiB = 0;

m = h = 1

(8.3.13)

so that the aerodynamic resistance in Equation (8.3.10)


reduces to the neutral value shown in Equation (8.3.9). The
sign of the temperature gradient in the surface layer
(numerator in Equation (8.3.12)) controls the sign of RiB
which corresponds to unstable or stable conditions. For
unstable conditions (negative temperature gradient), the
257

stability correction factors are less than 1.0 (which reduces the
resistance; enhanced turbulence), while for stable conditions
they are greater than 1.0 (which increases the resistance;
suppressed turbulence).

Using this notation the evaporation flux model can be
written as:

E=

(q surf qa )
ra

(8.3.14)

or equivalently in terms of latent heat flux:

LE = Lv

(q surf qa )
ra

(8.3.15)

and the sensible heat flux model can be written as:

H = c p

(Tsurf Ta )
ra

(8.3.16)

The resistance terminology is used because one can


conceptualize these fluxes using a circuit analogy, where the
current in this context is the flux, the potential dierence
is the dierence in either humidity or temperature and ra is
the resistance. This will be discussed more in the next section.

These so-called mass-transfer (or diusion analogy)
models show that the flux is proportional to the dierences in
the relevant variables between the surface and reference-level,
and the resistance describes how eciently the turbulence is

at transporting the vapor/heat away from the surface. From


the resistance model, it can be seen that as the reference-level
velocity increases, the resistance decreases (which causes the
flux to increase). This is because a larger reference-level
velocity implies more surface layer turbulence, which will
increase the turbulent transport. Similarly, if the roughness of
the surface increases, turbulence will also increase, reducing
the resistance and increasing the fluxes. It should be noted
that these flux models so far have been developed for bare soil
or open water surfaces only. The resistance analogy becomes
useful as other factors (most notably vegetation) become
relevant, the eects of which can also be embedded in a
resistance term. Then the law of resistances in series or
parallel used in circuits can be used to easily augment these
models. For example, in some models an additional resistance
is used to model the barrier in vapor leaving the soil in which
case the resistance in the denominator of Equations (8.3.15)
and (8.3.16) can be replaced by: r = rsoil + ra.

To estimate the fluxes using the above models requires
specification of time-varying variables including:
meteorological measurements (V, Ta, qa, i.e. from a weather
station), and surface states Tsurf and qsurf, as well as relatively
static surface characteristic roughness parameters (h, z0, d).
The surface states are often predicted as part of a land-surface
model (i.e. that models the surface energy/water balance).
The fluxes are driven by gradients in humidity and
temperature between the surface and atmosphere (larger
gradient yields larger fluxes) and regulated by turbulent
transport (smaller resistance yields larger fluxes).
258


Over well-watered surfaces (i.e. open water or moist soil),
the air in contact with the surface is expected to be saturated.
In this special case one can generally assume: qsurf = qs(Tsurf)
which is equivalent to E = Ep (i.e. potential evaporation).
Based on this, one model used for potential evaporation is:

Ep =

(q s (Tsurf ) qa )
ra

(8.3.17)

In the more general soil case, soil moisture limits qsurf (i.e. qsurf
< qs(Tsurf)), and E < Ep (i.e. actual evaporation is less than
the potential evaporation). In such cases, the soil moisture
impact is often modeled as:

q surf = ( )q s (Tsurf )

(8.3.18)

soil moisture reduction factor, 0 1


or via:

E = ( )E p

(8.3.19)

where in this example the soil moisture reduction factor


linearly increases between the wilting point and field capacity
and is otherwise 0 or 1 below the wilting point or above the
field capacity respectively.

Connecting these models back to the necessary
requirements for evaporation stated in Section 2:
i) Water availability: This comes into the above equations
implicitly through qsurf or explicitly via the soil moisture.
ii) Energy input: The net radiation implicitly drives the
evolution of surface temperature, which itself appears
implicitly or explicitly in the evaporation equations above.
iii) Transport of vapor away from surface: The aerodynamic
resistance is the term that expresses the regulation of
transport of vapor (or energy) away from the surface via
turbulent motions.
E XAMPLE 8.3.1

where in either case the soil moisture reduction factor is an


increasing function of soil moisture that varies between 0 and
1, e.g.:

Meteorological and surface measurements


collected at noon from a bare soil field site are:

#
1,
fc
%
%%
wp
, wp fc
=$

% fc
wp
%
0,
wp
%&

Air temperature =27.2C


Air relative humidity =69%
Surface air pressure = 973 mb
Windspeed = 5.8 m/s
Soil surface temperature = 26.9C

(8.3.20)

259

E XAMPLE 8.3.1 ( CONTINUED )

E XAMPLE 8.3.1 ( CONTINUED )

Assume the soil is saturated (i.e. after a heavy


rain) and has roughness elements with an average
height of 2 cm. Compute the instantaneous latent
heat flux. For simplicity assume neutral
conditions in your calculation. Comment on the
qualitative dierence you would expect if nonneutral conditions were included. What is the
equivalent evaporation rate in mm/day?

The specific humidity of the air and aerodynamic


resistance (for neutral conditions) are given by:

Given the measurement of surface soil temperature and


knowledge that the surface is saturated, we can safely
assume that:

q surf = q s (Tsurf ) =

es (Tsurf )
p

Using the integrated Clausius-Clapeyron equation, the


saturated soil surface vapor pressure can be computed
which yields es(Tsurf)=36.2 mb, which then gives:

q surf = 0.622

36.2 mb
= 23.1 10 3 kg/kg
973 mb

The evaporative mass flux is given by:

E=

(q surf qa )
ra

es (Ta )
p
36.9 mb
= (0.69)(0.622)
= 16.3 g/kg
973 mb

qa = RH q s (Ta ) = RH

raN

( " 2 m 0.7(0.02 m)%+


* ln $
') # 0.1(0.02 m) &,
=
= 51 s/m
(0.4)2(5.8 m/s)

The air density can be computed as:

p
RdT[1 + 0.608qa ]

97300 Pa
(287J/kg/K)(300.35K)[1 + 0.608(16.3 10 3 )]

= 1.12 kg/m 3
so that the evaporation rate is:

(23.1 16.3) 10 3 kg/kg


E = (1.12 kg/m )
(51 s/m)
3

= 1.49 10 4 kg/m 2 /s
The latent heat flux is determined by multiplying the
above mass flux by the latent heat of vaporization:

260

E XAMPLE 8.3.1 ( CONTINUED )

LE = Lv E = (2.5 10 6 J/kg)(1.55 10 4 kg/m 2 /s)


= 373 Wm 2
To get the equivalent flux density in mm/day simply
requires a unit conversion, i.e.:

E[mm/day] = (1.49 10

1 m3
kg/m /s)

1000 kg
2

1000 mm 86400 s

= 12.9 mm/day
1m
1 day

Note that these answers assumed neutral conditions. In


fact the conditions are near-neutral since the referencelevel temperature is almost the same as the surface
temperature. In this case, the air temperature is slightly
warmer than the surface. This corresponds to stable
conditions which generally suppresses turbulent flow.
Mathematically, this would correspond to a higher
aerodynamic resistance and hence a lower latent heat
(evaporation) flux.

261

S ECTION 4

Transpiration

The models presented in the previous section are relevant
for open water surfaces or bare soil surfaces. However in many
applications, vegetation is present and is expected to play a
key role in ET. Transpiration by plants occurs via the
vascular system of the plant structure (Figure 8.4). It is
helpful to first start with the motivation underlying the
plants role in water loss from the surface. In this context,
photosynthesis is the key driver. Plants use photosynthesis to
build plant structures (leaves, stems, roots, etc.) in order to
grow. The photosynthetic reaction can be written as:

H 2O+CO2 CH 2O+O2

(8.4.1)

where the key driver implicit in the forward progression of


this reaction are photons from the sun (photosynthetically
active radiation [PAR], which is essentially energy in the
visible part of the solar spectrum). The carbohydrate on the
right-hand-side (CH2O) is the compound that is used to
build/grow plant structure. So the key ingredients to make
this happen are water (H2O), carbon dioxide (CO2) and solar
energy (PAR). A by-product of the process is oxygen (O2).

The water source is soil moisture taken from the
rootzone and the carbon dioxide source is from the
atmosphere. The site of photosynthesis is inside the leaf,
where carbon dioxide can diuse into the leaf and solar

F IGURE 8.4 Schematic of key structures of a plant as relevant to transpiration flux (by Laurel Jules from

http://en.wikipedia.org/wiki/File:Transpiration_Overview.svg).

radiation is absorbed. Water is brought to the leaves via


absorption of soil water by plant roots and transport via the
roots, stem, branches, etc. (Figure 8.4). Diusion of carbon
dioxide into the stomatal cavities inside the leaves occurs via
small openings called stoma (Figures 8.5 and 8.6). These
stomatal openings are ringed with so-called guard cells,
which can open and shut to regulate the flow of carbon
262

dioxide into the leaves or the flow of water vapor out of the
leaves. The reason for this regulation has to do with the fact
that photosynthesis requires all of the ingredients listed above.
When the ingredients are unavailable (i.e. nighttime when
PAR = 0, or when water is limited due to soil moisture
shortage, etc.), there is little benefit to having the stoma open
and needlessly lose the water (which will diuse out into the
air). So in this context the transpiration is essentially a water
loss term that happens at the leaves of the plant.

The water is mobilized at the root structures via suction:

F IGURE 8.6 Microscope view of a leaf stomatal opening (from


sols.unlv.edu/Sculte/Anatomy/Leaves/PopulusStomata.jpg).

F IGURE 8.5 Illustration of leaf-level processes involved in


photosynthesis (from earthobservatory.nasa.gov/Features/LAI/
LAI2.php).

the water loss at the stoma provides a suction that propagates


through the system, ultimately applying a negative pressure
to the water in the soil. This negative pressure can act against
gravity as described in Chapter 7. Recall that the drier the
soil, the higher the suction pressure in the soil. Hence, when
soil moisture is low, the plant must exert more pressure to
mobilize the water. When overly high suction pressures
develop, the plant structure may fail (causing wilting) which
corresponds directly to the wilting point soil moisture defined
in Equation (7.3.15).

263


The evaporative loss occurs because the air inside the
stomatal cavity is essentially saturated (i.e. q = qs), while that
outside the leaf is generally sub-saturated (i.e. q < qs). This
gradient in vapor will drive a diusive flux as described in the
previous section. The guard cells attempt to regulate the
carbon dioxide influx in a way that minimizes environmental
stresses on the plant (chiefly water loss). This ability to
regulate transpiration will add an additional resistance to
the flux equations developed in the previous section. This can
be modeled in terms of an extra stomatal or vegetation
canopy resistance (Figure 8.7). The details of this
physiological behavior are complicated, but the way it is
handled in the modeling of ET is generally simplified. The
resistance for the entire vegetation canopy (i.e. all of the
leaves combined) is often scaled from the single-leaf scale
using the so-called leaf area index (LAI), which is defined as
the leaf area per unit area of ground surface. So dense
vegetation with a large leaf area to ground surface area may
have values of LAI of approximately 6.0 or greater. This
parameter has both a strong spatial and seasonal variation
(Movie 8.1).

The canopy resistance (rc) is then modeled as:


r
rc = s
LAI

F IGURE 8.7 Conceptualization of controls on evapotranspiration process using a resistance analogy approach (from

(8.4.2)

where rs is the stomatal resistance. In the same way that ra is


the resistance to evaporation due to turbulent transport, rc is
the resistance to transpiration due to stomatal control. Since

media.wiley.com/mrw_images/els/articles/a0003206/image_n/
nfgz002.gif).

the resistances are in series, the total resistance can be


thought of as just the sum of the two resistances. Hence the
mass-transfer (diusion analogy) model developed in Section 3
can be easily augmented to account for this eect:
264

place at the potential rate (Ep) since water is freely available


for evaporation from the leaves, and the contribution from the
leaf surface is often called interception loss. The amount of
precipitation that directly reaches the surface (i.e. is not
intercepted) is called throughfall. Additional water can fall to
the surface via drainage if the interception storage fills up in
conjunction with continuing precipitation.

M OVIE 8.1 Animation of satellite-derived global LAI showing


seasonal and interannual variability (from

earthobservatory.nasa.gov/Features/LAI/LAI3a.php).

E=

(q surf qa )
ra + rc

(8.4.3)


As mentioned above, the stomatal openings open/close
depending on environmental stresses being experienced by
the plant. So far these eects have not been included
explicitly in the model. The four stresses that are usually
accounted for are the incoming shortwave radiation (PAR), air
humidity, air temperature, and rootzone soil moisture. These
eects are usually modeled in the stomatal resistance
mentioned above as:
rsmin
rs =
(8.4.4)
fR fT fe f
s

which represents the evapotranspiration from a vegetated


surface. Note that the sensible heat flux is not regulated by
the stomata since it is still driven by temperature gradients,
hence the model for H is that given in Equation (8.3.16). This
particular form of model is often referred to as a one-source
model, assuming a homogeneous vegetation covered surface.
More complicated models can use a two-source model, where
evaporation comes from both vegetation and soil.

It should also be noted that the vegetation itself can
intercept water or snow as discussed in Chapter 6. In the case
where liquid water is on the leaves, the evaporation takes

rz

where rsmin is the (plant-dependent) minimum stomatal


resistance and the terms in the denominator are stress
factors that vary between 0 and 1 and depend on incoming
shortwave radiation, air temperature, vapor pressure deficit,
and rootzone soil moisture respectively. These functions are
constructed in a way that in the limit approaching fi = 0
(where i corresponds to the dierent variable subscripts in
Equation (8.4.4)) corresponds to fully stressed conditions and
the limit approaching fi = 1 corresponds to fully unstressed
conditions. For example, when rootzone soil moisture becomes
265

fR (R ) =
s

1.105Rs

1.007Rs + 104.4

0 R 1100 W m

-2

(8.4.5)

fe (e) = 1 0.000238e,

0 e 4200 Pa

(8.4.6)

Ta (40 Ta )1.18
fT (Ta ) =
,
a
690

0 Ta 40 C

(8.4.7)

#
1,
%
%%
wp
,
f (rz ) = $
rz

% fc
wp
%
0,
%&

fc
wp fc

(8.4.8)

wp

which are plotted in Figure 8.8.

F IGURE 8.8 Illustrative example of environmental stress factors used in canopy resistance model.

low, the stress factor for soil moisture will approach zero,
which will increase the stomatal resistance, and ultimately
make the transpiration equal to zero. The specific forms of
these functions are generally empirically-based. One example
of such equations is the set given by (Dingman, 2008):


For the incoming shortwave function, the stress function
goes to zero (i.e. stoma close) when the radiation is zero (at
night) and it goes to 1.0 when the stoma are fully open during
a sunny day with a significant amount of PAR. The air
temperature stress function is generally a function with a
peak or plateau region, which represents the optimal
temperature conditions for the plant. If the temperature gets
too cold or too hot the stoma will close completely. The vapor
pressure deficit (VPD) function is 1.0 when the VPD is zero
and decays as the VPD increases. This is because evaporation
losses will generally be high when the VPD is high and low
when the VPD is low. For soil moisture, when the rootzone
266

soil moisture is at or below the wilting point (i.e. 0.1 in Figure


8.8) the stoma will have to shut completely. When the
rootzone moisture is above the field capacity (i.e. 0.4 in Figure
8.8) then water is plentiful and the stoma can be open. Given
that all of these variables are functions of time, the stress
functions themselves are functions of time as is the stomatal/
canopy resistance.

In summary, the presence of vegetation primarily changes
evapotranspiration by adding an additional regulation on the
flux of vapor at the surface. Vegetation plays an active role in
the water vapor loss (transpiration) via the regulation of
stomatal openings in the leaves. When under perfectly
unstressed conditions, the stomatal resistance is equal to the
minimum stomatal resistance, which is then scaled by the LAI
so that the additional regulation is entirely determined by
plant-dependent parameters. In the more general cases,
environmental conditions (both meteorological and soil
moisture) can generally increase the canopy resistance as the
environmental conditions are sub-optimal for vegetation
photosynthetic functioning. In the extreme case where
environmental conditions are severely sub-optimal (e.g.
rootzone soil moisture becomes very dry approaching the
wilting point of the plant), transpiration can be shut-o
completely via the complete closing of the stoma.

E XAMPLE 8.4.1
A vegetated surface is adjacent to the bare soil
surface described in Example 8.3.1. The
vegetation consists of a grass surface with a
characteristic height of 5 cm, a leaf area index of
2.5 and a minimum stomatal resistance of 70 s/
m. The incident shortwave radiation at the
surface is measured to be 900 W m-2. Assume the
meteorological conditions and surface conditions
for the vegetated surface are the same as that in
Example 8.3.1. Compute the evaporation rate
from the vegetated surface.
The primary dierence for a vegetated surface is the
additional resistance due to stomatal control on vapor
loss. This requires computation of each of the
environmental stress factors in the stomatal resistance
function. Based on the saturated soil and meteorological
conditions and using functions illustrated in Figure 8.8
(Equations (8.4.5)-(8.4.8)), the stress factors can be
estimated as:

1.105(900 W m 2 )
fR (R ) =
= 0.98
2
s
1.007(900 W m ) + 104.4

(27.2 C)(40 27.2 C)1.18


fT (Ta ) =
= 0.80
a
690
f (rz ) = 1
rz

267

E XAMPLE 8.4.1 ( CONTINUED )

E XAMPLE 8.4.1 ( CONTINUED )

For the vapor pressure deficit term:

LE = (2.5 10 6 J/kg)(0.88 10 4 kg/m 2 /s)

e = es (Ta ) ea = es (1 RH ) = (36.9 mb)(1-0.69)


= 11.44 mb
fe (e) = 1 0.000238(1144 Pa) = 0.73

= 219 W m 2
which is less than the bare soil evaporation.

The canopy resistance is then given by:

rc =

rsmin
(70 s/m)
=
LAI(fR fT fe f ) (2.5)(0.98 0.80 0.73 1)
s

rz

= 49 s/m
The aerodynamic resistance also changes due to the
increased roughness of the vegetated surface, i.e.:
2

raN

( " 2 m 0.7(0.05 m)%+


* ln $
'0.1(0.05
m)
&,
) #
=
= 38 s/m
2
(0.4) (5.8 m/s)

which shows that the resistance is reduced due to the


increased roughness. The evaporation rate is then given
by:

(23.1 16.3) 10 3 kg/kg


E = (1.12 kg/m )
(38 + 49 s/m)
3

= 0.88 10 4 kg/m 2 /s

268

S ECTION 5

Using a finite dierence approach this can be written as:

Additional ET Models

c p (T2 T1 )
B=
Lv (q 2 q1 )


The mass-transfer models described in the last two
sections provide instantaneous (i.e. over 10-30 min.) models
for LE and H (over bare soil/open water or vegetated surface
respectively), but require measurements/estimates of Tsurf,
qsurf, Ta, qa, and V (as well as characterization of other surface
properties). Other methods/models have been developed for
cases where dierent measurements may be available.

The first approach discussed in this section is the socalled Energy Balance Bowen Ratio (EBBR) method. The
Bowen ratio is defined as the ratio of the sensible heat to the
latent heat flux: B = H/LE. Note that the Bowen ratio is a
parameter indicative of the state of the surface and its role in
the partitioning of energy. When the surface is very dry, most
of the available energy will go into sensible heat (high Bowen
ratio). When the surface is wet, most of the available energy
will go into latent heat flux (low Bowen ratio). From the
original diusion analogy models in Equations (8.3.2) and
(8.3.3), this can be written as:

c p T / z
B=
Lv q / z

(8.5.1)

(8.5.2)

which implies that if measurements of temperature and


humidity are available at two dierent levels (i.e. on a tower),
then one can estimate the Bowen ratio at a given time. The
method then invokes the surface energy balance equation,
which can be written in terms of the Bowen ratio:

Rn G = LE + H = LE(1 + B)

(8.5.3)

which can be solved for the latent heat flux:

LE =

Rn G
1+B

(8.5.4)

So given an estimate of the Bowen ratio and estimates of the


available energy (Rn - G), one can estimate the actual
evapotranspiration. Note this does not explicitly depend on
soil moisture measurement, wind speed, or other surface
properties aside from those needed to estimate Rn - G. Also,
no separate treatment is required for bare soil or vegetated
surfaces. Additionally, once latent heat flux is estimated, the
sensible heat flux can be estimated from the definition of the
Bowen ratio:

H = B LE

(8.5.5)

where the gradients are determined between two levels in the


atmosphere (i.e. could be over soil, open water or vegetation).
269

E XAMPLE 8.5.1
A meteorological tower takes temperature and
humidity measurements at two levels (where level
1 is at 5 m and level 2 is 10 m above the surface)
and available energy measurements:
Air temperature (level 1) =27.2C
Air temperature (level 2) =24.2C
Air specific humidity (level 1) =12 g/kg
Air specific humidity (level 2) =11 g/kg
Available energy (Rn - G) = 500 W m-2
Estimate the latent and sensible heat fluxes from
the surface.
From the two-level measurements, the Bowen ratio can
be estimated as:

B=

1004 J/kg/K (27.2 24.2) K


= 1.2
2.5 10 6 J/kg (12 11) 10 3

Based on the available energy, the latent and sensible


heat fluxes are:

LE =

500 Wm
= 227 Wm 2
1 + (1.2)
2


The second model described here is the so-called Penman
model. Penman (1948) combined the mass-transfer models
with surface energy balance to derive a model for potential
evaporation. The primary development of the model is as
follows:
1. Assume that the surface humidity in the mass-transfer
model for evaporation is equal to the saturated specific
humidity (this implies potential evaporation):

q surf = q s (Tsurf )

(8.5.6)

2. Linearize the saturated specific humidity around the air


temperature via a Taylor series expansion:
2
dq s
1 d qs
2
q s (Tsurf ) = q s (Ta ) +
(Tsurf Ta ) +
(T
T
)
+ ...
surf
a
dT T
2 dT 2
T
a

Keeping only the linear terms (i.e. the first two terms on the
right-hand-side) yields:

q s (Tsurf ) q s (Ta ) +

dq s
(T Ta )
dT T surf

(8.5.7)

where by definition:

dq s des Lv es
=
=
dT p dT p Rv T 2

(8.5.8)

H = (1.2)(227 Wm 2 ) = 272 Wm 2
both of which are away from the surface.

is simply the Clausius-Clapeyron equation and can be directly


evaluated (or tabulated) as a function of air temperature. It is
270

often written using the following simplified notation:

dq s des
=
=
dT p dT p

(8.5.9)

3. Rearrange the mass-transfer equation for sensible heat into:

(Tsurf Ta ) =

Hra
c p

(8.5.10)

4. Rearrange the SEB into: H = Rn + LE - G and substitute


into Equation (8.5.10).
5. Substitute the expression for (Tsurf - Ta) into the linearized
Equation (8.5.7), which can then be substituted into the
mass-transfer equation for E (Equation (8.3.14)).
6. Finally, solving for LE yields the Penman equation for
potential evaporation (i.e. from bare soils or open water
surfaces):

(Rn G) + v q

ra
LE p =

1+

psychrometric constant =
q = qs (Ta ) qa

(8.5.11)

pc p

Lv


The Penman model provides potential evaporation rates
given estimates of available energy, reference-level air

temperature, humidity, wind speed, and surface roughness. It


also shows two key drivers of potential evaporation, the first
term in the numerator is energy driven, while the second term
is associated with atmospheric demand depending on how
humid the air is. As mentioned in Section 3, the actual
evaporation is often then modeled as:

LE = ( )LE p

(8.5.12)

E XAMPLE 8.5.2
A reservoir has a nearby meteorological station
that takes the following measurements:
Air temperature =29.2C
Air relative humidity =76%
Surface air pressure = 980 mb
Windspeed (at 2 m height) = 10.3 m/s
Available energy (Rn - G) = 383.2 W m-2
Estimate the evaporation rate (in mm/day).
Assume the characteristic wave height on the
reservoir surface is 3 cm.
Note that in this case there is no data characterizing the
surface temperature. (If the surface temperature was
known, we could reasonably assume that the surface
specific humidity was equal to the saturated specific
humidity at that temperature.) Hence the mass-transfer
271

E XAMPLE 8.5.2 ( CONTINUED )

E XAMPLE 8.5.2 ( CONTINUED )

model cannot be used in this case. Since the problem


involves evaporation from an open-water surface we
know that the evaporation rate should equal the
potential evaporation rate. Given the measurements of
available energy, the Penman model is applicable to this
problem:

The Clausius-Clapeyron and psychrometric terms are


given by:

(Rn G) + v q

ra
LE p =

1+

where each term needs to be computed. The air density


can be computed from the ideal gas law for these
conditions, which yields a density of 1.11 kg/m3. The
aerodynamic resistance (under neutral conditions) is
equal to (from Equation (8.3.9)): ra = 25.6 s/m. The
specific humidity deficit is a function of air temperature
and relative humidity as given by:
# 2.5106 J/kg #
&&
1
1
es (Ta ) = (6.11 mb) exp %

%
((
$ 461 J/kg/K $ 273.16 K 302.35 K ''
= 41.57 mb
ea = (0.76) ( 41.57 mb) = 31.59 mb

q =

(0.622)
e =
(4157 3159 Pa) = 6.3103 kg/kg
p
(98000 Pa)

Lv es (2.5 10 6 J/kg) (4157 Pa)


=
=
Rv T 2
(461 J/kg/K) (302.35 K)2

= 246.6 Pa/K
(98000 Pa)(1004 J/kg/K)
=
= 63.3 Pa/K
(0.622)(2.5 10 6 J/kg)
246.6 Pa/K
=
= 3.9

63.3 Pa/K
Putting all terms together yields the latent heat flux:

(3.9)(383.2 W/m 2 )
LE p =
+
1 + 3.9
(1.11 kg/m 3 )(2.5 10 6 J/kg)
(6.3 10 3 )
(25.6 s/m)
1 + 3.9
= 444 W/m 2
The equivalent evaporation rate is then given by:

444 W/m 2
1 m3
E[mm/day] =

(2.5 10 6 J/kg) 1000 kg


1000 mm 86400 s

= 15.3 mm/day
1m
1 day

272


Other models have followed directly from Penman. For
example over a large moist surface with minimal advection the
air will become saturated under continued evaporation, i.e.:

q 0
Under these conditions one can define the equilibrium
evaporation, i.e. that driven solely by energy considerations:

(R G)
n
LEe =

1+

resistance term in the mass-transfer model. Following that


approach, one can show:

(Rn G) + v q

ra
LE =
r
1+ + c
ra

(8.5.15)

where experimentally it has been found that in many cases


the multiplicative coecient has a value of 1.26. Given this,
the Priestly-Taylor model is often used to estimate potential
evaporation (and does not require humidity or wind speed
measurements).

In functional form this is identical to the Penman model


except for the additional (last) term in the denominator which
includes the canopy resistance. In fact, in the limit that the
canopy resistance goes to zero, this equation reduces to the
Penman model as expected. Despite these similarities, some
care needs to be taken in terms of the meaning of the ET
predicted by the Penman-Monteith equation. Note that in its
general form the canopy resistance has a soil moisture stress
factor embedded in it (Equations (8.4.2) and (8.4.4)). Hence if
one wishes to use Penman-Monteith to model potential
evapotranspiration, then one needs to set the soil moisture
stress factor to 1.0 (which is equivalent to soil moisture not
being a limiting factor). The soil moisture impact on actual
evaporation could then be considered via Equation (8.5.12).
Alternatively, if one keeps the soil moisture stress function in
the canopy resistance, then the equation will not predict
potential evapotranspiration, but more closely actual
evapotranspiration.


Finally, it must be noted that the Penman model does
not consider vegetation cover. To account for the impact of
vegetation, the Penman-Monteith model is an extension which
follows the same recipe as above, but considers the canopy


The set of models described in this and the previous
sections, take various approaches to predict ET based on
various inputs. The model chosen depends on the surface type
as well as the measurements that are available. The specific

(8.5.13)

The Priestly-Taylor model postulates a relationship between


potential and equilibrium evaporation, i.e.:

(Rn G)

LE p = LEe =

1+

(8.5.14)

273

measurements that are available are often the result of


practical factors like cost of equipment and operation and
maintenance costs, proximity to existing infrastructure, etc.
E XAMPLE 8.5.3
A vegetated region adjacent to the reservoir in
Example 8.5.2 has the same meteorological data.
Based on its roughness and vegetation
characteristics, the corresponding aerodynamic
and canopy resistances are ra = 20 s/m and rc =
50 s/m. What is the expected latent heat flux
and evaporation rate (in mm/day)?
Based on the data, and since the surface is vegetated,
the Penman-Monteith model can be used:

(3.9)(383.2 W/m 2 )
LE =
+
50 s/m
1 + 3.9 +
20 s/m
(1.11 kg/m 3 )(2.5 10 6 J/kg)
(6.3 10 3 )
(20 s/m)
50 s/m
1 + 3.9 +
20 s/m
= 326 W/m 2
which is a lesser value than the potential rate from the
open-water surface in Example 8.5.2 due to the
additional stomatal control, despite having a lower ra.

274

S ECTION 6

MOD-WET Codes

Bulk Richardson number:



richardson_number.m


Relevant functions based on concepts introduced in
this chapter include:
Aerodynamic resistance to turbulent transport:

aero_resistance.m

psychrometric_constant.m

Surface energy balance model (prognostic solver):



soil_SEB_solver_prognostic.m
Stability correction factors:

stab_corr_factors.m

Canopy resistance due to stomatal control:



canopy_resistance.m
Slope of Clausius-Clapeyron equation:

clausius_clapeyron_slope.m
Energy Balance Bowen Ratio Method for ET estimation:

EBBR.m
Mass transfer model for evapotranspiration (and sensible
heat):

mass_transfer.m
Penman model:

penman.m
Penman-Monteith model:

penman_monteith.m
Psychrometric constant calculation:
275

S ECTION 7

Conceptual Questions
1. Describe the physical meaning of potential evaporation.
How does it generally compare to actual evaporation?
2. Is pan evaporation a measure of actual evaporation or
potential evaporation?
3. For a watershed mass balance, write the long-term average
evaporation as a function of other long-term average mass
fluxes.
4. Using an energy balance, write the long-term average
evaporation as a function of other long-term average energy
fluxes.
5. How do you convert between latent heat flux (in units of W
m-2) and evaporation mass flux (in units of kg m-2 s-1)? How
do you convert between an evaporation mass flux to units
of depth per unit time (i.e. mm/day)?
6. In the mass-transfer model for evaporation from bare soil,
what time-varying gradient is evaporation proportional to
(i.e. the driver of the flux). What gradient is sensible heat
flux proportional to?
7. What type of function describes the form of the mean
horizontal wind profile (i.e. as a function of z) in the neutral
surface layer?

8. What physical process in the near-surface layer of the


atmosphere is chiefly responsible for the transport of vapor
(or heat) away from the surface?
9. Describe what the aerodynamic resistance physically
represents.
10. How does the aerodynamic resistance change if wind
velocity at a reference-level increases? Explain why this is
and what impact it has on evaporation and sensible heat
fluxes.
11. What is a neutral surface layer, i.e. what does neutral
refer to?
12. Describe the basic physical mechanism whereby vegetation
exerts an additional control/regulation on
evapotranspiration.
13. What does canopy resistance refer to?
14. Describe what leaf area index is.
15. Describe the meaning of the environmental stress factors
used in models for canopy resistance.
16. If environmental conditions are such that there is no stress
on the plant, what will the canopy resistance reduce to?
17. Does high soil moisture stress refer to high or low soil
moisture? Explain.

276

18. Identify whether the EBBR, Penman, and PenmanMonteith models give actual or potential
evapotranspiration. Explain.

277

S ECTION 8

Sample Problems
Problem 8.1. Meteorological and surface measurements
collected at noon from a bare soil field site are:
Air temperature = 26.8C
Air relative humidity = 69%
Surface air pressure = 973 mb
Wind speed at 2 m height = 3.0 m s-1
Soil surface temperature = 26.2 C
The measurements are taken on the first sunny day after a
rainy period (you can assume the soil is saturated). The soil
roughness elements have a characteristic height (h) of 4 cm.
a) Using only the measurements/information provided, which
model could be used to estimate the instantaneous surface
evaporation? Which model/s cannot be used if you only have
the data provided? Explain why.
b) What additional information/input data would you need if
the soil were not saturated?
c) What additional information would you need if the surface
was vegetated?
d) In general, qsurf over bare soil can be modeled as:

q surf = q s (Tsurf );

Consider two cases: 1) a saturated soil where the


multiplicative coecient is equal to 1.0 and 2) a water-limited
case where the multiplicative factor is equal to 0.80. Use an
appropriate model to compute the latent and sensible heat
flux at the measurement time for each of these cases. How do
the values of the fluxes between the water limited case and
the moist soil case compare?
e) What is the equivalent evaporation rate in mm/day for the
saturated soil case? Is this actual or potential evaporation?
Problem 8.2. A meteorological station/tower is set up at a
field site to take the following measurements: net radiation
(Rn) at the surface, ground surface heat flux (G), and air
temperature and specific humidity at two dierent heights
within the atmospheric surface layer (T1 and T2 and q1 and
q2). Data at 1:00 PM for a particular day are given by:
Rn = 360 W m-2
G = 41. W m-2
T1 =27.C
q1 = 11. g/kg
T2 = 24.C
q2 = 8.6 g/kg
a) Use an appropriate model to compute the latent and
sensible heat flux at the measurement time.
b) What is the equivalent evaporation rate in mm/day? Is
this actual or potential evaporation? Explain.

0 ( ) 1
278

Problem 8.3.
a) Describe the meaning of potential evaporation. What
conditions are necessary for potential evaporation from a
surface? How does potential evaporation dier from actual
evaporation?
b) Evaporation from a reservoir can be obtained using the
Penman equation. If the available energy at the surface of the
reservoir is 220 W m-2, aerodynamic resistance is 30 s m-1, and
air specific humidity is 8 g/kg, what is the change in
evaporation (in mm/day) from the surface if the air
temperature is reduced from 20C to 15C as a cold air mass
passes over the reservoir? You may neglect any changes in net
radiation at the surface as a result of the dierent air
characteristics.
c) Qualitatively, explain how an increase in the surface
temperature would aect the evaporation rate from the
reservoir.
d) Qualitatively, explain how a lower humidity would impact
the evaporation rate from the reservoir.
Problem 8.4. Suppose the following measurements are made
over a bare soil surface (with a characteristic height of 2 cm
for surface roughness elements):
Reference-level (i.e. 2 meter height) wind-speed: 3 m/s
Reference-level air temperature: 25.9C
Reference-level air vapor pressure: 2217 Pa
Surface soil temperature: 37.9 C

Surface soil specific humidity: 15 g/kg


Surface pressure: 1000 mb
Estimate the surface latent and sensible heat fluxes under
these conditions. What is the corresponding evaporation rate
at this time (expressed in units of mm/day)?
Problem 8.5. A farmer is deciding between two crops to
plant for the upcoming growing season. For the two crops
being considered, the dierence in profit for the farmer is
exclusively related to the water needed to grow the plants.
Therefore, the farmer needs to know how much
evapotranspiration will occur from each crop on a typical day.
The characteristics corresponding to each crop type is listed
below:

CROP

STOMATAL
RESISTANCE (S/M)

VEGETATION
HEIGHT (CM)

30

30

60

100

From a meteorological station, measurements at 2 m height


are: Ta =20C, V = 1 m s-1, RH = 65%, and ps = 1000 mb.
The net radiation at the surface is 50 W m-2 and the ground
heat flux is 2 W m-2. Use the Penman equation to answer the
following:
a) What would the evaporation from each crop (in mm day-1)
be under these conditions?
279

b) How and why are they dierent?


c) Which crop should the farmer plant?

280

Chapter 9

Groundwater
Flow

S ECTION 1

Learning Objectives
By the time you finish this chapter you should be able to:
1. Write down Darcys Law for specific discharge in
saturated flow; show how it is a special case of the
expression used in unsaturated flow
2. Compute the average fluid velocity through an aquifer
based on the Darcy flux and soil porosity
3. List the key dierences between unconfined and confined
aquifers
4. Describe the dierent mechanisms by which water is
stored in unconfined and confined aquifers
5. Define the respective storage parameter relevant to
unconfined and confined aquifers
6. Write down the general 3D groundwater flow equation for
confined aquifers
7. State the key dependent variable that is being solved for
in any groundwater flow problem (i.e. once this variable is
known anything else can be solved for)

9. Solve the groundwater flow equation via integration for


simple steady-state one-dimensional boundary value
problems
10. Understand the dierence and meaning of a fixed-head
and no-flow boundary condition
11. Use the principle of superposition to decompose and
solve more complicated groundwater flow problems
12. Write down the 2D flow equation in cylindrical
coordinates for flow to a single pumping well
13. Understand the dierence between a monitoring well and
a pumping well
14. Solve the steady-state flow equation toward a single
pumping well for unconfined and confined aquifers
15. Use the single-well solutions to solve for aquifer
properties (conductivity/transmissivity) given measured
head values
16. Solve the transient flow equation toward a single
pumping well using the Theis equation
17. Use superposition to solve for head/drawdown in
problems with multiple wells, impermeable or fixed-head
lateral boundaries

8. Write down the general 2D groundwater flow equation


(using the Dupuit approximation) for unconfined and
confined aquifers
282

S ECTION 2

Basic Groundwater Characteristics



The term groundwater is generally used to refer to the
study of subsurface flow in saturated soil media. Such
groundwater reservoirs constitute approximately 30% of the
total global freshwater and approximately 99% of the total
global liquid freshwater, and hence form one of the key water
resources for water supply. We will start this chapter with
some of the basic definitions used in groundwater. It should
be noted that many are shared with unsaturated flow
problems (also flow in porous media) with the dierence being
that all the pores are filled with water. A more comprehensive
treatment of groundwater processes is given in Bear (2007).

An aquifer is a geologic unit (often of unconsolidated
soil) that is capable of storing and transmitting significant
amounts of water. A schematic of the subsurface is shown in
Figure 9.1. Aquifers are generally characterized into one of
two types: unconfined aquifers and confined aquifers.

In unconfined aquifers (generally near the surface just
below the unsaturated zone) the pores saturate via water
pooling up over time on top of an impervious or low
conductivity layer. These low conductivity layers that prevent
downward flow are generally referred to as aquicludes or
aquitards. The upper boundary for an unconfined aquifer is

F IGURE 9.1 Illustration of various layers in subsurface including vadose zone, an unconfined and confined aquifer and an
aquiclude.

the water table, which is open to the atmosphere and


therefore has a (gauge) pressure equal to zero (i.e.
atmospheric) at the water table. The main source of water for
unconfined aquifers is direct recharge from the overlying
unsaturated zone. Because the water table is a free surface,
changes in storage in unconfined aquifers correspond directly
to an increase or decrease in the water table elevation. In this
way, unconfined aquifers are analogous to streams where
unsteady flow involves changes in stream elevation. The
piezometric (or potentiometric) surface, i.e.:
283

h=

p
+z
g

(9.2.1)

corresponds directly with the height of the water table, where


z is defined relative to a datum.

In confined aquifers, the saturated soil is bounded above
and below by low conductivity layers. Often these layers may
be clay soil layers above and impervious bedrock below. As a
result, the physical boundaries of the confined aquifer are
essentially fixed, making the flow more analogous to pipe flow.
The source of water in confined aquifers is recharge from
upstream locations or in some cases via leaky aquitards
(Figure 9.2). As a result of the configuration (i.e. fixed
boundaries above/below), water in confined aquifers is often
under high pressure. By definition the piezometric surface is
generally above the top of the aquifer. In some cases, the
piezometric surface is actually above the ground surface
(Figure 9.2), in which case if a well were drilled into the
confined aquifer, water would flow up to the surface without
the need of a pump (artesian well).

Flow in aquifers is simply flow through porous media and
therefore Darcys Law still applies for the specific discharge/
Darcy flux vector:

q = Kh

(9.2.2)

where the conductivity is no longer a function of moisture


content since the soil is saturated (i.e. K=Ks). The gradient
operator is used in recognition of the fact that in general the

F IGURE 9.2 Typical aquifer configuration showing piezometric surface for unconfined and confined aquifers, which can be
measured via monitoring wells (from
artinaid.com/2013/04/types-of-aquifers) .

piezometric head is a 3D function in space, i.e. h=h(x,y,z,t).


The components of the Darcy flux vector can be written as:

qx K x

h
,
x

qy = K y

h
,
y

q z = K z

h
z

(9.2.3)

The pore velocity is another variable that is often of interest


in groundwater flow as it controls the advection of pollutants.
It represents the average velocity of water in the pores and
can be written as:

v=

Q
q
=
sA s

(9.2.4)
284

where the porosity in the denominator corresponds to the fact


that only that fraction of the cross-sectional area of flow is
occupied by pore space. As was discussed previously, while
flow in the unsaturated zone is primarily 1D (z-direction),
groundwater flow is primarily 2D (in the x-y plane). In such
cases this implies that qz = 0 and hence the head is constant
with depth in the aquifer.

The storage characteristics of aquifers dier depending
on the type of aquifer. For an unconfined aquifer, the storage
change corresponds directly to a change in water table level.
The storage parameter used to characterize unconfined
aquifers is the specific yield or the storage coecient: Sy
which is defined as the volume of water released per unit
decline in water table (per unit area). Physically, the specific
yield is dimensionless and varies between zero and the
porosity of the soil, with typical values between 0.05 and 0.35.
The specific retention is the dierence between the porosity
and specific yield.

In confined aquifers the storage mechanism is dierent
since there is no water table that can move up or down.
Rather storage changes correspond mainly to compression of
the aquifer as the weight of the overlying material is
transferred from liquid to solid grains when water is removed
(or vice versa). In essence this amounts to storage changes as
a result of porosity changes. The storage parameter used to
describe this behavior is the so-called specific storage:

Ss change in porosity/change in piezometric head

F IGURE 9.3 Map showing the large-scale groundwater aquifers across the U.S. (from USGS;

pubs.usgs.gov/ha/ha730/ch_a/gif/A004_us.gif).

which has dimensions of L-1. Typical values range between


110-5 and 110-3 m-1.

Together the storage and flow characteristics will be
combined to derive the groundwater flow equation. Some
examples of aquifer types and distribution in the U.S. are
shown in Figure 9.3. Of particular note is the Ogallala Aquifer
in the midwestern U.S. Much of the grain belt of the U.S.
draws its irrigation water via pumping from this one very
large aquifer.
285

E XAMPLE 9.2.1

A square island has an unconfined aquifer with a


piezometric head distribution (h(x,y); in meters)
as shown in the figure below. The aquifer is
homogeneous and has a saturated hydraulic
Problem #3: (6 points ~ 20 min.)
conductivity of 5.0 m/d and a porosity of 0.35.
The steady-state head field for a confined aquifer under a 5 km x 5 km square island (with a
is the
Darcy
flux and
poreis velocity
themap below.
homogeneousWhat
and isotropic
hydraulic
conductivity
of 1 m/day)
plotted in theat
contour
datum is such that the head on all sides of the boundary of the island is 0 m.
location: x = 3 km, y = 2 km?

E XAMPLE 9.2.1 ( CONTINUED )

The

middle of the island with values above 25 m, but below


30 m (since there is no 30 m contour). The head
generally decreases to zero at the aquifer boundaries.
The water will flow down the head gradient (i.e. from
high to low head). Based on the distribution of head, the
flow pattern will also vary in space. The Darcy flux is
given by:

q = Kh
where the gradient must be estimated from the head
distribution. The head gradient is mathematically in the
direction of steepest ascent, which graphically is
perpendicular to head contours. The negative sign in the
Darcy flux indicates that flow is in the down-gradient
direction. The down gradient direction at x = 3 km, y =
2 km is shown in the annotated figure below. The
gradient can be estimated using a finite dierence:

x =3km,y=2km

h
5m

= 0.01
L x =3km,y=2km 500 m

Based on the hydraulic conductivity, the magnitude of


the Darcy flux is given by:

q = (1.0 m/d)(0.01) = 0.01 m/d

a) On the map,
sketch
the flowof
direction
water starting
from point
The
pattern
head of
shows
a general
peak(3km,
near3km)
theto the edge of the island. in
(2 pts.)

b) Where is the Darcy flux the maximum and what is its approximate value? Explain your reasoning for
choosing where the flux is largest. (4 pts.)

the direction shown in the figure below. The pore


velocity is in the same direction with a magnitude:
286

E XAMPLE 9.2.1 ( CONTINUED )

Problem #3: (6 points ~ 20 min.)

The steady-state head field for a confined aquifer under a 5 km x 5 km square island (with a
homogeneous and isotropic hydraulicqconductivity
of 1 m/day) is plotted in the contour map below. The
0.01 m/d
v
=
=
= the
0.029
datum is such that the head on all sides of the boundary of
islandm/d
is 0 m.

0.35

+"

a) On the map, sketch the flow direction of water starting from point (3km, 3km) to the edge of the island.
Note that the Darcy flux and velocity will vary in space.
(2 pts.)

For example, the Darcy flux will be highest where the

b) Where is the Darcy flux the maximum and what is its approximate value? Explain your reasoning for
contours
closest
(i.e. largest gradient), which occurs
choosing where
the flux isare
largest.
(4 pts.)

near the point x = 2.5 km, y = 5 km. At this location


Problem #4:
(12head
pointsgradient
~ 30 min.)is about 3 times larger than above,
the
A farmer hasmaking
a row irrigation
system asflux
shown
in Figure
1 below.
The soil
is bounded on both sides by
the Darcy
and
velocity
3 times
larger.

drainage ditches which allow water to be removed from the soil through lateral groundwater flow. The
water in the drainage ditches is maintained at a depth H0. The irrigation rate I is applied at the top of the
soil profile. The water which is not evaporated via transpiration by the crops recharges (at a rate R) the
unconfined aquifer below (which is bounded by a horizontal impermeable layer at depth D). The crops
being grown by the farmer have a rooting depth of d and have an average evapotranspiration rate E. The
farmer must decide the irrigation rate such that the rootzone remains unsaturated (i.e. the water table

287

S ECTION 3

Ss

(9.3.3)

h
= (q)
t

Development of Groundwater Flow Equation

which can be expanded to:


The groundwater flow equation is analogous to the flow
equation for unsaturated flow, but under saturated conditions.
Here we will start with the general 3D flow equation using
mass balance and then simplify it to 1D and 2D versions as
special cases.

Substituting in Darcys Law yields the general form of the 3D


groundwater flow equation:


The statement of mass balance can be written as (same
starting point as in unsaturated flow):

M
= ( wq dx dy dz)
t

(9.3.1)

where again the left-hand-side term is the change in mass


storage and the right-hand-side term is the convergence of flux
(both for a dierential volume of soil). From the definition of
specific storage, we can expand the storage term as:

M
h
= wSs dx dy dz
t
t

(9.3.2)

which transforms the mass storage change into a piezometric


head change. Plugging this into Equation (9.3.1) and
assuming the density of water is constant yields:

Ss

Ss

q
q
q
h
= x y z
t
x
y
z

h

h
h
=
K
+
K
+
t x x x y y y z

(9.3.4)

h
K
z z

(9.3.5)

This is a second order PDE in h. For a general solution


h(x,y,z,t), there also needs to be two boundary conditions
(BCs) per coordinate and one initial condition (IC) specified.
Once h is known, then everything about the flow field can be
computed. In other words, the key point is that in any
groundwater flow problem what we are really solving for is the
piezometric surface (h) first. Once known we can determine
everything about the flow itself. This is analogous to the
unsaturated flow problem, where when soil moisture is known
everything else can be determined. In that case the head is a
function of soil moisture, whereas in groundwater it is not.

Before proceeding to the development of the 2D flow
equation, some special cases that are often encountered are
worth mentioning. First, an aquifer is said to be homogeneous
if its conductivity field does not vary in space, i.e.:
288

K x K y K z
K i K i (x,y,z)
=
=
=0
x
y
z
Second, an aquifer is said to be isotropic if there is no
directionality to the conductivity field, i.e.:

K x = Ky = K z = K
Finally, if the flow is in steady-state, then there are no head
variations in time, i.e.:

h
=0
t
Given these special conditions, one can derive simplified forms
of the 3D flow equation. For example, for a homogeneous/
isotropic aquifer, we can write the governing equation as:

2 h 2 h 2 h
h
Ss
=K 2 + 2 + 2
t
y
z
x

(9.3.6)

or for an homogeneous/isotropic aquifer in steady-state:

2 h 2 h 2 h
2
2 + 2 + 2=h =0
y
z
x

(9.3.7)

which is the well-known Laplace equation, which is seen not


only in groundwater problems but many other fields including
heat transfer.


In many natural groundwater systems, the flow is largely
horizontal (i.e. in the x- and y-directions) with little to no
vertical flow. As such, in many cases solving the 2D GW flow
equation is sucient. Here we develop the 2D flow equation
from the general 3D equation shown above. The usual
mechanism for doing so if via the so-called Dupuit
approximation, which essentially involves integrating the 3D
equation with respect to z, assuming no vertical flow inside
the aquifer (qz = 0) and that the aquifer has a horizontal lower
boundary. Together these mean that h=h(x, y, t). Note that
the lack of dependence on z implies that h is constant in any
vertical plane. This is consistent with the no vertical flow
assumption (i.e. if h varied with z then there would be a
gradient and therefore a Darcy flux). In such cases, as depth
increases into the aquifer, there are one-to-one tradeos
between the pressure and elevation head terms so that the
head is constant in the vertical. In other words, at the surface
the pressure is zero so the piezometric head is equal to the
elevation head. As one moves down vertically into the aquifer
the pressure increases such that the pressure head increases by
the exact amount that the elevation head decreases.

To develop the 2D GW flow equation we integrate
Equation (9.3.5) from z = 0 to z = H, where H = h (i.e. the
water table) in the case of an unconfined aquifer and H = b
(i.e. the aquifer thickness) in the case of an confined aquifer:
H

H

h
h
h
S
dz
=
K
+
K

0 s t
0 x x x y y y + z

h
K z z dz

289

This equation consists of four terms (one on the left-hand-side


and three on the right-hand-side). We will focus on integration
of them one-by-one. First, term #1:
H

Ss
0

h
h
dz =
S dz
t
t 0 s

h
[unconfined aquifers]
t
h
h
=Ss b
= S ; [confined aquifers]
t
t
= Sy

where the confined aquifer parameter: S = Ssb is called the


storativity. Term #2 is:
H

K
dz
=
0 x x x x

h
h
0 K x x dz = x K x x z

0 z


h
K
h
[unconfined aquifer]
x x x


h
h
K
b
=
T
[confined aquifer]
x x x x x x

where the confined aquifer parameter T = Kb is called the


transmissivity. Term #3 is similarly:
H
H


h

h
h
0 y Ky y dz = y 0 Ky y dz = y Ky y z


h
=
K
h
[unconfined aquifer]
y y y


h
h
K
b
=
T
[confined aquifer]
y y y y y y

h
h
K
z z dz = K z z

= Kz

h
h
Kz
=R0
z H
z 0

where R is the recharge through the top of the aquifer and it


is assumed there is no leakage out of the bottom of the aquifer
(impervious boundary).

Putting all four terms back together yields the 2D flow
equation. For a confined aquifer this can be written as:

h
h h
=
T
+
T
+R
t x x x y y y

Finally, term #4 is:

(9.3.8)

which is a linear PDE. The recharge R is non-zero (and


positive) only if the confined aquifer has a leaky upper
boundary. For an unconfined aquifer the 2D flow equation is
given by:

Sy

h

h
h
=
K
h
+
K
h
+R
t x x x y y y

(9.3.9)

which is a nonlinear PDE in h due to the product terms


involving h. As mentioned in the context of the unsaturated
flow equation, nonlinear PDEs are generally more dicult to
solve. To make the equation linear we can use the following
identities:

h 1 2
=
(h );
x 2 x

h 1 2
=
(h )
y 2 y
290

along with a linearization (approximatino) of the storage term


about a nominal head h0:

Sy

S 2
h
h
Sy
= y
(h )
t
h0 2h0 t

Substituting these terms into the original Equation (9.3.9)


yields:

Sy h 2

1 h 2
1 h 2
=
K
+
K
+R
2h0 t
x x 2 x y y 2 y

d 2 R
0= 2 +
; subject to BCs
C2
dx

(9.3.12)

(9.3.10)

which is now a linear PDE with respect to the variable h2.



By expressing both equations in linear form, we can
write a unified governing linear 2D GW flow equation for a
homogeneous/isotropic aquifer:

2 2 R
C1
= 2+ 2+
t x
y C 2

direction), which is the same thing as saying that there is no


variation in head in the y-direction. In such cases, the lefthand-side term will be zero (steady-state) and the second
term on the right-hand-side, which involves derivatives with
respect to y, will also be zero. Therefore the 1D (steady-state)
flow equation will look like:

(9.3.11)

= h 2 2 [unconfined]; = h [confined]
C 1 = Sy (Kh0 ) [unconfined]; = S T [confined]
C 2 = K [unconfined]; = T [confined]
In many applications, especially those types of problems that
can be solved analytically, we will be dealing with steady-state
and/or 1D problems. In the case of steady-1D problems, we
can use the above unified equation as a starting point. A 1D
flow problem involves flow only in one direction (i.e. the x-

where it should be noted that since it is 1D and steady-state


the governing equation is no longer a PDE, but instead an
ODE. This equation can be integrated (twice) via separation
of variables, i.e.:

(x) =

R 2
x + a 0x + a 1
2C 2

(9.3.13)

The two integration constants (a0 and a1) can be determined


via application of the two BCs. Common BCs are fixed head
boundaries (which generally correspond to a lake/river
boundary where the water level is in steady-state) or a no-flux
boundary (where there is an impervious boundary like a
bedrock outcrop). To reiterate: the solution to the above
equation will be the piezometric surface (in this case h(x)),
which can then be used to determine fluxes, pore velocities,
travel times, etc.

291

E XAMPLE 9.3.1

E XAMPLE 9.3.1 ( CONTINUED )

A farmer has a row irrigation system as shown in


the cross-section figure below:

such that the rootzone remains unsaturated (i.e.


the water table remains below the rooting depth
throughout the entire length L). The soil has a
saturated hydraulic conductivity K. Assuming
steady-state conditions with no change in storage
in the unsaturated zone:
a) Write an expression relating the (steady-state)
irrigation, evapotranspiration, and recharge rates.
b) Determine the expression for the steady-state
water table profile beneath the crops as a
function of x (and in terms of R, K, H0, and L).

Figure 1: Cross sectional view of the unconfined aquifer for problem #1

c) Based on your answers to parts a) and b),


The soil is bounded on both sides by drainage
rite an expression relating the (steady-state) irrigation, evapotranspiration, and recharge rates. (2 pt.)
write the expression for the maximum irrigation
ditches which allow water to be removed from
etermine the expression for the steady-state water table profile beneath the crops as a function of x rate the farmer can apply to maintain
the
through lateral groundwater flow. The
in terms of R,
K, Hsoil
0, and L). (5 pts.)
unsaturated conditions in the rootzone.
inparts
thea)drainage
is maintained
atirrigation
a
sed on your water
answers to
and b), writeditches
the expression
for the maximum
rate the
er can apply depth
to maintain
conditions inrate
the rootzone.
(2 pt.) at the
H0unsaturated
. The irrigation
I is applied
a) The mass balance for the unsaturated zone control
-1
-1
ofLthe
soil
The
which
ssuming H0 =top
1.0 m,
= 6 m,
D =profile.
4 m, d = 1m,
K =water
3.0 cm day
, and is
E =not
8 mm day , what is this volume (i.e. below the land surface and above the
-1
ation rate in mm day ? (3 pts.)
evaporated via transpiration by the crops
water table) can be written as:
recharges (at a rate R) the unconfined aquifer
dS
blem #5: (15below
points (which
~ 45 min.)is bounded by a horizontal
= I E R
dt
re 2 (shown below)
is a cross-sectional
illustrating
the steady-state
positions and conditions
impermeable
layerdiagram
at depth
D). The
crops being
pond and river relative to a one-dimensional confined aquifer.
grown by the farmer have a rooting depth of d
In steady-state:
and have an average evapotranspiration rate E.
dS dt = I E R = 0
The farmer must decide the irrigation rate
292

E XAMPLE 9.3.1 ( CONTINUED )

E XAMPLE 9.3.1 ( CONTINUED )

b) The steady-state irrigated system yields a onedimensional unconfined aquifer which is governed by
(from Equation (9.3.11)):

c) The form of the derived head profile indicates that


there will be a peak in head at the mid-point of the
aquifer. This can be proven by taking the derivative of
the head profile and setting it to zero, and is true in this
case because the boundary conditions are the same at
each boundary. The maximum allowable irrigation is
such that the head is the requisite level shown in the
figure, i.e.:

d 2h 2
2R
=

K
dx 2
Integrating twice yields:

h 2(x) =

R 2
x + a 0x + a 1
K

which has two (unknown) integration constants. They


can be determined by applying the boundary conditions
(BCs) at each end of the aquifer:

h 2(x = 0) =

R 2
(0) + a 0 (0) + a1 = H 02
K

a1 = H 02

R
(L)2 + a 0 (L) + H 02 = H 02
K
RL
a0 =
K

h 2(x = L) =

R " L % RL " L %
2
2
2
h (x = L 2) = $ ' +
$ ' + H 0 = (D d)
K #2&
K #2&
RL2

+ H 02 = (D d)2
4K
(I E)L2

+ H 02 = (D d)2
4K
where solving for the irrigation rate yields:

I =E+

4K
(D d)2 H 02
2
L

so that the steady-state head profile is given by:

h 2(x) =

R 2 RL
x +
x + H 02
K
K

293

S ECTION 4

Groundwater Flow to
Pumping Wells

As mentioned above, groundwater is a primary source of
freshwater. To extract this supply from an aquifer generally
requires a pumping well (Figure 9.4) whereby the well taps
into the existing (natural) flow conditions. A pumping well
generally consists of a lined bore hole (i.e. with a pipe) that
has a screened end within the aquifer and a pump on the
other end of the pipe. Turning on the pump takes water out of
the aquifer and by mass balance induces flow toward the well
as will be described in more detail below. Note that a
monitoring well is quite dierent, as it does not have a pump
and is generally used solely for monitoring the piezometric
surface and/or constituents in the water (i.e. is passive rather
than an active part in the flow system). A pumping well may
be drilled into a near-surface unconfined aquifer or more
deeply into a confined aquifer. In this section we focus on the
hydraulics associated with flow toward pumping wells.

For mathematical tractability we will focus on a single
pumping well in an aquifer of infinite horizontal extent and for
conditions where the original piezometric surface (i.e. prior to
pumping) was horizontal. Extensions to multiple pumps and
more realistic cases can be built up from these single-well
solutions via superposition as discussed in the next section.
The basic schematic for such single well configurations is

F IGURE 9.4 Illustration of impact of pumping on the regional natural groundwater system.

shown for a confined and unconfined aquifer in Figures 9.5


and 9.6 respectively. The pumping rate is generally denoted
by Q and is given a positive value for extraction out of the
aquifer (or negative for pumping into the aquifer). The
original piezometric surface (h0) is shown with a dashed line.
Note for the confined aquifer it occurs above the top of the
aquifer due to the large pressure head term, while for the
unconfined aquifer it corresponds to the water table itself. The
spatial coordinate useful for well problems is the radius (r)
from the well. Additional parameters include the outer radius
of the well (rw) and the head within the well (hw). The water
294

F IGURE 9.5 Schematic of piezometric surface solution

around a single pumping well in a horizontally infinite confined aquifer (from Mays, 2005).

being extracted by the pump must be supplied by flow


through the aquifer. As is well known by now, flow in the
aquifer is driven by head gradients. For flow to be toward the
well, the head must be sloping downward toward the well.
This radially symmetric pattern is usually referred to as the
cone of depression. The key variable that is most often solved
for in pumping problems is the so-called drawdown:

s(r,t) = h0 h(r,t)

(9.4.1)

which is known if the head is known. A radius of influence is


often defined as the radius away from the well at which there
is no drawdown (i.e. at r = R, s = 0). An aquifer of infinite
horizontal extent is really just one that extends beyond the

F IGURE 9.6 Schematic of piezometric surface solution

around a single pumping well in a horizontally infinite unconfined aquifer (from Mays, 2005).

radius of influence of the pumping well. For a constant


pumping rate, the drawdown will initially be changing with
time (transient) such that the drawdown is a function of time
(and radius). Ultimately, after enough pumping occurs, the
amount being withdrawn will be exactly balanced by the flow
from the aquifer (steady-state) in which case the drawdown
will only be a function of radius.

In developing the flow equation for a single pumping well
in an infinitely extending aquifer we focus on the 2D flow
equation for a homogeneous/isotropic aquifer (Equation
(9.3.1)). This equation can be expressed in cylindrical
coordinates (centered at the well) as:
295

1 R
C1
=
r
+
t r r r C 2

(9.4.2)

To further simplify the problem we will assume no recharge


(R=0) in which case this becomes:

C1

1
=
r
t r r r

(9.4.3)

which is the governing equation for radial flow for a single


pumping well (without recharge). We will use this as a
starting point to develop solutions for cases with a constant
pumping rate.

We will start with the case of transient flow in a confined
aquifer, which corresponds to the expected behavior early on
after pumping begins. For the transient case (i.e. non-steadystate) the governing equation is:

S h 1 h
=
r
T t r r r

(9.4.4)

This PDE cannot be solved via separation of variables, but


has a solution which is referred to as the Theis solution
(solved in terms of drawdown):

s(r,t) = h0 h(r,t) =

Q
W (u)
4T

where W(u) is the well function which is given by:

(9.4.5)

e u
W (u) = du
u
u

(9.4.6)

where the dimensionless argument for the well function is:

Sr 2
u=
4Tt

(9.4.7)

TABLE 9.1. VALUES OF WELL FUNCTION W(u) AS A


FUNCTION OF THE NON-DIMENSIONAL PARAMETER u
u

0.219

0.049

0.013

3.8
10-3

1.1
10-3

3.6
10-1

1.2
10-4

3.8
10-5

1.2
10-5

10-1

1.82

1.22

0.91

0.7

0.56

0.45

0.37

0.31

0.26

10-2

4.04

3.35

2.96

2.68

2.47

2.3

2.15

2.03

1.92

10-3

6.33

5.64

5.23

4.95

4.73

4.54

4.39

4.26

4.14

10-4

8.63

7.94

7.53

7.25

7.02

6.84

6.69

6.55

6.44

10-5

10.94

10.24

9.84

9.55

9.33

9.14

8.99

8.86

8.74

10-6

13.24

12.55

12.14

11.85

11.63

11.45

11.29

11.16

11.04

10-7

15.54

14.85

14.44

14.15

13.93

13.75

13.6

13.46

13.34

10-8

17.84

17.15

16.74

16.46

16.23

16.05

15.9

15.76

15.65

10-9

20.15

19.45

19.05

18.76

18.54

18.35

18.2

18.07

17.95

10-10

22.45

21.76

21.35

21.06

20.84

20.66

20.5

20.37

20.25

10-11

24.75

24.06

23.65

23.36

23.14

22.96

22.81

22.67

22.55

10-12

27.05

26.36

25.96

25.67

25.44

25.26

25.11

24.97

24.86

10-13

29.36

28.66

28.26

27.97

27.75

27.56

27.41

27.28

27.16

10-14

31.66

30.97

30.56

30.27

30.05

29.87

29.71

29.58

29.46

10-15

33.96

33.27

32.86

32.58

32.35

32.17

32.02

31.88

31.76
296

Based on the aquifer properties (S and T), u can be computed


for any radius and time and used as input to the well
function. The well function is typically evaluated via its
tabulated solution (Table 9.1) or numerically (i.e. the MATLAB
command expint is the well function integral). Alternatively,
the well function can be expanded via a Taylor series
expansion, i.e.:
1 u2
W (u) = 0.5772 ln u + u
+ ...
(9.4.8)
2 2!
where the small higher-order terms can be dropped depending
on the size of u. In addition to the Theis solution providing
drawdown information, it can also be used to estimate aquifer
properties using measured drawdown from monitoring wells.

Strictly speaking, an analytical solution for transient flow
toward a pump in an unconfined aquifer is not available. This
has to do with the nonlinearity of the governing PDE.
However based on the linearized version, an approximate
solution can be obtained by using (K h0) in place of the
transmissivity (T) in the above solution. This is equivalent to
assuming the aquifer is quite thick so that the amount of
drawdown is relatively small compared to the original
piezometric surface (h0).

Next we focus on the analytical solutions for steady-state
drawdown. For a confined aquifer the governing equation is
given by:

S h 1 h
=
r
=0
T t r r r

(9.4.9)

d dh
r
=0
dr dr

Note that, because there is no time dependence, the equation


goes from a PDE (hard to solve) to an ODE that can be
solved easily via separation of variables. If the above equation
is integrated (twice) we get:

h(r) = c 0 ln(r) + c1

(9.4.10)

where c0 and c1 are integration constants that can be


determined from BCs. If available, head measurements at two
locations could be used to identify the integration constants.
Instead, generally one is related to the pumping rate and the
second to a measurement. The flow toward the pumping well
can be equated to the pumping rate via:

dh
Q = 2 rb K
dr

(9.4.11)

where the negative sign in front of the right-hand-side is


because flow is in the negative r-direction. The first term on
the right-hand-side represents the cross-sectional area of flow,
which is a cylinder at a radius r of height b, and the last term
in parentheses is the Darcy flux (which by definition is the
flow per unit cross-sectional area). Substituting the first
derivative of the general solution (Equation (9.4.10)) yields:
297

c
Q
Q
Q = 2 rb K 0 c 0 =
=
2 Kb 2T
r

(9.4.12)

which provides the first integration constant in terms of the


pumping rate and the transmissivity of the aquifer. Solving for
the second integration constant is generally done via the
specification of a measured head h1 at a known radius r1 (i.e.
from a monitoring well) where one can write:

h(r1 ) = h1 =

Q
Q
ln(r1 ) + c1 c1 = h1
ln(r1 )
2T
2T

(9.4.13)

Substituting for the two integration constants with these


conditions yields:

!r $
Q
h(r) =
ln # & + h
2T " r1 % 1

(9.4.14)

which is the solution for the piezometric head in the case of a


single pumping well in a confined (infinite) aquifer in steadystate conditions. For the special case where the specified
radius and head are the radius of influence (R) and the
corresponding (original) head (h0), yields:

h(r) =

r
Q
ln + h0 ;
2T R

rw r R

(9.4.15)

which can be rearranged to obtain an expression for the


steady-state drawdown:

r
Q
s(r) = h0 h(r) =
ln ;
2T R

rw r R

(9.4.16)

where it is noted that strictly this solution is valid starting at


the radius of the well extending out to the radius of influence.
Beyond determining drawdown fields, the above solutions can
be used to determine the aquifer transmissivity from two head
measurements at dierent locations (monitoring wells), i.e.:

T =

Q ln[r2 r 1 ]
2 h2 h1

(9.4.17)

Such an approach might first be used to characterize the


aquifer (i.e. determine its properties) which can then be used
to generally estimate drawdown fields throughout the aquifer.

Finally, the last case to examine is the steady-state flow
in an unconfined aquifer with a single pumping well. The
governing equation is:

S h 2 1 1 h 2
=
r
=0
Kh0 t
r r 2 r

(9.4.18)

d 1 dh 2
r
=0
dr 2 dr
which can be integrated (twice) and using the same procedure
for identifying integration constants as shown above (for the
confined aquifer case) yields:

298

r
Q
h (r) =
ln + h12 ;
K r1
2

rw r R

(9.4.19)

Because the solution is in terms of the square of the


piezometric head, there is not as clean of a solution for
drawdown as there is in the confined aquifer case, but it still
can be determined via:

s(r) = h0 h 2(r);

rw r R

(9.4.20)

Analogously to the confined aquifer, these solutions can also


be used to characterize the conductivity of the aquifer from
measurements at two monitoring wells:

K =

Q ln[r2 r1 ]
(h22 h12 )

(9.4.21)


So to summarize, we have defined the solutions to the
groundwater flow problem for three single-well infinite aquifer
cases: i) transient flow in a confined aquifer, ii) steady-state
flow in a confined aquifer, and iii) steady-state flow in an
unconfined aquifer. Additionally, while we do not have an
analytical solution for the fourth case of transient flow in an
unconfined aquifer, the Theis solution can be used if the
unconfined aquifer is relatively thick compared to the
drawdown. Finally, it should be noted that the above
development defined a positive pumping rate as one that
extracts flow from the aquifer. These cases yield a cone of
depression based on the given pumping rate. All of the above

solutions can also be used to model the case of a recharge


well, which is simply one that is used to pump water into the
aquifer. Mathematically, this just involves using a negative
value for the pumping (recharge) rate. The result is a cone of
build-up rather than one of depression as the water being
pumped into the aquifer initiates a head gradient away from
the well that drives flow laterally outward into the aquifer.
E XAMPLE 9.4.1
A single well in a (horizontally infinite) confined
aquifer begins pumping at a rate of 10 m3/day.
The aquifer has a storativity of 0.0001 and a
transmissivity of 50 m2/day. What is the
drawdown at a radius of 25 meters from the well
after 10 days?
This problem involves transient flow in a confined aquifer
and can therefore use the Theis solution. The nondimensional input argument for the well function is given
by:

(0.0001)(25 m)2
u=
= 0.00003
4(50 m 2 /day)(10 day)
From Table 9.1 the Well function at this value is: W(u)
= 9.84. The drawdown is:

(10 m 3 /day)
s(r,t) =
(9.84) = 0.16 m
2
4 (50 m /day)
299

E XAMPLE 9.4.2
A single well pumps from a horizontally infinite
confined aquifer at a rate of 5 m3/day until
steady-state is reached. The well is of radius 5 cm
and shows a drawdown of 1 m. At a monitoring
well located 100 meters from the pumping well
the drawdown is 0.1 m. Estimate the
transmissivity of the aquifer. Based on the
estimate, what is the expected steady-state
drawdown at a radius of 50 meters?
The two dierent head measurements can be used to
estimate the transmissivity:

(5 m 3 /day) ln[100 m/0.1 m]


T=
= 6.1 m 2 /day
2
1 m 0.1 m
Based on the estimated transmissivity and the measured
drawdown, the radius of influence can be determined:
" 2 (6.1 m 2 /day)(0.1 m) %
" 2Ts %
R! = r exp $
'
' = (100 m)exp $
(5 m 3 /day)
# Q &
#
&
= 215 m

The drawdown at 50 meters is then:

" 50 m %
(5 m 3 /day)
s(r) =
ln
$
' = 0.19 m
2 (6.1 m 2 /day) # 215 m &

300

S ECTION 5

Superposition of Groundwater Solutions



The linearity of the governing equation in groundwater
flow (most notably for confined aquifers) allows for the
application of the principle of superposition of solutions. What
this means is that if a complicated problem can be broken
down into a series of additive simpler ones (each with their
own solution), the solution to the complicated problem can be
obtained by adding up the solutions. This can be used for any
number of groundwater flow problems. Several types of
problems involving pumping wells are commonly solved via
this approach. Examples where this is useful include:
1. An infinite aquifer with multiple wells (instead of just a
single well)
2. Non-infinite aquifers with a no-flux boundary condition
3. Non-infinite aquifers with a fixed-head boundary condition
4. Combinations of the above cases
Additionally, cases where a regional natural groundwater flow
exists to which pumping wells are added can be solved via
superposition. Here we will focus on the above examples, with
the understanding that even more complicated problems can
be solved via superposition.


To start, the easiest example of superposition is the case
of a confined aquifer of infinite extent with multiple wells.
Conceptually, we know that a single well generates a
symmetric drawdown around the well. Multiple wells will have
multiple cones of depression that can form a complicated
head/drawdown field. The linearity of the single well solution
(with respect to h) allows for the determination of the total
drawdown as simply the summation of drawdown from each
well in the well field. (Note for an unconfined aquifer,
superposition does not strictly apply since the governing
equation is not linear in h). So the actual drawdown (s) is
given by:

s(x,y,t) = s1(x,y,t) + s2(x,y,t) + ... + sN (x,y,t)

(9.5.1)

where this assumes there are N wells. The x- and ycoordinates are used rather than r simply because each well
will have its own radial coordinate system. In practice, the
drawdown at a given location can be determined by finding
the distance from that point to each pumping well and using
that distance as the argument in the single-well solution (i.e.
in Equations (9.4.5) or (9.4.16)) to get each of the individual
drawdown predictions in Equation (9.5.1).

Figure 9.7 shows an example of a case with 2 wells with
pumping rates Q1 and Q2. In general, the pumping rates may
dier. Even for this relatively simple case, the composite
drawdown curve is complicated. For example, there is a local
maxima in between the two wells, with local minima at the
two wells. The local maxima corresponds to the location
where the head gradient is zero (implying no flow). It
301

E XAMPLE 9.5.1
It is required to de-water (i.e. lower the water
table) below a construction site that 80 m by 80
m in area as shown in plan below.

F IGURE 9.7 Drawdown solution for multiple wells via superposition (solid line) from single well solutions (dashed line).

represents the location (inflection point) separating regions of


flow from one well to another well. The specific locations of
these features depend on the relative pumping rates and
distances between wells. It is also important to keep in mind
that the resulting head is a 2D field (map) which could be
represented as a contour map of piezometric head or
drawdown. In the simple one-well case, the contours would be
concentric circles around the well. For multiple wells the
drawdown becomes much more complicated. Superposition
can be applied for any number of wells in a confined aquifer.
In the case of an unconfined aquifer, superposition can be
applied using confined aquifer solutions (as an approximation)
if the drawdown is relatively small compared to the
undisturbed head.

The bottom of the construction will be 1.5 m


below the initial water surface elevation of 100 m
(see cross-section figure shown below). Four
pumps are to be used in 0.5 m diameter wells at
the four corners of the site. The aquifer has
K=15 m/day and the wells each have a radius of
influence of 600 m. Assuming all four wells pump

302

E XAMPLE 9.5.1 ( CONTINUED )

E XAMPLE 9.5.1 ( CONTINUED )

at the same rate, determine the required steady


state pumping rate that has to be extracted from
each of the four wells to meet the desired
objective.

location will be in the center point of the construction at


a radius from one of the wells:

r = (40 m)2 + (40 m)2 = 56.7 m


The total drawdown from all four wells must be:

s(r) = s1(r) + s2(r) + s3(r) + s4 (r) = 1.5 m


where if the pumping wells are operating at the same
rate, the drawdown due to each well will be 1.5m/4
=0.375 m. Inverting the drawdown equation for a single
well yields the corresponding pumping rate:

Q =

2Ts(r)
2 (15 m/day)(100 m))(0.375 m)
=
ln(r / R")
ln(56.7 m / 600 m)

= 1498 m 3 /day

First, note that since the unconfined aquifer is relatively


thick compared to the drawdown, the solution will
behave like that of a confined aquifer with T=KH0 (i.e.
the governing equation is linear). This simplification is
useful because it allows for the application of
superposition of single-well solutions for a confined
aquifer. Next, we can note that the lowest drawdown


Building upon the example of multiple wells, we can
apply superposition to solve problems that involve non-infinite
aquifers. In this context, non-infinite means there is some
feature which impacts the flow to the well that is within its
radius of influence. There are generally two examples of this:
no-flux and fixed head BCs.

The first case we will discuss involves a no-flux boundary
within the radius of influence of the well. An example of such
a system is shown in the top panel of Figure 9.8.
303

Conceptually, for a given pumping rate, the aquifer must


supply that amount of water to preserve mass balance. In the
simple case of an infinite aquifer, the water comes
symmetrically from all directions with the result that
drawdown is a symmetric cone of depression around the well.
The primary consequence of the no-flux boundary is that flow
is cuto from a portion of the aquifer so that the water must
be supplied by other portions of the aquifer, leading to nonsymmetric and generally higher drawdown. While this
complicated problem could be solved numerically, the question
is can we use superposition to our advantage to solve the
problem using the addition of single-well solutions?

The use of superposition to solve such problems is often
referred to as the method of images. We can start with a
proposed solution as the summation of two single-well
solutions, i.e.:

s(r,t) = sreal (r,t) + simag (r ,t)

(9.5.2)

where the left-hand-side represents the true drawdown, the


first term on the right-hand-side is the predicted drawdown of
a single-well solution at the location of the real well and the
second term on the right-hand-side is a single-well solution for
an image (or imaginary) well located at a particularly chosen
location to satisfy the real conditions. The key condition that
needs to be satisfied in the no-flux boundary case is that the
head (or drawdown) gradient must be zero everywhere along
the no-flux boundary.

The location and pumping rate of the imaginary well can

F IGURE 9.8 Solution of flow to a pumping well with an im-

permeable boundary using the image well approach. The top


panel is the real system and the bottom panel is the conceptual
model using superposition with a pumping image well to ensure the no-flow boundary. The dashed lines in the bottom image are the single well solutions which when superposed yield
the real cone of depression (adapted from Mays, 2005).

be arbitrarily chosen to satisfy this condition. The simplest


way of doing this is to use symmetry to our advantage. This
304

involves putting an imaginary well at the same distance as the


real well on the other side of the boundary (hence it is an
image well with the boundary serving as the mirror), with
the same pumping rate as the true well. The result of this for
a no-flux boundary is shown in the bottom panel of Figure
9.8. Due to the symmetry of the choice made, adding the
drawdowns together ensures the proper BC at the location of
the impermeable boundary (i.e. dh/dr = 0). Moreover adding

F IGURE 9.9 Contour map of a single well solution in the presence of an impervious boundary. Note this is obviously incorrect as the gradient is non-zero at the boundary.

the two solutions over the entire domain provides the actual
cone of depression for this case (shown with the solid line).

Another example is shown in Figures 9.9 - 9.11 in terms
of the drawdown fields. Figure 9.9 illustrates that a single well
solution for the real system violates the no-flux BC. Figure
9.10 shows the contours for the drawdown fields predicted by
each of the (real and image) well drawdown solutions. Figure
9.11 shows the summation of the two drawdown fields, which

F IGURE 9.10 Contour map of the single-well solutions for


the real and image wells.

305

distance equal to the radius of influence of the image well.


Finally, note that the solutions used in Equation (9.5.2) could
be either transient or steady-state depending on the particular
problem.

F IGURE 9.11 Actual drawdown obtained via the summation


(superposition) of the real and image well solutions shown in
Figure 9.10.

indeed has a zero head gradient everywhere perpendicular to


the impermeable boundary. Additionally it is clear that the
overall drawdown is significantly increased, most notably in
the region between the real well and the impervious boundary.
As one moves away from the impervious boundary on the far
side of the real pumping well, the drawdown approaches zero.
Strictly this occurs at a distance equal to the radius of
influence. The impact of the boundary will disappear at a


The other type of non-infinite aquifer condition that
often must be dealt with is that of a stream boundary. For
simplicity, the stream is often modeled as a fixed-head BC
(Figure 9.12). The primary consequences of the fixed-head
boundary is that the boundary acts as a water source that
would not exist in an infinite aquifer, so that less water is
required to be supplied by the rest of the aquifer leading to
non-symmetric and generally less drawdown. Superposition
can again be used to solve this problem. We again choose to
add an image well at the same distance as the real well, but
on the other side of the fixed-head boundary. However, rather
than making the image well a pumping well (to represent the
water loss and no-flux boundary condition provided by the
impervious boundary) we make it a recharge well (to represent
the extra source of water provided by the stream boundary).
By symmetry, since the real and image wells are the same
distance and the same magnitude of pumping rate, the
amount of drawdown predicted by the single-well solution for
the real well is exactly balanced by the same amount of
build-up as a result of recharge. Therefore, everywhere along
the fixed-head boundary the drawdown is zero. This is shown
in the bottom panel of Figure 9.12. Moreover we see that
generally there is less overall drawdown everywhere in the
aquifer due to the extra source provided by the stream.

306


Another example analogous to that shown in Figures
9.9-9.11 for the no-flux BC is shown in Figure 9.13 for the
fixed-head BC. The setup is exactly the same as before except
that the image well involves recharge so that there is a sign
change in the image well drawdown contours. In this case the
drawdown from the image well is negative (build-up). By
summing up the two single-well solutions, we get the result
shown in Figure 9.13. As can be seen, the drawdown is

F IGURE 9.12 Solution of flow to a pumping well with a fixed-

head boundary using the image well approach. The top panel is
the real system and the bottom panel is the conceptual model
using superposition with a recharge image well to mimic the
no-flow boundary The dashed lines in the bottom image are the
single well solutions (one discharge and one recharge) which
when superposed yield the real cone of depression (adapted from
Mays, 2005).

F IGURE 9.13 Actual drawdown contours (plan view) ob-

tained via the superposition of a real pumping well solution


and an image recharge well solution.

307

eectively zero everywhere on the boundary and the overall


drawdown is less than for a single well in an infinite aquifer.
E XAMPLE 9.4.2
A pumping well for a confined aquifer is located
100 meters away from a valley wall as shown
below. The well is pumping 0.06 m3s-1. The
storativity of the aquifer is 10-4 and the
transmissivity is 950 m2day-1. Find the total
piezometric surface (head) drawdown at the
midpoint between the pumping well and the
valley wall after 2 days from the start of
pumping.

E XAMPLE 9.4.2
The drawdown due to the pumping well if there were no
boundary would be (using the Theis solution):

Q
W (u1 );
4 T
Sr12
(10 4 )(50 m)2
u1 =
=
= 3.3 10 5
2
4Tt 4(950 m /day)(2 day)
W (u1 ) = 9.76

s1 = s(r = 50 m, t = 2 days) =

s1 =

86400 s
)
1d
9.76 = 4.24 m
4 (950 m 2 /day)

(0.06 m 3 s 1

A mirror image well on the other side of the boundary


would yield a drawdown of:

Q
W (u2 );
4 T
Sr12
(10 4 )(150 m)2
u1 =
=
= 2.96 10 4
2
4Tt 4(950 m /day)(2 day)
W (u1 ) = 7.55

s2 = s(r = 150 m, t = 2 days) =

Assuming the radius of influence of the well is greater


than 100 meters, then the bedrock outcrop will need to
be handled as a no-flux boundary condition using an
image well. Further, the pumping is in the transient
phase (2 days after pumping).

s2 =

86400 s
)
1d
7.55 = 3.28 m
2
4 (950 m /day)

(0.06 m 3 s 1

So that the total drawdown (via superposition) is:

s = s1 + s2 = 4.24 + 3.28 m = 7.5 m

308


These examples show some of the power of the
superposition principle to solve relatively complex problems in
a relatively easy way. Even these examples/concepts of
superposition could be used to build up solutions for more
complicated problems. For example, if a pumping well is in a
river valley with a horizontal head distribution prior to
pumping, the aquifer may have a stream BC condition on one
side and a impervious boundary on the other. Depending on
the radius of influence of the well (i.e. given its pumping rate)
both boundaries may impact the flow patterns. In such a case
one may use two image wells and have an additive solution
with three terms instead of two. Or even more complicated
yet, the original head distribution (prior to pumping) may not
be horizontal, but itself the solution of a 1D flow problem
between the two BCs with recharge. In this case the solution
could be the summation of the original piezometric head
distribution combined with the drawdown from the real and
imaginary wells. Another example would be the case with
multiple real wells with a no-flux or fixed-head BC (or both).
In this case each real well would have a corresponding image
well. In these types of problems the hardest part is breaking
down the real problem into its constituent parts (each of
which has an analytical solution), with the easy part being
determining the solution via superposition of the individual
solutions.

As groundwater flow problems become more complex,
one may instead need to solve the problem numerically. In
such a case, the aquifer is discretized into small dierential
units and the GW flow equations are solved on the finite

dierence or finite element grid. Figure 9.14 shows an example


corresponding to a square island (i.e. with fixed head
boundary on all four sides) that is subject to a non-uniform
steady-state recharge (i.e. more recharge in the northeast
corner of the island). Numerical solution of the 2D flow
equation yields the steady-state piezometric head on the grid,
showing a characteristic recharge mound with flow implicitly
down the head gradient to the island boundaries. Figure 9.15
shows a solution for the same recharge pattern, but with five
pumping wells extracting water from the aquifer. As expected,
the solution is a superposition of the solution in Figure 9.14

F IGURE 9.14 Steady-state piezometric head (in meters) on a


square island with non-uniform recharge.

309

F IGURE 9.15 Steady-state piezometric head (in meters) for


the same recharge as in Figure 9.14, but with five pumping
wells.

with drawdown surrounding each well location. This


illustrates the more complicated flow patterns that may exist
in real groundwater flow problems.

310

S ECTION 6

Conceptual Questions
1. Name the two types of groundwater aquifers. Describe how
they dier.
2. What type of aquifer generally occurs closest to the land
surface (i.e. is shallowest in the subsurface)?
3. In an unconfined aquifer, what physical boundary does the
piezometric surface correspond to?
4. In an unconfined aquifer, is the piezometric surface above
or below the upper boundary of the aquifer? Explain.

9. Suppose you have a no-flux boundary condition in a


groundwater problem. What do you know about the head
at the no-flux boundary?
10. What is the dierence (in meaning) between a transient
and steady-state pumping problem?
11. What is the general shape of the piezometric surface
surrounding a single pumping well?
12. What are the underlying assumption/s used in the
development of the three single-well solutions derived in
this chapter?
13. Define the meaning of radius of influence in the context of
a pumping well.

5. Name the storage parameters used to characterize an


unconfined and confined aquifer.

14. Describe what superposition is and how it is useful in


groundwater problems. For which type of aquifer is
superposition strictly applicable or not applicable?

6. Why is the 2D groundwater flow equation usually sucient


for groundwater problems? What two dimensions are used
in this equation?

15. Describe the purpose of using an image well to solve a


groundwater flow problem (i.e. why is it used?).

7. In any groundwater flow problem, what variable are you be


solving for?
8. In the 2D groundwater flow equation, what is the name of
flux entering the top of the aquifer? Is it typically higher for
an unconfined or confined aquifer?

311

a) On the map, sketch the flow direction of water starting


from point (3 km, 3 km) to the edge of the island.

S ECTION 7

Sample Problems

b) Where is the Darcy flux the maximum and what is its


approximate value? Explain your reasoning for choosing where
the flux is largest.

Problem 9.1. The head field for a confined aquifer


underneath a 5 km x 5 km square island (with a homogeneous
and isotropic hydraulic conductivity of 1 m/day) is plotted in
the contour map below. The datum is such that the head on
all sides of the boundary of the island is 0 m.

Problem 9.2. The picture below shows the terrain and


steady-state groundwater cross-section (marked by the thick
line on the inset map) of Long Island, New York. The island is
composed of glacial sedimentary deposits that have a known
saturated hydraulic conductivity of 60 meters/day. If the
length of the cross-section is 30 kilometers and the
groundwater mound peak is 6 meters above sea level, what is
the average groundwater recharge over Long Island (in mm/
year)?

Hint: As with any such boundary value groundwater problem,


your solution should start with a clear identification of what
the governing equation for this groundwater flow problem is
followed by a general solution of the governing equation. The
specifics of the problem can then be used to find the
particular solution.
312

Problem 9.3. The figure shown below is a cross-sectional


diagram illustrating the steady-state positions and conditions
of a pond and river relative to a confined aquifer.

A chemical company accidentally dumps a conservative


chemical substance into the pond. Assuming that the chemical
instantly mixes with the pond water, estimate how long (in
days) it would take for the contaminant to get to the river.
You can reasonably assume that the chemical moves with the
water (i.e., it doesnt adsorb to the soil). The aquifer
properties are: porosity=0.426, b = 18 m, and T = 0.85 m2 s-1.
The height of the pond surface above the datum (hp) is 537 m,
the height of the river above the datum (hr) is 531 m, and the
length of the aquifer (l) is 7 km.
Problem 9.4. A common method of cleaning up
contaminated aquifers is via a pump-and-treat approach
where a pumping well is installed to extract the polluted
water where it can then be treated above ground and either
recharged back into the aquifer or disposed of o-site. There is
a proposal to pump-and-treat water from a partially polluted
confined aquifer (see figure shown below). A key feasibility
question is the time it will take to pump out the polluted
plume. In order to first characterize the aquifer properties, a

well is pumped from the b = 88 m thick aquifer at a rate of


0.11 m3 s-1 until steady-state is reached. In two observation
wells located 15 m and 65 m away from the pumping well, the
dierence in water elevation at the two observation wells is
measured to be 2.3 m. The porosity is estimated to be 0.492
using the soil cores obtained during the well construction.
a) What is the hydraulic conductivity of the aquifer material
in m day-1?
b) What is the expression for pore-velocity (v) toward the
well?
c) Find the expression for travel-time t(r) from radius r to the
pumping well. Hint: For non-constant pore-velocity v:

t(r) =

dr
v

d) Evaluate the travel-time for r = 50 and 500 m. Comment


on whether it is practical to pump-and-treat large pollutant
313

plumes (greater than 500 m from pumping well) in this


aquifer.

groundwater near the river. For the vegetation to survive it is


determined that the maximum steady-state drawdown in the
riparian zone must be less than or equal to 0.1 meters. The
municipal agency would like to pump 65 m3/day from the
well. The radius of influence under these pumping conditions
is 5000 m. Will this pumping rate satisfy or violate the
steady-state riparian zone drawdown constraint? Justify your
answer. The transmissivity of the aquifer is 100 m2/day, and
L1 = 2500 m, L2 = 1500 m and L3 = 1000 m.

Problem 9.5. A pumping well is being designed for


installation in a confined aquifer in a valley that is bounded
by a river on one side and a bedrock (i.e. impermeable)
outcrop on the other (shown in plan below). A riparian zone
borders the river and contains habitat that is deemed vital to
a particular migratory bird species. The vegetation draws its
rootzone soil moisture via capillary forces from the shallow

!
314

Chapter 10

Runoff and
Streamflow

S ECTION 1

Learning Objectives
By the time you finish this chapter you should be able to:

10. Apply the unit hydrograph approach to predict


stormflow for a particular storm given the UH for that
duration storm
11. Describe the key dierences between using a physicallybased distributed modeling approach and the UH method

1. Describe the basic mechanisms for overland runo and


how they work
2. Describe the basic mechanisms for subsurface runo and
how they work
3. Define and describe the dierences between distributed
hydrologic modeling and lumped hydrologic modeling
4. Dierentiate between stormflow and baseflow
5. Understand the concept of baseflow separation and be
able to apply the basic empirical approaches for baseflow
separation
6. Describe the primary assumptions of the unit hydrograph
(UH) method
7. Construct a D-hour unit hydrograph from data
8. Understand and describe the concept of eective (or
excess) rainfall and how it relates to stormflow
9. Construct a D*=nD -hour unit hydrograph from a D-hour
unit hydrograph
316

S ECTION 2

Basic Runo and Streamflow Definitions



Traditionally, runo and streamflow are a major focus of
hydrology due to their relation to both floods and water
supply. Simply put, runo is a lateral flux (either at the
surface or in the subsurface that transports water toward the
stream channel network. High (rare) runo events lead to the
potential for flooding, which can be mitigated via flood
protection measures (levees, detention basins, culverts, etc.).
Streamflow can be captured behind a dam in a reservoir to
store water that would otherwise be lost downstream.

Runo into streams (where it becomes streamflow) is an
integration of several upstream processes in a watershed. It
can be thought of as an integrated response of the watershed
to storm and/or snowmelt events. Specific mechanisms for
generating runo will be described in the next section. To
start, it is worth noting that topography generally plays a
significant role, not just in the amount of runo, but in
directing it downstream. An example of a DEM is shown in
Figure 10.1, where one can readily see mountain peaks and
valley troughs across the landscape. In Chapter 1 we
described how one can delineate a watershed by choosing an
outlet point (on a stream) and tracing out all points upstream
of that location that would ultimately flow to it. Implicit in

F IGURE 10.1 Sample DEM over a region that can be used to


derive stream channel network.

this is that water flows down the steepest topographic


gradient (slope).

A stream (or river) network is simply an expression of
this fact, where a stream is generally just defined as points
in a basin where water will generally collect via the routing of
flow over the topography or through the groundwater system.
We can define this more precisely by computing a quantity
like the contributing area of each point in the basin. The
contributing area (Ac) is simply the area (or equivalently the
number of pixels in a DEM) upstream of a given location, i.e.
all the points that will flow into that pixel. Some limiting
cases are the watershed outlet, which by definition has the
entire watershed area as its contributing area, and the points
317

on the watershed ridge, which have no pixels upstream of


them (i.e. Ac = 0). A stream pixel can therefore be defined as
any pixel that has a contributing area above some threshold
(e.g. 1% of the basin area). One can compute and plot the Ac
(actually plotted as log(Ac)) as shown in Figure 10.2. What
clearly stands out is low values (dark blue) for the ridges and
higher values (yellow/red) for the valley areas. If one only
accepts stream pixels as those with contributing areas greater
than 1% of the total basin area, the stream network can be
identified simultaneously with the watershed delineation
(Figure 10.3). Given this construct, one can identify two kinds

F IGURE 10.3 An example watershed delineated from DEM

shown in Figure 10.1. The stream channel is defined using


those regions with a flow accumulation above a specified threshold (i.e. Ac = 1%).

F IGURE 10.2 Flow accumulation patterns across terrain derived from DEM in Figure 10.1. Areas of high flow accumulation (red) correspond to stream channels while areas of low
flow accumulation area (dark blue) represent watershed
boundaries.

of pixels within a basin: hillslope pixels and stream network


pixels. Water falling on hillslope pixels will infiltrate and/or
runo to the nearest stream channel pixel and then water will
flow down the stream channel network toward the outlet.
Movie 10.1 illustrates this eect. Processes that occur on the
hillslopes generate runo and the streamflow at any point in
the stream channel network is an integration of upstream
runo processes. Runo amounts will vary by location in the
watershed and will have variations in travel time from any one
location to the outlet. The amount of flow crossing the outlet
at any point in time is generally referred to as a hydrograph
(described more in Section 4).
318

M OVIE 10.1 Animation showing varying travel time of water


from different locations in a basin (from COMET program).

319

S ECTION 3

Runo Generation
Mechanisms

Runo is generated as a result of a water flux being
applied to the surface. This flux can either be in the form of a
rainstorm or snowmelt (Figure 10.4). Storm events can vary in
terms of intensity and duration, both of which should impact
the runo. For conceptual purposes, one can think of

F IGURE 10.4 Conceptual picture of runoff resulting from either rainfall or snowmelt.

snowmelt as a long-duration storm event. The subject of this


section is on the various mechanisms that are responsible for
generating runo. Note that they should depend on the
characteristics of the storm and the characteristics of the
basin (soils, slopes, etc.). The motivating questions for this
section are: What are the mechanisms that generate runo?
When, why, and where do they occur in a basin? How do they
contribute to the overall flow in a stream?

Figure 10.5 shows a conceptual picture of hydrologic
fluxes in a basin with particular emphasis on runo
mechanisms contributing streamflow. Runo can be classified
into two main categories: i) surface (overland) runo and ii)

F IGURE 10.5 Components of various runoff and related processes contributing to streamflow.

320

subsurface runo. Surface runo is either generated by


infiltration excess runo or saturation excess runo. The
former was already discussed to some extent in Chapter 7
since it is directly connected to the infiltration capacity of the
surface. Subsurface runo is generated by either interflow or
baseflow and is simply related to groundwater flow processes
discussed in Chapter 9. Each of these mechanisms are
elaborated on below.

Infiltration excess runo, often referred to as Hortonian
runo after Horton (1933), is related to the infiltration
process at the surface. As described in Chapter 7, depending
on the antecedent soil moisture condition and soil properties,
a soil has a given infiltration capacity. Depending on the
storm intensity (and duration), the soil may reach a state in
which the precipitation intensity exceeds the infiltration

capacity of the soil. Movie 10.2 shows a conceptual


representation of this phenomenon. The amount of water that
is reaching the surface cannot be fully infiltrated into the soil.
As such, there will generally be ponding (saturation) at the
surface (even though the subsurface soil is not saturated),
which, given a sloped surface, will result in runo. As shown
in Chapter 7, it is important to remember that this particular
process is a threshold process. When the precipitation rate is
less than the infiltration capacity (i.e. before ponding) or less
than the minimum infiltration capacity of the soil (i.e.
saturated hydraulic conductivity), all water will infiltrate,
meaning there will be no runo. Movies 10.3 and 10.4
illustrate this by showing a hillslope experiencing two dierent
precipitation intensities, one of which results in no runo and
the other results in infiltration excess runo.

As discussed in Chapter 7, the infiltration rate for the
general case (where ponding occurs at some point during the
storm) can be written as:

P, t 0 t t p

f (t) =
fc (t tc ), t p t tr

M OVIE 10.2 Animation of infiltration excess runoff mechanism at a given location on a hillslope.

(10.3.1)

where after ponding a shifted infiltration capacity model is


used via the time compression approximation. By definition,
the infiltration excess runo rate is simply given by: Qie(t) =
P - f(t). So the infiltration excess runo is zero up until
ponding occurs. Figure 10.6 illustrates the cumulative
infiltration and cumulative infiltration excess runo that
occurs over the course of the storm. The cumulative depth of
321

M OVIE 10.3 Animation of infiltration excess runoff under

light rainfall conditions relative to soil properties (from COMET


program).

F IGURE 10.6 Cumulative infiltration and infiltration excess


runoff for a general case.

infiltration excess runo can be computed via integration:


tr

t0

M OVIE 10.4 Animation of infiltration excess runoff under

ie

(t) =

tr

tr

[P f (t)]dt = [P f (t t )]dt

t0

tp

(10.3.2)

where the integral represents the blue shaded region in Figure


10.6. It should be noted that this runo generation mechanism
occurs only during the storm and is often in localized areas of
low conductivity soils (i.e. P >> Ks) where ponding is
generated at the surface. If ponding does not occur, then no
infiltration excess runo can occur.

heavy rainfall conditions relative to soil properties (from


COMET program).

322

E XAMPLE 10.3.1
Compute the cumulative infiltration excess runo
for the case described in Example 7.7.1.
The cumulative infiltration excess runo can be
computed via the integral in Equation (10.3.2).
Alternatively if the cumulative infiltration is already
computed (as in Example 7.7.1), then the runo is
simply the dierence between the cumulative rainfall and
the cumulative infiltration. So for the Philip solution the
estimated runo would be:
tr

t0

ie

(t) =

tr

P dt F(t ) = (30 mm/h)(8 h) 177 mm

t0

= 63 mm

of the precipitation initially infiltrates, but due to a shallow


water table, the pores quickly fill up until the surface is
saturated from below. At that point, any additional
precipitation is falling on saturated soil (i.e. no ability to
infiltrate) and therefore completely runs o. Movie 10.6 shows
another illustration of the process. The key point is that for
this mechanism to work, the initial water table must be
relatively shallow. This condition generally occurs near the
stream, so contributing areas that generate saturation excess
runo are most often localized around the stream network.
Moreover, these areas are often referred to as variable
contributing areas in reference to the fact that, based on the
groundwater dynamics, they grow and contract both
seasonally and during and after the course of a storm. An
example of this is shown schematically in Figure 10.7. In this
example, before a large storm there is little saturated area

Similarly for the Green-Ampt model the estimated runo


would be:
tr

t0

ie

(t) =

tr

P dt F(t ) = (30 mm/h)(8 h) 187 mm

t0

= 53 mm

The second overland flow generation mechanism is
referred to as saturation excess runo. It is sometimes referred
to as Dunne runo in reference to the work that first
identified the mechanism (Dunne, 1975). This mechanism
occurs when the groundwater table (due to recharge from
above) saturates the soil from below. Movie 10.5 shows a
conceptual representation of this phenomenon. In this case all

M OVIE 10.5 Animation of saturation excess runoff mechanism.

323

rainfall events, the saturation excess runo mechanism will


generally dominate over infiltration excess runo. The fact
that Dunne runo is the primary contributor was a new
discovery, where prior to those studies it was thought that
Hortonian (infiltration excess) runo was the primary
overland runo generating mechanism. It turns out that only
in areas of low conductivity soils and/or as a result of very
high intensity storms will Hortonian runo tend to be a
significant contributor to total overland flow. In areas of
snowmelt, either mechanism can generate runo depending on
the melt rate, conductivity, and saturation of the soil
underlying the snowpack. In both mechanisms, due to high
surface flow velocities, and/or proximity to channel network

M OVIE 10.6 Animation of saturation excess as a result of a


rising groundwater table.

that would be expected to contribute to saturation excess


runo. After a large storm the variable contributing area is
significantly larger. The areas often surround the stream
channel network where the groundwater table intersects the
surface. During the drydown the groundwater table recedes
and the variable area shrinks. During the wet season when
contributing area is larger, more saturation excess runo can
be generated.

The relative importance of infiltration excess and
saturation excess runo depends on several factors including
soils and climate. In humid areas with relatively low intensity

F IGURE 10.7 Illustration of variable contributing areas growing and contracting during and after a storm. These areas are
the locations expected to generate saturation excess runoff
(adapted from Dingman, 2008).

324

(in the case of Dunne runo), overland runo tends to reach


the channel network relatively quickly, contributing to
stormflow during or shortly after the storm event.

In addition to overland flow, two runo generation
mechanisms are associated with subsurface processes. These
are illustrated conceptually in Figures 10.8 and 10.9.
Interflow, or perched stormflow, is the lateral movement of
water through the unsaturated zone. Since the hydraulic
gradients in the unsaturated zone are generally largest in the
vertical direction (i.e. driving vertical flow), lateral flow is

F IGURE 10.9 Interflow generation as the result of perched


groundwater (adapted from Beven, 2004).

often the result of temporary perched water (i.e. a perched


water table) on a low conductivity soil lens. Another interflow
mechanism, that is much more dicult to characterize, is that
occurring via so-called macropores. These network of pores
are generally the result of root growth or biological factors
that eectively create channels in the soil that can eciently
transport water (Figure 10.10). Generally speaking, interflow
is often a relatively small component of total runo.
Additionally, due to relatively low subsurface velocities,
interflow may reach the stream network after the storm ends.

F IGURE 10.8 Illustration of subsurface runoff processes: interflow and baseflow.


The final subsurface runo mechanism is generally
referred to as baseflow. It is simply lateral groundwater flow
into the stream channel network (Figures 10.8 and 10.9). The
flow can be from either unconfined or confined aquifers. In
Chapter 9, streams were treated as boundary conditions in the
groundwater flow problem. As seen in those examples the flux
is generally into the stream channels, which is exactly what
baseflow is. In the steady-state groundwater flow problems
325

that can be solved analytically, the stream is a static feature


in the problem. In reality of course the two systems are
coupled via mass balance. Baseflow from the groundwater
system supplies the river network which in general may cause
an increase in river stage (height).

Finally, with respect to storms it is important to keep in
mind the time scales of subsurface processes (Table 1.3). First,
the percolation and recharge resulting from a given storm can
take a significant amount of time (porous media velocities are

generally low). Second, depending on the size of the basin, the


low groundwater flow velocities imply that the recharged
water often reaches the stream significantly after the storm.
This is precisely why perennial streams exist, where
streamflow occurs even in the dry season when storms are not
occurring. During these low flow conditions, the streamflow is
completely made up of baseflow. The various components of
the streamflow hydrograph will be discussed in more detail in
the next section.

F IGURE 10.10 Interflow as the result of flow through macropores.

326

S ECTION 4

Streamflow Hydrographs

The mechanisms described in the previous section
ultimately lead to runo that contributes to streamflow. One
could envision measuring streamflow at a given point in the
stream network. If measured, the amount of flow crossing that
point (i.e. cubic meters per second) would simply be the result
of all upstream flow processes (both runo to the closest
downstream channel location and flow down the stream to the
measurement point) as a function of time. The streamflow
crossing a given point over time is generally referred to as a
hydrograph, where one can conceptualize the total flow (Q)
as:

Q(t) = Qie (t) + Qse (t) + Qi (t) + Qb (t)

F IGURE 10.11 Typical hydrograph during and after a storm.

Baseflow corresponds to groundwater flow, while other components together are often referred to as stormflow. The fractional components depend on the storm and basin characteristics (adapted from Mays, 2005).

(10.4.1)

where the four terms on the right-hand-side are respectively


amounts of runo related to the infiltration excess, saturation
excess, interflow, and baseflow runo mechanisms.

Figure 10.11 shows a schematic of a streamflow
hydrograph (at a basin outlet) resulting from a rain storm.
The figure shows the typical characteristics of the hydrograph
curve as well as hypothetical contributions from dierent
runo mechanisms. Prior to the storm, and before any runo
reaches the outlet, the hydrograph is composed entirely of
baseflow from the basins groundwater system. The rising limb
of the hydrograph generally occurs shortly after the storm

begins and results primarily from overland flow mechanisms


(infiltration excess and/or saturation excess, i.e. Qie(t)+Qse(t).
Typically the overland runo contribution peaks early in the
hydrograph response. After the peak, the falling limb of the
hydrograph will ultimately recede (generally much more
slowly than the rising limb) back to baseflow levels.
Secondarily, the interflow and baseflow tend to respond more
slowly to the storm, often peaking after the overland flow has
moved past the outlet.

How much each mechanism contributes to overall
streamflow, as well as the hydrograph shape, depends on both
327

dynamic factors related to the individual storm characteristics


(intensity, location, etc.) as well as static factors related to
basin characteristics (topography, soil types, vegetation types,
etc.).

Figure 10.12 illustrates how the hydrograph can vary as
a result of the spatial distribution of a storm over a basin.
Storm A is isolated over an extreme location of the basin and
Storm B is uniformly distributed over the basin. For Storm A,
the runo is generated far from the outlet and therefore takes
a significant amount of time to reach peak flow. For Storm B
the peak occurs much earlier since areas nearer the outlet
generate runo. Additionally, the total volume of runo (i.e.
integration of the hydrograph) is larger for Storm B since the
contributing area is so much larger.

Figure 10.13 illustrates how hydrograph shape can vary
as a result of the temporal distribution of precipitation during
the course of a storm. The left panel shows a case where the

F IGURE 10.12 Impact of rainfall distribution on stormflow


hydrograph (adapted from Mays, 2005).

F IGURE 10.13 Impact of rainfall timing on stormflow hydrograph (adapted from Mays, 2005).

highest intensity rainfall occurs late in the storm and the right
panel shows a case where the highest intensity rainfall occurs
earlier in the storm. In the former case the rising limb is less
steep with a later peak.

Figure 10.14 illustrates how particular basin
characteristics can impact the hypothetical storm hydrograph.
The top panels show the response for basins of diering slope;
steeper basins will have hydrographs that respond and recede
more quickly. The bottom panels show the response for basins
of diering roughness; basins with higher roughness will
respond and recede less quickly. Another factor that is not
shown is the level of imperviousness, which is especially
relevant in urbanized basins. Basins with high levels of
urbanization will have more infiltration excess runo and
therefore the hydrograph will respond and generally recede
more quickly. Additionally, the available storage in the basin
can have an impact; basins with little storage will respond and
recede more quickly.
328

F IGURE 10.14 Impact of basin slope and roughness on stormflow hydrograph (adapted from Mays, 2005).

M OVIE 10.7 Animation showing travel time from extreme



Some of these factors are highlighted in the form of
animations shown in Movies 10.7-10.9. The travel time of
water falling on a basin, as one measure of basin response, is
generally going to be impacted to first-order by key static
characteristics like basin shape, size, and slope.

Lastly, in discussing the streamflow hydrograph it is
important to define the stormflow hydrograph. For flood
forecasting and flood mitigation design, engineers are most
often interested in the quick response or stormflow portion of
the hydrograph. Nominally, this is the portion of the
hydrograph that is in direct response to the storm. Since
baseflow is often relatively small and responds with significant
lag to a storm, the stormflow generally refers to the
infiltration excess, saturation excess, and interflow
components of the hydrograph (i.e. stormflow neglects

portions of the basin for varying basin shapes.

M OVIE 10.8 Animation showing travel time from extreme


portions of the basin for varying basin sizes.

329

M OVIE 10.9 Animation showing travel time from extreme


portions of the basin for varying basin slopes.

baseflow). In practice, since the interflow component is also


often small, stormflow and overland runo are sometimes used
synonymously. The next two sections describe some examples
of modeling hydrograph response of a basin, where implicitly
the goal is the modeling of the stormflow hydrograph.

330

S ECTION 5

Rainfall-Runo Modeling:
Unit Hydrograph

The modeling of streamflow for a basin is not only of
implicit hydrologic interest, but has deep roots in the
engineering practice of flood forecasting and the need for
design flood estimation. In either of these applications, the
peak flow and the timing of the flow are key input parameters
to the design/forecast. Models that can be used for predicting
hydrographs span a large range of complexity. A detailed
treatment of modeling runo is given in Beven (2012).

In this section we will focus on a classical method used
in hydrology called the unit hydrograph (UH) method, which
is specifically designed to predict the stormflow hydrograph
for a given storm. Some of the underlying choices in this
method include: treating the basin as a lumped response unit,
using historical (archived) data to develop the model response
function, and making some simplifying assumptions including
using a systems (black-box) response approach. The primary
benefit of these choices is that the model is very
computationally ecient. In the next section we will focus on
a more physically-based approach that is becoming more
commonly applied in research and/or practice.

The UH approach is an empirically-based method to
predict stormflow response for a given storm. Conceptually,

the method treats the basin as a black-box where for a given


input (i.e. precipitation) one can develop a response function
that yields an output (stormflow) hydrograph. In particular,
the method uses excess (or eective) precipitation as the
input, which corresponds to the cumulative amount of
stormflow runo from a given storm. In other words, the
excess precipitation is the total precipitation minus the depth
that infiltrates (and therefore does not contribute to runo. So
one way to look at the method is that if one knows the total
amount of runo, the UH method transforms that cumulative
amount into its temporal distribution (including rising limb,
peak flow, recession, etc.).

The basic idea of the UH approach is to derive the unit
response function for an excess precipitation storm of D-hour
duration. In this context the unit response refers to the
response to 1 cm (or 1 inch) of excess precipitation. The
length of the storm that generated excess precipitation
(runo) will have a first-order impact on the shape of the
hydrograph, which is why all UHs are tied to a particular Dhour event. The key assumption of the UH method is that the
basin responds linearly to excess precipitation inputs. If the
basin does respond linearly, then superposition can be used (a
key aspect of the UH method). The other assumptions in the
UH method are generally implicit in the linear response
assumption. As described in the previous section, the shape of
the hydrograph will change if the storm varies in space across
the basin (Figure 10.12). For a useful response function to be
derived, its shape must be representative and reproducible.
Hence an implicit assumption in the UH method is that the
331

precipitation occurs uniformly over the basin. If the storm


occurs uniformly over the basin, then the shape of a measured
streamflow hydrograph should be representative of the static
characteristics of the basin (i.e. soils, slope, roughness, etc.)
and therefore have useful predictive utility. Note that the
shape is thus inherently tied to a basins characteristics and
therefore dierent basins should have dierent unit
hydrographs. Similarly, if the excess precipitation is uniformly
distributed over the D-hour event, the linearity assumption
will be most valid. In addition to the above assumptions, one
would expect that antecedent soil moisture and groundwater
table conditions before the storm are important (i.e. will
determine how much infiltration vs. runo occurs and how it
is distributed across the basin). Hence it is conceivable that
dierent unit hydrographs could be developed for dierent
antecedent conditions. Real-world departures from these
simplifying assumptions are relatively common and will
undoubtedly introduce errors into UH stormflow predictions.

We will first discuss the recipe used to construct a Dhour UH. Then we will discuss the various ways in which they
are applied. The key inputs needed for construction of a UH
are shown schematically in Figure 10.15 and consist of a
streamflow hydrograph, the basin area, the basin-averaged
excess rainfall, and the duration (D) of the excess rainfall.
Note that the excess rainfall and duration of excess rainfall
are in general less than the actual rainfall and actual storm
duration. You should be able to convince yourself of this by
considering the special case where only infiltration excess
runo occurs, while keeping in mind that in this special case

F IGURE 10.15 Inputs necessary for constructing a unit hydrograph.

the excess rainfall is simply the cumulative infiltration excess


runo. As shown in Figure 10.6, runo would only begin after
ponding, so that in this case the duration of excess rainfall
would be D = tr - tp and the excess rainfall would be the
integral shown in Equation (10.3.2). The more general case is
more complicated as the excess rainfall also includes
saturation excess rainfall (and perhaps interflow) as well, but
the concept is similar in that some of the precipitation will
infiltrate and not contribute to stormflow. The steps of a UH
construction are outlined sequentially below.
332

function of time so that we could represent the total


hydrograph as: Qt , where t corresponds to discrete time
periods (e.g., every 30 minutes or every hour).
Step 2: The second step is to remove the baseflow at each
time step (Bt) from the streamflow hydrograph. This is done
because the UH method makes no attempt to predict baseflow
(only stormflow). If desired, the baseflow can be added back
to the predicted stormflow to get the total predicted
hydrograph. The conceptual picture is illustrated in Figure

F IGURE 10.16 Illustrative example of a streamflow hydrograph that can be used to derive a unit hydrograph.

Step 1: The construction of a UH starts with a historical


streamflow hydrograph for the basin (e.g. Figure 10.16). Note
this implies that there are streamflow measurements for the
basin of interest. If no measurements exist, then synthetic
UHs must be used (Mays, 2005). Here we focus only on the
case where measurements are available. The hydrograph is the
result of runo mechanisms described in Section 3 that are
driven by a storm of a given duration and intensity. From
historical records of precipitation, one may know the upper
bound of the excess rainfall duration, but it will need to be
identified specifically as described below. From a data
perspective, the hydrograph can generally be tabulated as a

F IGURE 10.17 Illustrative example showing the stormflow

(quick-response runoff) and baseflow components of the


streamflow hydrograph. The duration of excess precipitation is
a key input that defines a D-hour response.

333

10.17 which shows the duration of excess precipitation (i.e. D


hours) and the same hydrograph shown in Figure 10.16. The
baseflow is shown in green. Note that in general we do not
actually know what the baseflow is. Figure 10.18 shows a
similar picture, but with the time to peak and the time of
concentration labeled on the graph. The time of concentration
represents the time it would take for water from the most
distant part of the basin to reach the outlet and corresponds
to the inflection point in Figure 10.18. By definition it
represents the time at which the quick-response runo
components go to zero leaving only the baseflow component of

the hydrograph. Several empirical methods exist for


attempting to estimate the baseflow. The simplest method
assumes that there is no baseflow response to the storm event,
which is equivalent to assuming Bt = Qt0 (i.e. the baseflow is
equivalent to the streamflow prior to the storm event). Under
this simplified model, the baseflow is simply a constant in
time. Slightly more sophisticated empirical methods attempt
to estimate the inflection point to account for increases in
baseflow during the response. One example (Mays, 2005)
assumes the hydrograph trend before the storm response
(which may be a recession curve from the previous storm)
continues until the time to peak and that the inflection point
is Ndays after that time, where:

N days = 1.21A0.2 ;

A basin area (km 2 )

(10.5.1)

Many other empirical functions for the time of concentration


are given in Mays (2005). Note that these are all empirical
relationships and therefore generally not dimensionally
consistent. Using any of these or other empirical techniques,
one can estimate a discrete time series for baseflow: Bt.

F IGURE 10.18 Illustrative example showing the time to peak


and time of concentration for the streamflow hydrograph.

Step 3: Determine the cumulative stormflow (excess


precipitation) by removing the baseflow and integrating under
the stormflow hydrograph. The removal of baseflow yields the
stormflow hydrograph (Figure 10.19), i.e.: Qt - Bt. If the
baseflow was estimated accurately, then the integral under
this curve should exactly match the cumulative quick-response
runo (stormflow). When transformed to depth, this is by
definition equal to the excess precipitation, which therefore
can be calculated by:
334

F IGURE 10.19 Plot of the stormflow hydrograph with baseflow removed from the original streamflow hydrograph.

Pe =

1
(Q B)dt
A

F IGURE 10.20 Conceptual picture of the process of converting total volume of excess runoff to the equivalent amount of
uniform runoff depth over the basin.

(10.5.2)

where A is the basin area in proper units. The integral yields


a volume and the area transforms it to a depth (Figure 10.20).
This will yield the total depth of the excess precipitation
(runo). In practice, the curve is not known analytically, but
at discrete times, so the above integral is estimated via a
numerical approximation (i.e. the rectangular or trapezoidal
methods).
Step 4: Normalize the storm flow hydrograph by the eective
rainfall to get the D-hour unit hydrograph (Figure 10.21). In

other words, each ordinate of the storm flow hydrograph can


simply be divided by Pe, so that the unit hydrograph is given
by:

ht =

Qt Bt
Pe

(10.5.3)

where the units of the UH are flow units per unit depth of
stormflow, i.e. m3/s/cm. The UH simply represents the
response function that would occur if 1 cm of eective rainfall
occurred over D hours.
335

stormflow runo. In this particular example, only three hours


in the middle of the storm exceeded the threshold. Hence in
this example, we would have an excess rainfall duration of D
= 3 hours. Therefore the UH shown in Figure 10.21 is the 3hour UH. Note that this particular method assumes that the
loss function is constant in time, which may be an
oversimplification (i.e. in the case of infiltration excess runo).
Additionally, it shows that the excess precipitation is not
spread uniformly in time over the 3 hours, which is a violation
of the underlying UH assumptions. Nevertheless, this is
representative of a typical situation where the UH is applied.

F IGURE 10.21 The stormflow (quick-response) hydrograph

is converted to the unit hydrograph by dividing by the equivalent depth of runoff. Note the units of the unit hydrograph are
volumetric flow per unit stormflow depth (e.g. m3/s/cm).

Step 5: Determine the duration of excess precipitation. While


the above UH construction is implicitly in response to a Dhour excess rainfall event, the actual duration must be
determined prior to application. The rainfall hyetograph (time
series) gives some indication of the duration of the storm, but
the duration of excess rainfall will generally be less than the
storm duration due to infiltration. Figure 10.22 shows an
example of one method to determine D. In this example the
storm was a 12-hour storm. To determine D, a precipitation
depth is found such that the amount above the line equals the
excess precipitation found in Step 3. Implicitly, the
precipitation below this cuto line infiltrated into the soil and
went into storage, while the amount above the line went to

F IGURE 10.22 Example of one method for determining the

duration of excess precipitation based on the known amount of


excess precipitation.

336

E XAMPLE 10.5.1 ( CONTINUED )

E XAMPLE 10.5.1
The hydrograph shown below was measured at
the outlet of a 1 km2 watershed in response to a
storm event that corresponded to 2 hours of
excess precipitation. Construct the 2-hour UH
from this hydrograph. For simplicity assume the
flow at the start of the hydrograph is
representative of the baseflow.
TIME
(HOURS)

FLOW
(m3/s)

0.05

0.10

0.15

0.13

12

0.10

15

0.08

18

0.05

Given the historical hydrograph, the first step is baseflow


removal. In this case the flow at the beginning of the
storm is assumed representative of baseflow and
subtracted to generate the stormflow hydrograph:

TIME
(HOURS)

STORMFLOW
(m3/s)

0.0

0.05

0.10

0.08

12

0.05

15

0.03

18

0.0

Next, the excess precipitation can be estimated by


integrating under the stormflow hydrograph and
normalizing by the basin area (Equation (10.5.2)). Using
the trapezoidal rule, the integral can be estimated as:

Pe = 0.5[(0 + 0.05) + (0.05 + 0.10) + (0.10 + 0.08) +


(0.08 + 0.05) + (0.05 + 0.03) + (0.03 + 0.0)m 3 /s]
(3h)(3600s / h) / (1 km 2 / (1000 m)2 )
= 0.0067 m = 0.67 cm
The original hydrograph can then be normalized by the
excess precipitation to get the UH (Equation (10.5.3)):
337

E XAMPLE 10.5.1 ( CONTINUED )


TIME
(HOURS)

UH (m3/s/cm)

0.0

0.074

0.148

0.118

12

0.074

15

0.044

18

0.0

Note that this is a 2-hour UH because the excess


precipitation event is of a 2-hour duration and not
because of anything about the time base of the
hydrograph itself (which is a function of basin size and
other characteristics). The shape of the UH is the same
as the original hydrograph with the peak occurring at 6
hours after the start of the storm event. The UH can be
used to estimate peak flow for storms associated with
other excess precipitation amounts (and other storm
durations) via superposition as described below.


The end result of the UH construction process is the
determination of the unit response of the basin (i.e. to 1 cm of
excess rainfall) over a D-hour duration. Once constructed, the
UH can be used in applications, including the prediction of
response to other D-hour events of varying intensity as well as
response to storms of dierent excess precipitation intensity
and duration. This is done via superposition based on the
assumed linear response of the system. Three examples of
applications of the UH method are presented here including:
prediction of stormflow from a dierent D-hour storm,
construction of an (nD)-hour UH from a D-hour UH (where n
is an integer multiple), and prediction of response from a more
complicated excess precipitation event.

The simplest application of the UH method involves
using a D-hour UH to predict the response of a dierent Dhour event. For example, suppose for design or flood
prediction purposes you want to predict the resulting
hydrograph for a 6-hour eective rainfall event with Pe of
excess precipitation. The entire basis of the UH approach is
that the system behaves linearly so that the shape should be
invariant, but the magnitude can be scaled. So in this case, if
you have a 6-hour UH, you would simply scale it by Pe:

Qt = Pe ht

(10.5.4)

By construct, the integral under the scaled stormflow


hydrograph will exactly equal Pe. Examples of this for storms
with 0.5 and 2.0 units of eective precipitation (e.g. cm) are
shown in Figure 10.23. Note that implicit in this is that the
starting and ending times of stormflow do not change, only
338

cm that is lagged by 1 hour. This is just a statement of


superposition. One can use this to construct an (nD)-hour
UH. Suppose n = 2 and D = 1 hour. By starting a UH at
time zero and lagging a second UH by one hour (blue curves
in Figure 10.24) and summing the two together would yield a
hydrograph with 2 cm of stormflow. This summed hydrograph
can then simply be rescaled (in this case divided by 2) to yield
the 2-hour UH. This operation is the same as taking the
average of the two lagged UHs and results in the red curve
shown in Figure 10.24, which is the 2-hour UH (constructed

F IGURE 10.23 Example of predicted stormflow hydrograph


for different amounts of excess precipitation based on scaling
the unit hydrograph.

the magnitudes change. To get a full streamflow hydrograph


for the scaled event, one could add back an estimate of the
baseflow.

Another application of the UH method is the
construction of an (nD)-hour UH from a D-hour UH. This
approach again uses linearity to its advantage. Suppose for
example a 2-hour excess precipitation event occurred with 1
cm of rainfall in each hour. Linearity suggests that the
response to this should be identical to the summed response of
a 1-hour event of 1 cm followed by a second 1-hour event of 1

F IGURE 10.24 Illustrative example of constructing a 2-hour


unit hydrograph from two 1-hour unit hydrographs lagged by
one hour and re-scaled to correspond to a unit amount of
stormflow.

339

from two 1-hour UHs). This could then be used to predict the
response to any 2-hour event.

The final application discussed here is the case where
there is an excess precipitation event of longer duration than
the available UH and of varying excess precipitation
throughout the event. An example is that shown in Figure
10.25, which shows an excess precipitation time series over a
24-hour period. For the purposes of illustration, we will
assume we have already constructed a 6-hour UH for this
basin. One approach would be to construct the 24-hour UH
from four lagged 6-hour UHs as described above and then

F IGURE 10.25 Example of a 24-hour excess precipitation

time series. The non-uniformity of the time series suggests the


use of superposition of multiple UHs (e.g. four 6-hour UHs)
might perform better than the use of a single 24-hour UH.

scale the 24-hour UH by the amount of excess precipitation.


Another approach is to break the event into smaller pieces,
apply UHs to each, and then sum them up to get the full
response. One reason to do the latter over the former is if
there are considerable variations in excess precipitation over
the course of the event. In the special case that the excess
precipitation were constant over the 24 hours, then the two
approaches would be expected to be identical. In this example
we could propose to break the excess precipitation hyetograph
into four 6-hour pieces. Again, linearity lets us treat each as
independent events where the responses to each can
ultimately be summed up (often referred to as convolution).
By breaking the event up, each sub-period comes closer to the
assumption of constant excess precipitation over its duration.
For each sub-period, the response will be the 6-hour UH
starting at the beginning of the event scaled by the excess
precipitation taking place over that period. In Figure 10.26,
the four periods are color-coded as green, blue, red, and
yellow and the respective responses (i.e. scaled 6-hour UHs) to
each are shown in the bottom panel. Using superposition, the
stormflow hydrograph resulting from all of the events over the
24-hour period can be obtained by adding up the individual
responses as shown schematically with the black line in Figure
10.27. This is illustrative of the power of superposition where
a single UH is used to build up a model to a more
complicated input. The net result is a prediction that is
considerably more complicated than that from a single event
response and includes multiple local maxima. Whether the
prediction is accurate depends on the validity of the
underlying UH assumptions.
340

F IGURE 10.26 Illustration of superposition of UHs. The ex-

cess precipitation time series is conceptualized as four different


6-hour storms. The linearity assumption allows for the predicted response of each 6-hour event. The bottom panel
shows the stormflow hydrographs (i.e. scaled UHs) corresponding to each 6-hour event. The color of each hydrograph corresponds to the same color excess precipitation event.

F IGURE 10.27 The bottom panel shows the stormflow hy-

drographs (i.e. scaled UHs) corresponding to each 6-hour


event. The color of each hydrograph corresponds to the same
color excess precipitation event. The bottom panel shows the
predicted total response (in black) which is simply the summation (superposition) of each of the individual (colored) response hydrographs shown in Figure 10.26.

341

E XAMPLE 10.5.2

E XAMPLE 10.5.2 ( CONTINUED )

A small watershed is located in an area that is


frequently hit by intense thunderstorms. You
decide to investigate the peak flows for two
possible extreme thunderstorm events (which are
the largest expected amount to fall in each
duration storm): 1) a one-hour storm with a
cumulative eective rainfall of 24 cm and 2) a
three-hour storm with a cumulative eective
rainfall of 45 cm. The 1-hr unit hydrograph (UH)
constructed for the watershed is shown in the
table below.

b) Based on further investigation it is determined


that the basin is composed of soils with a large
saturated hydraulic conductivity and that the
unit hydrograph was constructed from a typical
intensity storm. What runo generation
mechanism can you reasonably hypothesize was
primarily responsible for the basin response seen
in the unit hydrograph?

TIME
(HR)

UH FLOW
(m3/s/cm)

10

a) Estimate the peak stormflow corresponding to


each storm (based on a unit hydrograph
analysis). Which of the two storms yields a
higher peak runo?

c) For the extreme events analyzed above, should


we expect the same runo generation mechanism
to be responsible for the peak stormflow? How, if
at all, should we expect this to impact the
accuracy of the UH-based prediction?
a) The first storm has duration of 1 hour, which allows
us to directly use the UH ordinates while scaling them
for the actual eective precipitation; this also means that
the peak time occurs at the same time for which the UH
peaks:

Qpeak = Q(t = 1 h) = h(t = 1 h) Pe


= (10 m 3 /s/cm)(24 cm) = 240 m 3 /s
For the 3-hour storm, the three 1-hour UHs can be used
with the first UH starting at t = 0 h, the second UH
starting at t = 1 h and the third UH starting at t = 2 h.

342

E XAMPLE 10.5.2 ( CONTINUED )


The one-hour lag for each UH is possible if superposition
is invoked whereby the 3-hour event can be
conceptualized as three separate one hour events. Each
UH can be scaled by 15 cm (i.e. 45/3 cm) so that the
total cumulative eective precipitation is equal to the
desired total. At each hour the scaled runo values can
be summed via superposition. For this particular case
the peak flow is 225 m3/s at both t = 2 and 3 h.
b) If the soil is highly conductive and the UH was
derived from a typical storm (i.e. precipitation intensity
was not too high) it is highly likely that runo is
generated via the saturation excess mechanism (i.e.
Dunne runo) rather than the infiltration excess
mechanism (i.e. Hortonian runo) which generally
requires P >> Ks.
c) Since the events analyzed are extreme events it is
likely the precipitation intensity is much larger than a
typical storm. This increases the likelihood that there
will be infiltration excess runo (instead of saturation
excess runo). If the runo mechanism generating runo
for the storm of interest is a dierent mechanism than
that underlying the original UH, it is possible that the
UH will not be of appropriate shape and therefore the
predictions could be erroneous. A physically-based model
may be more appropriate in this case.


While in the UH construction the excess precipitation is
determined directly from the historical hydrograph,
applications to other events require some estimate of Pe. In
reality it is determined by the runo generation mechanisms
occurring in the basin, which depend on the precipitation
intensity, soil type, and antecedent conditions (i.e. initial soil
moisture). In practice it is often estimated empirically. One
commonly applied method is the so-called SCS method
developed originally by the U.S. Soil Conservation Service
(now known as the Natural Resources Conservation Service
(NRCS)). A schematic of the assumed processes is shown in
Figure 10.28, where Ia is the initial abstraction (i.e. where all
water infiltrates), Fa is the continuing abstraction (potentially
decaying infiltration rate), P is the total precipitation and Pe
is the excess precipitation. Based on empirical evidence from
many small experimental watersheds, the excess precipitation
can be estimated by:

(P 0.2S)2
Pe =
P + 0.8S

(10.5.5)

where S is the potential maximum retention storage in the


watershed. The storage is typically estimated by:

S=

1000
10
CN

(10.5.6)

where CN is the so-called SCS curve number which is


dimensionless and varies between 0 and 100. A value of CN =
100 corresponds to an impervious surface (i.e. no storage and
343

1. All of the above can be theoretically applied to snowmelt


runo modeling just as easily as a rain storm with the
appropriate modifications.
2. The linearity assumption makes for a very computationally
ecient rainfall-runo model that can be easily
implemented on a spreadsheet and modified for many cases
as illustrated above.
3. By the same token, there is no guarantee that the linearity
assumption is a good one. Any model is only as good as its
underlying assumptions, so the user must be careful in
application of the UH method to minimize errors that could
be introduced as a result of improper assumptions.

F IGURE 10.28 Illustration of SCS empirical method for esti-

mating excess precipitation from precipitation data (adapted from


Mays, 2005).

4. The key tradeo to computational eciency in the UH


method, is a complete lack of knowledge of the underlying
physical mechanisms responsible for the runo and
sensitivity to errors as a result of unaccounted for changes
in the basin. For a more physical treatment of the rainfallruno processes, physically-based models like those
described in the next section may be required.

Pe = P). The curve numbers are generally a function of land


use, soil types, and antecedent moisture conditions. Tabulated
values can be found in Mays (2005) or other sources. Given
the estimate of excess runo for a given storm, the UH
method can then be applied using a derived UH of the
appropriate duration.

A few final comments on the UH method:

344

S ECTION 6

Rainfall-Runo Modeling:
Physically-based Models

The unit hydrograph approach to rainfall-runo
modeling described in the previous section has been and
continues to be widely used. If one needs only basin outlet
runo predictions and the underlying assumptions of the UH
method are reasonably valid for the basin of interest, then
UH-based predictions may be sucient (and often are, e.g. for
design purposes). The empirical nature of the approach
however includes limitations.

The primary alternative to empirical modeling is
modeling using physically-based approaches. This simply
means that the processes within the basin are modeled using
the physics that have been the primary basis of this book. As
has been made clear in earlier chapters, the primary drivers of
watershed processes are precipitation and net radiation. Given
these inputs, a set of processes ensue that include infiltration,
evaporation, unsaturated zone moisture redistribution,
recharge, groundwater flow, and runo. Each of these is
governed by physical processes that can be expressed in terms
of models. If tied together into an integrated unit, the model
becomes a physically-based hydrologic watershed model. One
should be plainly aware of the tradeos between models.
Physically-based models have the potential for a more robust
modeling framework (that may include modeling of interior

basin processes not just basin outlet streamflow), but


generally require significantly more input data, both
meteorological data and characterization of the basin (soils,
vegetation, terrain, etc.). At least part of the reason (beyond
simplicity) that empirical models have been the traditional
approach to rainfall-runo modeling is that hydrology used to
be a very data-limited enterprise, where a given basin may
have had a single stream gauge and perhaps a few rain
gauges. For such a limited data environment, simpler models
make sense. However, with the advent of remote sensing that
has been discussed in earlier chapters, many key hydrologic
inputs (precipitation, radiation, topography, etc.) are available
in a much more comprehensive way. Such new data streams
raise the possibility of implementing more complicated
models.

The first step in physically-based modeling is generally a
decision about to what extent processes are represented in
space. Figure 10.29 illustrates three distinct ways in which a
watershed can be discretized in space for modeling and
analysis. Choosing a method of discretization has tradeo
implications related to explicitly modeling processes, accuracy
due to both the degree of underlying spatial variability and
available data, and computational demand. Many existing
models are available that span various levels of discretization
as well as how processes are modeled.

The simplest approach is generally referred to as using a
lumped modeling approach (left panel in Figure 10.29). In a
lumped approach the entire basin is lumped or grouped
together in one unit. The UH method is an example of a
345

F IGURE 10.29 Illustration of various ways of representing a


basin in model: lumped (left panel), semi-lumped or semidistributed (middle panel), or distributed (right panel) (from
COMET Program).

lumped (albeit empirically-based) approach. In lumped


models, the physical processes may be represented (i.e.
evaporation, infiltration, runo) with basin-scale mass balance
and flux equations (in varying ways), but no attempt is made
to account for spatial distributions within the basin. Inputs to
lumped models would include mean areal estimates of
precipitation, radiation, etc. and the outputs would be basin
runo as well as basin-averaged estimates of evaporation and
storage in the various modeled reservoirs. An example of a
widely used lumped model is the Sacramento Soil Moisture

F IGURE 10.30 Model structure for the conceptual SACSMA


model that is used by the National Weather Service (NWS)
River Forecasting Centers.

Accounting (SACSMA; NWS, 2002) model, which is


represented schematically in Figure 10.30. Sometimes such
models are referred to as conceptual models in that the
physical processes may be represented in a more
parameterized way. The primary goal of such lumped models
is the outlet hydrograph.

An intermediate approach that attempts to discretize the
basin to represent some of the spatial variability, is referred to
346

as a semi-lumped (or semi-distributed) approach (middle


panel of Figure 10.29). In this approach the basin is
discretized into several sub-units (either sub-watersheds or
areas between elevation contours). In such a framework, each
sub-unit is treated as a lumped unit. Mean areal inputs are
required for each sub-unit and mean areal outputs are

F IGURE 10.31 Semi-distributed schematic representation of


basin as used in the USACE HEC-HMS model.

provided for each sub-unit (including runo). To get the


basin-outlet runo, the runo from each sub-unit must be
routed to the outlet. Techniques for routing hydrographs are
described in more detail in the next section. An example of a
widely-used semi-lumped model is the U.S. Army Corps of
Engineers (USACE) HEC-HMS model (USACE, 2010) which
is shown schematically in Figure 10.31. In this model the full
watershed is discretized into a series of sub-basins (lumped
units) that are connected via a series of river reaches.

Finally, the last approach generally used involves socalled distributed hydrologic models (right panel in Figure
10.29). In this approach the basin is discretized into relatively
small units (typically either on a regular grid or via a
triangular irregular network [TIN] yielding Thiessen polygontype units). The attractiveness of distributed models is that
increasingly available spatially-distributed datasets (i.e.
NEXRAD RADAR, satellite-based radiation, topography,
etc.) can be used to model processes at a high resolution. The
desire to model within-basin processes or how changes to a
basin (i.e. urbanization) impact the hydrology can be
addressed with such a model in ways that lumped models
cannot. The drawback is the increased need for input data
and specification of model parameters. An example of a
raster-based (i.e. grid-based) distributed model is the MIKESHE model (Graham and Butts, 2005), which is shown
schematically in Figure 10.32. Such grid-based models are
consistent with many of the relevant datasets (e.g.
topography) that are also raster-based. Other models take
advantage of the potential for increased computational
347

F IGURE 10.33 TIN-based distributed schematic representaF IGURE 10.32 Distributed schematic representation of basin

tion of basin used in the tRIBS model (from vivoni.asu.edu/


tribs/tinmodeldiagram.jpg).

as used in the MIKE-SHE model (from mikebydhi.com/Products/


WaterResources/MIKESHE.aspx).

eciency (i.e. fewer cells) if one uses a TIN to generate a


computational mesh. In such models, the individual modeling
units are the triangular elements of the TIN or the Thiessen
(or Voronoi) polygons formed by the TIN. An example of one
such model is the so-called tRIBS model (Ivanov et al, 2004),
which is represented schematically in Figure 10.33.

For the rest of this section, the focus is on distributed
hydrologic modeling, where the simpler modeling approaches

can be thought of as a special case of representing processes in


a single or handful of lumped units. Figure 10.34 shows a
single cell (or pixel) in a distributed model. Implicit in any
discretization is that the processes at the pixel-scale can be
treated as a homogeneous unit. The spatially-distributed
inputs include all relevant meteorological data (precipitation,
radiation and reference-level air temperature, humidity, wind
speed, etc.). At the surface, surface mass and energy balances
can be solved, which partition the incoming precipitation into:
infiltration, evapotranspiration, interception, and unsaturated
348

zone storage and net radiation into: surface sensible and latent
heat fluxes, ground heat flux and soil energy storage.
Subsurface processes that are modeled may include: one
dimensional flow in the unsaturated zone (including
redistribution and groundwater recharge), and groundwater
mass balance. Lateral fluxes modeled at the pixel-scale include
overland flow (infiltration excess and saturation excess) and
subsurface runo (interflow and baseflow). These runo
generation mechanisms yield an outflow hydrograph for each

F IGURE 10.35 Illustration of saturation excess runoff generation predicted by the tRIBS distributed model (from Enrique Vivoni personal communication).

F IGURE 10.34 Illustration of surface and subsurface compo-

nents at a pixel within the tRIBS distributed model (from Enrique


Vivoni personal communication).

pixel. One example of output from a distributed model that


cannot be obtained in a lumped approach is shown in Figure
10.35, which illustrates the fraction of saturation excess runo
occurrence from each pixel in a basin.

In applying a distributed model, not only are the
hydrologic point-scale processes at each unit modeled
349

explicitly, but hydrologic connectivity between cells becomes


an important component. Figure 10.36 shows an example of

F IGURE 10.37 Illustration of how distributed models provide


hydrographs at both internal and outlet nodes.

F IGURE 10.36 Illustration of stream channel (flow patterns)


in a distributed model.

the cell connectivity for the example basin shown in Figure


10.29. Such connectivity can be derived from topographic
information as described in Section 2. The routing scheme
used by the model takes hydrographs generated at a given
interior point and routes them downstream (Figure 10.37).
The explicit routing takes into account increase in peak flow
due to downstream accumulation as well as translation and
attenuation in time. The end result is the outflow hydrograph
(which is also output by a lumped model) along with
knowledge of all of the interior processes (and their spatial/
temporal variability) that led to the streamflow hydrograph.
350

S ECTION 7

Streamflow Routing: Unsteady Flow



Streamflow routing is simply the solution for the
hydrograph at a given point (or at all points) along a stream
channel as a function of time. The most general form of
solution is often termed hydraulic routing and is simply the
solution of unsteady flow in a channel. The mean flow in a
stream channel is downstream and hence the flow can
generally be treated as one-dimensional with the stream
channel as the single spatial coordinate (i.e. the x-direction),
as represented schematically in Figure 10.38.

One can derive the one-dimensional unsteady flow
equation starting with the continuity (mass balance) equation
originally shown in Equation (1.5.2), which for mass
conservation yields:

0=

d
dV + CS V dA
dt CV

(10.7.1)

where the control volume can be a dierential element of


length dx in the channel. The mass inflow rate is given by:

inlet

V dA = (Q + qdx)

(10.7.2)

F IGURE 10.38 Relevant variables defining flow in an open


stream channel (adapted from Mays, 2005).

351

where Q(x,t) is the flow in the stream channel and q is the


lateral inflow (per unit length of channel) and represents, e.g.
the overland runo into the stream. The mass outflow rate is
given by:
Q

V
dA
=

(Q
+
dx)
(10.7.3)

x
outlet
where the second term on the right-hand-side represents the
increment in flow due to lateral inflow and/or any storage
changes. Finally, the rate of change of mass stored in the
control volume is given by:

dV
=
( Adx)

dt CV
t

(10.7.4)

where A dx is the volume of the channel element. Putting all


three terms back together yields:

( Adx)
Q
(Q + qdx) + (Q +
dx) = 0
t
x

(10.7.5)

Assuming fluid density is constant yields the 1D unsteady


continuity equation:

A Q
+
q = 0
t
x

(10.7.6)

which is subject to initial and boundary conditions in flow


and cross-sectional channel area. Note however that this

equation has two unknowns: A and Q. Therefore to solve it we


need to apply another independent constraint via the
momentum equation.

The momentum equation can also be derived from
Reynolds Transport Theorem (Equation (1.5.1)) and written
as:

F =

d
V dV + CS V V dA
dt CV

(10.7.7)

where the left-hand-side represents the net force, the first


term on the right-hand-side represents momentum storage and
the last term represents the divergence of momentum. The net
force can be written as:

F = F

+ Ff + Fe + Fp

(10.7.8)

where Fg is the gravitational force, Ff is the friction force, Fe


is the force associated with expansion/contraction of the
channel, and Fp is the unbalanced pressure force. The sum of
all forces can be expressed as:

F = gAS 0dx gAS f dx gASedx gA

y
dx
x

(10.7.9)

where S0 is the channel bed slope, Sf is the friction slope (head


loss per unit length of channel), Se is the head loss due to
expansion/contraction and the unbalanced pressure force is
due to dierences in hydrostatic pressure on either side of the

352

channel element as a result of dierences in water elevation.


The momentum inflow rate is given by:

V V dA = (VQ + v qdx);

inlet

(10.7.10)

momentum coefficient

where the first term on the right-hand-side is the momentum


entering the upstream face and the second is the momentum
entering laterally (i.e. due to lateral influx q) where vx is the xcomponent of the velocity of the lateral influx. The
momentum coecient accounts for the nonuniform
distribution of velocity in the channel cross-section (Mays,
2005). The momentum flux at the outlet of the elemental
volume is given by:
(VQ)
V

V
dA
=

VQ
+
dx)
(10.7.11)

x
outlet
where the last term represents a change in momentum flux
(either as a result of the lateral influx of momentum or due to
a storage change). Finally, the rate of change of momentum
storage is given by:

d
Q
V

dV
=

dx

dt CV
t

(10.7.12)

Putting all of the terms back together (and simplifying)


yields:

# y
&
Q (Q 2 A)
+
+ gA % S 0 + S f + Se ( qvx = 0
t
x
$ x
'

It should be noted that the flow depth y is directly tied to the


cross-sectional area (A) and therefore is not an independent
(new) variable. So Equations (10.7.6) and (10.7.13) provide
two equations for the two unknowns: Q and A (or depth y).
Together they are often referred to as the Saint-Venant
equations which are a coupled set of PDEs that describe the
1D unsteady flow in a channel (subject to initial/boundary
conditions and lateral inflow) and are often solved numerically
(Mays, 2005).

Many of the commonly used equations for flow in open
channels/streams can be derived as special cases of the SaintVenant equations. Simplifications may include steady-state, no
lateral inflow, uniform flow, etc. Because of their wide use
they are quickly covered here.
1. Steady-flow with no lateral inflow: For steady flow with no
lateral inflow (i.e. q = 0), the continuity equation becomes:

A
Q
+
q =0 = 0
t =0 x
which simply states that Q = constant from one cross-section
to the next (i.e. Q1 = Q2 or V1A1 = V2A2), which is the
commonly applied steady-state mass balance. Under the same
assumptions, the momentum equation simplifies as well:

Q
t

# y
&
(Q 2 A)
+
+ gA % S 0 + S f + Se ( q =0vx = 0
x
$ x
'
=0

(10.7.13)
353

which can be simplified further by eliminating the dQ/dx term


implicit in the second term and noting that:

dy
dh
S0 =
dx
dx

(10.7.14)

where the Froude number (Fr) is defined as:

Fr =

gD

D=

A
;
B

B top width =

dA
dy

(10.7.17)

Substituting into Equation (10.7.16) and rearranging yields:


where h is the height of the water surface above the datum
(Figure 10.38). This yields:

dh d V 2
+

= S f Se
dx dx 2g

2. Steady-flow (with no lateral inflow) for a prismatic channel:


Under the same assumptions as above, but for a prismatic
channel (i.e. of a regular shape that does not change with x),
by definition Se = 0 (and the momentum coecient is
approximately 1.0), which yields a further simplification to the
momentum equation (expressed in terms of water depth y):

where by definition the second term is equivalent to:

d V
2 dy
=
F
r
dx 2g
dx
2

(10.7.18)

(10.7.15)

which is the commonly applied equation for gradually varied


flow in a natural channel. Note this is an ODE (not a PDE)
and can be integrated rather easily (most often numerically).

dy d V 2
+
= S0 S f
dx dx 2g

dy S 0 S f
=
dx
1 Fr2

(10.7.16)

which is the commonly used equation for gradually varied flow


in prismatic channels. Note that this equation is an ODE
(rather than a PDE) and can be integrated relatively easily.
3. Steady-uniform flow (with no lateral inflow): Simplifying
further for the case of uniform flow (with the same simplifying
assumptions as above), i.e. where there is no variation in the
x-direction, the momentum equation can be simplified:

Q
t

=0

y
(Q 2 A)
+
+ gA
x
x
=0

=0

S 0 S f Se q =0vx = 0

which yields:

S 0 = S f + Se

(10.7.19)

for natural channels, and:

S0 = S f

(10.7.20)

for prismatic channels. Both of these simply state that the


frictional and expansion/contraction head losses exactly
354

balance the elevation head gradient. Similar expressions can


be derived from energy considerations.

(empirical) Manning equation which can be expressed either


in terms of velocity (here using SI units):


The special case of uniform flow is one of particular
interest and worthy of additional discussion. From dimensional
analysis one can show that the bed shear stress associated
with friction is given by:

V =

V2
0 = Cf
= gRS 0
2g

(10.7.21)

where Cf is a resistance coecient and R is the so-called


hydraulic radius of the channel (cross-sectional area divided
by wetted perimeter [i.e. perimeter of cross-section in contact
with water]). Solving Equation (10.7.21) for the mean velocity
yields:

V =

2g
RS 0 = C RS 0
Cf

(10.7.22)

1 23 12
R S0
n

(10.7.24)

or in terms of flow (also SI units):

Q=

1
AR 2 3S 01 2
n

(10.7.25)

It is important to remember that these equations are valid for


steady-uniform flow (with no lateral inflow to the channel).
For nonuniform flow, the Manning equation can be generalized
by using Sf in place of S0. The Manning roughness parameter
depends on the type of channel bed and is often tabulated
(e.g. Mays, 2005).
E XAMPLE 10.7.1

which is often referred to as the Chezy Equation for uniform


flow. The Chezy coecient (C) represents a measure of the
roughness of the stream channel bed. The Manning equation
involves an empirical expression for C:

C =

1 16
R
n

(10.7.23)

where n is the Manning roughness coecient. Putting this


together with Equation (10.7.22) yields the so-called

Estimate the uniform flow in a prismatic


rectangular channel with a width of 3 m, a water
depth of 0.5 m, a bed slope of 0.001 and a
Manning roughness coecient of 0.02.
For uniform flow we can use the Manning equation. For
the channel/water depth conditions, the cross sectional
area is equal to: A = 1.5 m2. The hydraulic radius is
given by the ratio of the area to wetted perimeter:

355

E XAMPLE 10.7.1 ( CONTINUED )

E XAMPLE 10.7.2 ( CONTINUED )

A
(1.5 m 2 )
R= =
= 0.375 m
P (2(0.5 m) + 3 m)

the solution must satisfy:

The discharge is then given by:

1
Q=
(1.5 m 2 )(0.375 m)2 3(0.001)1 2 = 1.2 m 3 s 1
(0.02)

2 3

! (3 m)y $
((3 m)y) #
&
" 2y + (3 m) %

= (0.1 m 3 s 1 )(0.02)(0.001)1 2
= 0.0632

where again it should be noted that the Manning


equation is not dimensionally consistent.

Using an iterative solver, the depth that satisfies the


above equation is y = 0.101 m. Note that this depth is
much smaller than the width (i.e. 0.1 m vs. 3 m). In such
cases, a wide-channel assumption is often invoked
whereby:

E XAMPLE 10.7.2

P = B + 2y B, if B << y

Estimate the depth of uniform flow for the


channel in Example 10.7.1 if the discharge is 0.1
m3 s-1.

which yields a simpler (non-iterative) form of the


Manning equation:

First, the Manning equation can be expressed in terms of


depth of flow given a rectangular cross-section of width
B:
2 3

! By $
1
12
Q = (By) #
& S0
n
" 2y + B %
which shows that even for a simple cross-sectional
geometry, the nonlinearity of the Manning equation
requires an iterative solution for the depth y. Specifically,

2 3

" By %
1
1
Q (By) $ ' S 01 2 = By 5/3S 01 2
n
n
#B &
y = (QnS 01 2B 1 )3/5
For the conditions shown above this would yield a depth
estimate of 0.099 m, which is a close approximation to
the real solution. This approximation is only valid under
the wide-channel assumption mentioned above.

356


The Saint-Venant equations are sometimes referred to as
the dynamic wave equation because they fully describe the
dynamics of 1D unsteady flow in a channel. In this regard
they can be used to model any number of phenomena
including floods, tides, nonuniform flow, uniform flow, etc.
The price for this generality is one of computational demand
since the Saint-Venant equations are expensive to solve.

The kinematic wave equation is another special case of
the Saint-Venant equations that is generally more easily
solved. By definition a kinematic wave (as opposed to a
dynamic wave) is one where the acceleration terms, pressure
term, and lateral influx of momentum are all negligible, i.e.


The result of the above simplifications is essentially a
reduction of the two equation governing system to a single
governing equation. For uniform flow the momentum equation
can also be expressed in the general form relating A and Q as:

A = aQ b Q = cAd

If true, the momentum equation simplifies to the uniform flow


condition so that the two governing equations (continuity and
momentum) for the kinematic flow equation are:

A Q
+
= q(x,t)
t
x

(10.7.26)

S0 = S f

(10.7.27)

3 5

3
;b =
5

or

12
1 S0
c=
;
n P2 3

d=

5
3

(10.7.29)

where P is the wetted perimeter. The above form can be used


to derive an expression for the kinematic wave equation
(either in terms of the dependent variable Q or A). To get the
governing equation in terms of Q, we can use:

A = aQ b

where the above form of the momentum equation implies a


prismatic channel. So the kinematic wave equation is an
unsteady flow equation, but one where the wave motion is
determined primarily from mass balance. For many normal
floods in natural rivers the dynamic component of the wave

(10.7.28)

where, for example, with the Manning equation:

nP 2 3
a=

S0

Q (Q A)
y
+

gA
qvx 0
t

x
x

attenuates relatively quickly making the kinematic


assumptions reasonably accurate. From Equation (10.7.27) it
is clear that the flow is uniform.

A dA Q
Q
=
= abQ b1
t dQ t
t

(10.7.30)

where if this is substituted into the continuity equation yields:

abQ b1

Q Q
+
= q(x,t)
t
x

(10.7.31)

which is one equation in one unknown. Note that the equation


357

is still a PDE and therefore is generally solved numerically on


a discretized channel (Mays, 2005). Similarly, to get the
governing equation in terms of A we can use:

Q = cAd

Q dQ A
A
=
= cdAd 1
x dA x
x

(10.7.32)

which can be substituted into the continuity equation to get:

A
A
+ cdAd 1
= q(x,t)
t
x

(10.7.33)

which is again a single equation in a single unknown. In either


case, the streamflow hydrograph Q(x,t) can be determined,
either directly via solution of Equation (10.7.31) or indirectly
via solution of Equation (10.7.33) and then using Equations
(10.7.28) and (10.7.29).

358

S ECTION 8

Streamflow Routing: Hydrologic Routing



An alternative to hydraulic streamflow routing using the
dynamic wave method (or one of its special cases) is to use a
so-called hydrologic routing method. The primary dierence
between the two methods is that hydrologic routing is a
lumped approach, where a length (reach) of river is treated as
a lumped unit of storage (yielding flow at the reach outlet at
a given time), while hydraulic routing is a distributed
approach where the reach is discretized into many small pieces
to get the flow as a function of both space and time.

F IGURE 10.39 Stream channel reach storage conceptualization used in the Muskingum hydrologic streamflow routing
method (adapted from Mays, 2005).


Hydrologic routing starts with a lumped mass balance
equation for a specified reach of channel (Figure 10.39):

dS
= I(t) Q(t)
dt

(10.8.1)

where S is the stored water in the reach, I(t) is the inflow


hydrograph, and Q(t) is the outflow hydrograph. It is
generally assumed that the inflow hydrograph is known and
the goal is to obtain an estimate of the outflow hydrograph.
However Equation (10.8.1) alone is insucient, as a storageinflow-outflow relationship is needed to yield two equations in
two unknowns. The Muskingum method is a popular
hydrologic routing method that conceptualizes the storage in

the channel reach as two components (Figure 10.39): prism


storage (uniform cross-section across the reach) and wedge
storage. As a flood wave enters the reach, the wedge storage is
positive and when the flood recedes the wedge storage is
negative. The Muskingum method assumes the prism storage
is proportional to the outflow, i.e. KQ and the wedge storage
is given by KX(I -Q) where X is a weighting coecient
between 0 and 0.5. Note that if I > Q the wedge storage is
positive and if I < Q the wedge storage is negative. The total
storage is the sum of the prism and wedge storage so that:

S = KQ + KX(I Q) = K[XI (1 X )Q]

(10.8.2)
359

which shows that the storage is a linear function of the


weighted average of inflow and outflow. The parameter X is
typically on the order of 0.3 for natural streams. The
parameter K must also be determined and is roughly equal to
the travel time of the flood wave through the channel reach.

To perform the routing, the original mass balance
equation can be discretized as:

S j +1 S j
t

I j +1 + I j
2

Q j +1 + Q j
2

(10.8.3)

where j is the time index, the left-hand-side represents a finite


dierence approximation to the time derivative, and the righthand-side fluxes are approximated by their respective average
over the time step. Additionally, the storage change can be
represented by:

S j+1 S j = K[XI j+1 + (1 X )Q j+1 ] K[XI j + (1 X )Q j ]

(10.8.4)

Combining Equations (10.8.3) and (10.8.4) and simplifying


yields the following routing equation:

Q j +1 = C 1I j +1 + C 2I j + C 3Q j

(10.8.5)

where the constants defined above are given by:

C1 =

t 2KX
2K(1 X ) + t

(10.8.6)

C2 =

t + 2KX
2K(1 X ) + t

(10.8.7)

C3 =

2K(1 X ) t
2K(1 X ) + t

(10.8.8)

where the C1, C2, and C3 coecients sum to 1.0, meaning that
Equation (10.8.5) is simply a weighted average of the righthand-side terms. Provided the inflow hydrograph (as a
function of time), the initial outflow, and the parameters K
and X, Equation (10.8.5) can be applied recursively to yield
the outflow hydrograph. Methods for determining the
parameters generally require historical inflow and outflow
hydrographs or estimation via other means (Mays, 2005).

Cunge (1969) provided a connection between the
Muskingum method and the kinematic wave method. Recall
that the kinematic wave model is a hydraulic routing method
that provides a distributed estimate of discharge (i.e. at
discretized locations along the channel). One can express the
discretized solution of the kinematic wave model as:
j+1
j
Qi+1
= C 1Qij+1 +C 2Qij +C 3Qi+1

(10.8.9)

where the i index represents the location in space and the j


index represents the time step, which is in the exact form of
Equation (10.8.5). The use of Q at varying locations
represents the distributed nature of the estimate. The
coecients in Equation (10.8.9) are the same as those in
Equations (10.8.6)-(10.8.8). Cunge (1969) showed that when
K and the time step are constant, the above equation is an
approximation to the kinematic wave equation (Mays, 2005).
This representation is often referred to as the MuskingumCunge method.
360

S ECTION 9

Measurement of Streamflow

In the preceding sections it was assumed in various
locations that hydrograph data was available, i.e. at a basin
outlet. Here we briefly outline the primary ways those
measurements are made.

As with many hydrologic variables, streamflow is
generally not measured directly, but estimated indirectly from
measurements that are more straightforward to make. In a
stream, the measurements which are easiest to make include
water depth and velocity. In general, streamflow is estimated
via a rating curve, which is a derived or known from a
relationship between streamflow depth and discharge (flow).

Examples where a rating curve is known theoretically (or
empirically) include flow over or through certain constructed
structures (often called weirs or flumes; Mays, 2005; WMO,
2010). Figure 10.40 shows an example of a weir which is
constructed into a channel reach. The primary point of such
structures is to force the flow to go through a known flow
state transition, i.e. critical flow, which is a transition
between supercritical and subcritical flow conditions (see
Mays, 2005 for a thorough discussion of critical flow and
depth). This is usually accomplished via a constriction in the
flow geometry. When the flow is critical and the structure
geometry is known, there are empirical relationships between
flow and the critical depth or measured head. In such cases,

F IGURE 10.40 Schematic (top panel) and picture of an in-

stalled (bottom panel) v-notch weir in a stream channel for estimating streamflow (from WMO, 2010).

the measurement of depth at some point in the structure can


be used to estimate the corresponding flow. In such structures,
flow depth is often measured automatically via a pressure
sensor or other mechanism.
361


Another mechanism for deriving rating curves and
streamflow measurements is via the manual measurement of
mean velocities in a channel cross-section. Velocity in a stream
can be measured using vertical axis mechanical current meters
(WMO, 2010; Figure 10.41), which are devices placed
perpendicular to the flow. The flow velocity turns a propellor
or anemometer (attached to a wading rod) which can then be
used to back out the velocity at the measurement point. It
should be noted that flow in a stream is generally turbulent,
which means that the velocity field can vary significantly (and
somewhat randomly) in space and time. To get an accurate
measurement of the mean velocity, measurements should be
taken over a long enough period to average over the turbulent
eddies. While flow is often written as: Q = VA, the V is the
average velocity across the entire stream cross-section, so

F IGURE 10.41 Schematic of a Price current meter (from WMO,


2010).

generally multiple measurements are required. More precisely,


one can write the flow as the integral (over the cross-sectional
area):

Q = V dA
A

(10.9.1)

In practice, the cross-section can be discretized as shown in


Figure 10.42. Based on this, the flow could be estimated via:
n

Q = Viyi wi

(10.9.2)

i =1

where Vi is the average velocity across each vertical strip and


the product of depth and discretized width is the area of the
strip. Note that the velocity generally has a profile in the
vertical, with zero velocity at the stream bed and highest
velocity at the surface. A simple rule of thumb for the average

F IGURE 10.42 Illustration of methodology for estimating

streamflow from stream cross-section area and multiple velocity measurements (adapted from Mays, 2005).

362

velocity is to take the average of two velocity measurements


taken at a height of 20% and 80% of the depth at that
location. Using this manual approach, streamflow discharge
can be determined for a given river stage (depth). If done
several times spanning dierent flow conditions, one can
develop a rating curve. Once developed, a single manual depth
measurement could be used (along with the rating curve) to
estimate the discharge.

363

S ECTION 10

13. What is excess or eective precipitation?

Conceptual Questions

14. Suppose you have a constructed 1-hour unit hydrograph.


Describe schematically how you would generate a 2-hour
unit hydrograph from your 1-hour unit hydrograph.

1. Describe the mechanism behind infiltration excess runo.

15. Describe the main dierences between the UH method and


physically-based lumped and distributed modeling
approach.

2. Describe the mechanism behind saturation excess runo.


3. Describe the mechanism behind interflow.
4. Describe the mechanism behind baseflow.
5. What is a variable contributing area?

16. What two equations are used to derive the Saint-Venant


Equations?
17. Describe the primary methods for how streamflow is
measured.

6. What are the components of a streamflow hydrograph?


7. Describe how storm characteristics can impact an outlet
hydrograph.
8. Describe how basin characteristics can impact an outlet
hydrograph.
9. What is the primary assumption in the unit hydrograph
method? What principle does that assumption allow us to
apply?
10. What inputs are required for the construction of a D-hour
unit hydrograph?
11. What does D-hour refer to in the context of a unit
hydrograph?
12. What is stormflow?
364

Construct the 1-hr unit hydrograph from the known 0.5-hr


unit hydrograph. Your answer should be in tabulated form.

S ECTION 11

Sample Problems
Problem 10.1. You are hired to assess the flash flood
potential of a given basin, where flash floods are generally
associated with short duration (high intensity) storms that
generate infiltration excess runo. The basin is composed of a
homogeneous silt loam soil. To comply with regulations for
the region, dierent downstream infrastructure must be
designed for various return-period storm events, where the
return-period is associated with the probability (and therefore
magnitude) of the storm event. The 1-hour duration design
storm events for the 10-year and 25-year return periods in this
region are:
STORM RETURN PERIOD

10-year

25-year

PRECIP. INTENSITY

2 cm/hour

7 cm/hour

a) For this particular basin, what is the expected (if any)


flash flood stormflow (i.e. infiltration excess runo) for each
return period design storm? Design regulations dictate that it
should be assumed that 75% of the soil pore space is filled
prior to the storm. The Philip model can be used as an
infiltration capacity model (as needed).
b) A previous analysis of the basin found the 0.5-hr unit
hydrograph to be given by:

TIME
(HOUR)

FLOW (m3/s/cm)

0.5

1.0

10

1.5

15

2.0

2.5

3.0

c) Based on the constructed 1-hr unit hydrograph, what is the


predicted peak flash flood stormflow (i.e. design flow) for the
two return period storms analyzed in part a). The infiltration
excess runo values computed above can be treated as the
eective (excess) precipitation.
Problem 10.2. An environmental monitoring agency has a
rain gauge and a river gaging station installed at the outlet of
a basin. A rainstorm with 2 hours of eective rainfall
produced 2.6 cm of runo and resulted in the following
observed total hydrograph for the stream:
a) Assuming that the baseflow can be calculated by drawing a
straight line connecting the end points of the hydrograph,
365

TIME
(HOUR)

DISCHARGE
(m3/s)

3.11

3.45

6.51

16.36

18.25

12.28

8.29

5.72

4.53

10

3.11

separate the baseflow and the direct runo contributions to


runo. On the same plot, show the baseflow, direct runo, and
total discharge in m3/s.

Problem 10.3. A small watershed is located in an area that


is frequently hit by intense thunderstorms. As the engineer
responsible for designing a new flood protection structure, you
are asked to determine the peak design flow for which the
structure will be built. From your excellent hydrologic training
you know that shorter duration events have the possibility of
being of higher intensity and vice versa. So you decide to
investigate the peak flows for two possible extreme
thunderstorm events (which are the largest expected amount
to fall in each duration storm): 1) a one-hour storm with a
cumulative eective rainfall of 24 cm and 2) a three-hour
storm with a cumulative eective rainfall of 45 cm. The 1-hr
unit hydrograph for the watershed is shown in the table
below.
a) Estimate the peak stormflow corresponding to each storm

1-HR UH
TIME
(HOUR)

UH FLOW
(m3/s/cm)

b) Find and plot the 2-hour unit hydrograph for the basin.

10

c) Estimate and plot the 6-hour unit hydrograph.

d) What is the total hydrograph resulting from a storm with


12-hour eective rainfall that delivers 2 cm of eective rain in
the first half and 3 cm of eective rainfall during the second
half? Plot the resulting total hydrograph.

(based on a unit hydrograph analysis). Which of the two


storms yields a higher peak runo?
366

b) Based on further investigation you determine that the


basin is composed of soils with a large saturated hydraulic
conductivity and that the unit hydrograph was constructed
from a typical intensity storm. What runo generation
mechanism would you reasonably hypothesize was primarily
responsible for the basin response seen in the unit
hydrograph?
c) For the extreme events analyzed above, would you expect
the same runo generation mechanism to be responsible for
the peak stormflow? Explain your reasoning and what, if any,
additional mechanism/s might contribute to stormflow. How,
if at all, would you expect this to impact the accuracy of the
UH-based prediction? Explain.

367

Chapter 11

A Simple
Watershed
Model

S ECTION 1

Learning Objectives
By the time you finish this chapter you should be able to:
1. Understand the key components (i.e. inputs, governing
equations, and outputs) of a distributed watershed model
2. Describe the basic idea behind the TOPMODEL approach
3. Understand how analytical mass/energy balance equations
can be discretized and solved numerically
4. Setup and run watershed simulations using the provided
MOD-WET numerical watershed model code
5. Reconcile model simulation output (i.e. watershed
response) with your hydrologic understanding developed
in previous chapters.

369

S ECTION 2

Development of a Distributed Watershed Model Using MOD-WET



The previous chapters have covered the various physical
processes involved in hydrology. Chapter 10 (Section 6)
provided an introduction into how these process-level physics
can be tied together into a unified framework via a distributed
watershed model. Many existing examples of distributed
models have been developed and are available for use in
hydrologic modeling (HEC-HMS, MIKE-SHE, tRIBS, etc.).
These models vary in many ways including: level of
conceptualization, physical process representation, degree of
lumped vs. distributed representation, and numerical
implementation. In this chapter an example of such a model is
put forth that is implemented in MATLAB and ties together
many of the processes covered throughout the book using
MOD-WET, thereby illustrating the modularity of the
framework. The goals are to: 1) develop a unified framework
that builds on what has been used previously and 2) provide a
model for qualitative and quantitative understanding of
hydrological processes and how they change as a function of
watershed properties. The MOD-WET model may not be
suitable for all applications, but provides a relatively userfriendly framework for developing and testing basic

understanding and intuition of distributed hydrologic


processes in a watershed.

In developing a distributed watershed model,
computational expense is generally an important factor.
Physical process computations must be performed at each
pixel in a watershed so that, depending on the model
resolution and area (i.e. total number of pixels),
computational expense is many orders of magnitude greater
than that of a lumped model. Since the model to be used here
is meant to foster understanding of sensitivities of model
response to inputs/parameters, hypothesis testing, etc., we
intentionally choose a distributed framework that attempts to
minimize computational expense. As such, several simplifying
assumptions are made. Every attempt is made to clearly
identify the primary assumptions and simplifications in the
development below.

The underlying framework used here for runo is based
on one of the first distributed watershed models, typically
referred to as the TOPography based hydrologic MODEL
(TOPMODEL) developed by Beven and Kirkby (1979). The
model was motivated by the increased availability of spatially
distributed topographic data, i.e. the DEM data we have used
in earlier chapters. TOPMODEL was developed primarily to
predict saturation excess runo due to shallow groundwater
within a basin. So the TOPMODEL framework is used to
represent unsaturated zone-saturated zone interactions with
an emphasis on runo. Other components can be coupled to
TOPMODEL including snow accumulation and melt and
370

surface energy balance (including distributed radiative


forcings).

The first key assumption used in the TOPMODEL
framework is that there is a similarity function that describes
hydrologic runo response. The similarity function originally
proposed was the so-called topographic index for pixel i:

! a $
i = ln # i &
" tan Si %

(11.2.1)

where ai is the upstream (drainage) area per unit contour


length for a given pixel in the basin ([ai]=L2L-1) and Si is the
slope of the pixel. In the case of a raster grid, the per unit
contour length amounts to the grid resolution. Note that this
index is a static map for a basin that can be computed from
DEM data. Pixels where flow is expected to converge (large
upstream area and/or shallow slopes) will yield large
topographic index values, while areas of divergence will yield
small topographic index values. Figure 11.1 shows the
topographic index computed for the DEM shown in Figure
10.1. Note that the highest index values tend to occur at
expected stream locations and the lowest values occur at
upstream areas with steep slopes. Note that based on
Equation (11.2.1), the index is not dimensionless (i.e. [ln(L)]).
This is generally an undesirable property, however this is not
overly problematic since only the relative value of the index is
used to determine the type of response of one pixel compared
to another. Specifically, it is assumed that pixels with the
same value of the index will respond identically with respect

F IGURE 11.1 Topographic index for an example DEM.

to runo. As will be shown below, this yields significant


computational savings in that repeated computations are not
necessary for pixels with the same index value. This
assumption will not always be a good one.

The second key assumption used in TOPMODEL is that
the saturated hydraulic conductivity in the unsaturated soil
zone decays exponentially with depth, i.e. for pixel i:

K s = K 0 exp(z i / m)

(11.2.2)

where K0 ([K0]=L T-1) is the surface saturated hydraulic


conductivity and m is an exponential decay parameter which
has dimensions: ([m]=L). This assumption has been shown to
be valid for some soils and invalid for others (Beven, 1997).
The model can be generalized for other profiles as shown in
371

Ambroise et al. (1996). It should also be mentioned that the


parameters are often treated as calibration (or eective)
parameters and hence might not match tabulated values for a
given soil type. From the above assumption, one can
alternatively define the corresponding transmissivity
(Sivapalan et al., 1987):

T(z i ) =

K(z)dz K m exp(z

zi

/ m) = T0 exp(z i / m)

(11.2.3)

figure of the pixel-scale processes is shown in Figure 11.2. In


this version, there is no vegetation represented in the model.
Future versions could include vegetation processes. The
conceptualization connects to the original TOPMODEL via
the prediction of the saturation deficit (SD; where [SD] = L )
and baseflow (qb) as shown below (where the pixel index is
dropped for simplicity). The key meteorological inputs to the

where zi is the depth to the groundwater table (which is often


replaced by the storage (or saturation) deficit SDi (detailed
below) and T0 is the saturated surface transmissivity (equal to
K0 m) which has dimensions: [T0]=L2T-1. From this, the
topographic index can be replaced with the soil-topographic
index, i.e.:

! a
$
i
i = ln #
&
T
tan
S
" 0
i%

(11.2.4)

where this can alternatively be used as the similarity function


to predict hydrologic response. From these two key
assumptions, various forms of the TOPMODEL approach
have been developed. The original approach was primarily
developed to predict baseflow and, by tracking the locations
where the water table intersected the surface (using the
topographic index), to identify variable contributing areas and
saturation excess runo.

Here the TOPMODEL approach is expanded in a similar
way to that shown in Takeuchi et al. (1999). A schematic

F IGURE 11.2 Schematic of the pixel-scale processes in the distributed model.

372

model are rainfall/snowfall (P), and the resulting infiltration,


and those that drive evapotranspiration (E) at the surface.
The key outputs are overland flow which consists of saturation
excess runo (qse) and infiltration excess runo (qie) and
groundwater runo or baseflow (qb). Mass balances are applied
using unsaturated and saturated zone moisture states as
described below. Note that the model is conceptual in nature
in that mass balances are applied to various buckets or
reservoirs rather than via a vertical discretization of the
domain and solution of the unsaturated zone moisture budget
using Richards Equation (Equation (7.4.7)). Such a bucket
model approach is typically taken for computational savings.

The form of the surface mass and energy balance
depends on whether or not there is snow cover. For the case
with no snow cover, the prognostic (soil) surface energy
balance is solved using the so-called force-restore method
(Noilhan and Planton, 1989):

dTs
= CT [Rn LE H ] C d (Ts Td )
dt

(11.2.5)

where Ts is the surface soil temperature, CT is related to the


surface heat capacity and Cd is the diurnal periodicity (i.e. 24
hours), and Td is a deeper soil temperature. The first term
serves to force the surface temperature via the surface heat
conduction and the second term restores the surface
temperature exponentially to Td. The deep layer temperature
is computed as a low-pass filter applied to the surface
temperature (Noilhan and Planton, 1989). The surface soil

water mass balance is built into the TOPMODEL framework


described below.

For a snow covered surface (see Chapter 6), a simple 1layer snow mass and energy balance is used. The snow mass
balance at each pixel is given by (Equation 6.4.1) in terms of
the snow water equivalent (SWE):

dSWE
= P E M
dt

(11.2.6)

where P, E, and M are the (snowfall) precipitation,


evaporation/sublimation, and melt output/runo respectively.
Computational eort is saved by not simulating snow density
or depth. The surface energy balance for the snowpack is
applied using a prognostic equation for the single-layer snow
temperature (Tsnow) at each pixel:

C snow

dTsnow
= Rn LE H wLf M
dt

(11.2.7)

where Csnow is the snow-layer heat capacity (J m-2 K-1) and the
right-hand-side terms are the net radiation, latent and
sensible heat fluxes, and latent heating due to melt. Advected
energy by rain could also be included.

For simplicity, the soil surface energy balance is only
solved for prognostically when snow disappears. In doing so,
the soil surface temperature is set equal to the snow
temperature (i.e. 0C) just before the snow disappears. This is
obviously a simplification as it ignores soil dynamics under
373

snow, but this choice is made for computational savings. More


complicated models can be used to solve the energy balance
across the entire snow-soil continuum and include soil freezethaw processes.

The unsaturated zone is represented by two moisture
states: the rootzone soil moisture storage (Srz) and the
unsaturated zone storage (Suz). The rootzone storage is the
near-surface reservoir that is filled by infiltration (f ; see TCA
method in Chapter 7; Section 7) and is depleted by
evapotranspiration (E ; see Chapter 8) and drainage (qdrain) to
the main unsaturated zone reservoir:

dSrz
= f E qdrain ;
dt

Srz min Srz Srz max

(11.2.8)

Implicitly, what does not infiltrate generates infiltration-excess


runo, qie. In the MOD-WET model this is done using a
simplified form of the TCA method using the precipitation
over the time step as input and the time step as the storm
duration. The loss from the lower boundary of the rootzone
is assumed to be dominated by gravity drainage and is
modeled to only occur when the rootzone storage overfills the
maximum storage capacity. Hence the drainage flux involves a
threshold process and the storage in the rootzone is always
less than or equal to the specified upper limit Srzmax (e.g. field
capacity). If the upper limit is set to a value lower than the
soil porosity, this implies that the rootzone never fully
saturates and this can have implications on infiltration excess
runo. It is also assumed that the storage cannot go below a

specified lower limit Srzmin (e.g. permanent wilting point). The


larger the rootzone reservoir, the more infiltration will
ultimately be partitioned into evaporation. The smaller the
reservoir or more intense the infiltration rate, the more
drainage flux there will be.

The unsaturated zone storage is fed by the rootzone
drainage flux and depleted by recharge (qv) to the underlying
groundwater system:

dSuz
= qdrain qv ;
dt

0 Suz SD

(11.2.9)

where the unsaturated zone storage can go to zero if fully


depleted by recharge or if the groundwater table rises to the
surface (i.e. SD = 0). Implicit in the latter case is the
generation of saturation excess runo, which is another
threshold process (only non-zero when the groundwater table
reaches the surface). The recharge flux is modeled as:

qv = K 0 exp(SD / m)

(11.2.10)

assuming there is adequate storage in the unsaturated zone.


This drainage flux is eectively equal to the assumed
saturated conductivity at the water table (i.e. only the gravity
term is relevant) and increases with increasing surface
conductivity (and m) or decreasing saturation deficit.

The groundwater storage is depleted by the baseflow
(volumetric flow per unit width, i.e. [qb]=L2 T-1) leaving the
pixel, which is given by:
374

qb = T0 exp(SD / m)tan S

(11.2.11)

which indicates that baseflow increases with increasing


transmissivity or slope (S) and decreasing saturation deficit.

Finally, a basin average mass balance and similarity
arguments (using the topographic or soil-topographic index)
are used for the saturation deficit. The basin average
saturation deficit mass balance is given by:

d SD
dt

= Qb Qv

(11.2.12)

where the angled brackets denote a spatial average over the


basin and Qb and Qv are basin-averaged baseflow and recharge
fluxes respectively. It is then assumed, using similarity
arguments, that the pixel-scale saturation deficit is a function
of the basin average deficit and the departure of the local soiltopographic index from its basin-average, i.e.:

SD = SD + m "# i $%

(11.2.13)

where the second and third term on the right-hand-side are


constant in time. What this similarity argument implicitly
states is that SD at a given pixel will go up or down with the
basin-average saturation deficit, but that the pattern within
the basin is dictated by the topographic or soil-topographic
index. Pixels with a soil-topographic index greater than the
mean will have a lower saturation deficit (i.e. convergent areas
will have a higher water table) and vice versa. Physically, the
saturation deficit should be greater than or equal to zero,
where a value of zero corresponds to the water table being at

the surface. Pixels at which a negative value is obtained using


Equation (11.2.13) indicate regions where the groundwater
table intersects the surface and therefore generate saturation
excess runo as described in more detail below.

The model described above generates runo fluxes: qie,
qse, and qb at each pixel in the domain. To generate the outlet
hydrograph, the pixel-wise runo fluxes must be moved
downstream via a so-called routing scheme (see Chapter 10,
Sections 7 and 8). A key input to the routing scheme is the
channel topology of the basin. In the MOD-WET model, we
choose to have a network that covers the entire basin as
shown in Figures 10.36 and 10.37. The connectivity of the
cells is implicit in the automated basin delineation used in the
MOD-WET function:
watershed_area_and_stream_delineation.m, which has an
output called flowdir. From that variable all upstream and
downstream nodes for each link in the network are identified.

The routing equations are applied to each link of the
network, taking into account both flow in the stream and
inflows from pixel runo. For simplicity, it is assumed that the
channel is of rectangular (prismatic) cross-section. The spatial
distribution of channel width (B) across the network is
modeled via a two-parameter power law (Takeuchi et al.,
1999):

B = Acc

(11.2.14)

where Ac (km2) is the contributing upstream area flowing into


a given pixel. This expression, given the specification of the
375

two parameters, provides a simple means of estimating the


increase in channel width downstream. However, using the
equation does not prevent cases where (for large upstream
areas) the stream channel width could be larger than the pixel
resolution. The Manning roughness parameter for a given
pixel is given by the model (Takeuchi et al., 1999):

!
tan S
n = n0 #
# tan S
"

1/3

$
& ; S mean slope
&
%

K=

dx
ck

(11.2.17)

(11.2.15)

where n0 is the mean Manning roughness over the basin. This


expression indicates that areas of higher than average slope
have higher roughness values and vice versa. Note that such
an expression does not incorporate explicit dierences as a
result of dierent landcover types.

For simplicity, the Muskingum-Cunge method (Equation
10.8.9) is used for routing as done in Takeuchi et al. (1999). In
this case there is an implied river reach connecting a given
upstream and downstream pixel. The routing of flow is
determined by the relationship:
j+1
j
Qi+1
= C 1Qij+1 +C 2Qij +C 3Qi+1


Aside from the spatial discretization of the river network
(dx) and temporal discretization (dt), the C coecients
depend on the dynamic parameters K and X. The parameter
K is given by:

(11.2.16)

where the i index represents the location in space (i.e.


upstream or downstream pixel), the j index represents the
time step and the C coecients are weighting factors that
depend on flow characteristics (Equations (10.8.6-10.8.8)). In
this formulation, upstream flows are the sum of those already
in the channel from upstream and those contributing runo
from the upstream pixel.

where ck is the so-called wave celerity, which is defined by the


derivative of flow in the channel with respect to cross-sectional
area:

ck =

dQ
dA

(11.2.18)

which can be determined via the use of the Manning equation


(Equation (10.7.25)):

Q=

1
1
AR 2 3S 01 2 = A5/3P 2 3S 01 2
n
n

(11.2.19)

The celerity is then given by:

ck =

dQ
5 2/3 2 3 1 2
=
A P S0
dA 3n

3/5

1/2
5 " S0 %
= $ 2/3 ' Q 2/5
3 $#nP '&

(11.2.20)

which is itself a function of the nominal flow which can be


estimated using a so-called 3-point average:
j
Q = (Qij+1 +Qij +Qi+1
)/3

(11.2.21)
376

Note that the general form in Equation (11.2.20) is iterative


since the wetted perimeter depends on depth. To make the
above formulation non-iterative, a wide channel assumption
is invoked, whereby the wetted perimeter in the Manning
equation is assumed approximately equal to the width of the
channel, i.e. (Takeuchi et al., 1999):

P = B + 2y B (for B >> y)

the soil/snow surface, etc. Future versions of the MOD-WET


model will embed these processes in the model.

(11.2.22)

where the width is given by Equation (11.2.14). The X


parameter is given by:

"
%
Q
X = 0.5 $$1
''
Bc
tan(S)dx
#
&
k

(11.2.23)

For a given time step in the model, Equations (11.2.14)(11.2.23) can be applied to each reach of the stream network
from upstream nodes all the way downstream to the basin
outlet. In some cases the flow velocity may be larger enough
that it can cover the distance between two pixels in less than
the specified time step. A dynamic time step can be used to
reduce the routing time step for such cases. The streamflow at
the basin outlet is the outlet hydrograph.

It should be reiterated that the version of the MODWET model described herein does not contain explicit
representation of vegetation. As such, it is best suited to nonvegetated basins. We would expect vegetation to modify the
processes in a variety of ways including: rainfall/snowfall
canopy interception, additional control on evapotranspiration
via stomatal resistance, attenuation of solar radiation reaching
377

S ECTION 3

Numerical Implementation
of the MOD-WET Model

The MOD-WET model defined in Section 2 needs to be
implemented numerically since analytical solutions are not
generally available. Numerical solutions are obtained by
discretizing the equations in space and time. Here the spatial
discretization is chosen to coincide with the underlying DEM
or other spatially-distributed fields that are necessary inputs
for the model. So, for example, a typical spatial discretization
may be on the order of 30 m x 30 m using readily available
DEMs. Alternatively, the DEMs may be coarsened to a lower
resolution to reduce computational expense. The time
discretization is done via a time step dt. The choice of the
time step is generally governed by the dynamics of the system.
To resolve the surface energy balance, a time step of less than
an hour may be necessary, while for more slowly varying
dynamics a time step as large as a day may be acceptable.

For numerical implementation, the above equations are
posed by representing all storage (or deficit) states in terms of
equivalent depth of water and the energy state at the surface
is represented by surface temperature (either snow or soil). In
the discretization used below, the indices i and t are used to
represent spatial and temporal dependence respectively. In the
solution of all dierential equations, an explicit forward finite
dierence scheme is used for computational eciency. This is

done mostly because it allows for easy vectorization of the


state equations which allows one to avoid using for loops in
MATLAB, which can be a very expensive proposition. It
should be noted that this introduces the possibility of
numerical accuracy and stability issues. For stability, a short
duration (15 min) time step is used. As mentioned above, for
additional computational eciency, the model solves a single
surface energy balance model at each pixel, either for snow or
soil, depending on the surface conditions. The surface
condition is tracked via an evolving snow mask.

For non-snow-covered pixels the discretized surface
energy balance is given by (based on Equation 11.2.5):

Ts (i,t + 1) = Ts (i,t) +

(11.3.1)

dt{CT [Rn (i,t) LE(i,t) H(i,t)]

C d (Ts (i,t) Td (i,t))}


For snow-covered pixels the discretized single-layer snow
mass balance and snow surface energy balances are given by
(based on Equation 11.2.6):

SWE(i,t + 1) = SWE(i,t) + dt[P(i,t) E(i,t) M(i,t)]

(11.3.2)

and (based on Equation 11.2.7):

Tsnow (i,t + 1) = Tsnow (i,t) +

dt
[R (i,t) LE(i,t) H(i,t) wLf M(i,t)]
C snow n

(11.3.3)


378

As noted above, there are several threshold processes involved


in surface and subsurface mass balance equations. The way
these are handled below is through the usage of intermediate
calculations which are then checked to satisfy thresholds and
bounds. All intermediate variables (i.e. before being finalized
at the end of the time step) are denoted using a primed
superscript. Equation (11.2.8) can be written by first
computing an intermediate storage amount assuming there
was no drainage:

Srz! (i,t + 1) = max{Srz (i,t) + dt[f (i,t) E(i,t)], Srz min }

(11.3.4)

and then correcting for the drainage flux:

qdrain (i,t) = max{(Srz! (i,t + 1) Srz max ), 0}

(11.3.5)

Srz (i,t + 1) = Srz! (i,t + 1) qdrain (i,t)

(11.3.6)

where f(i,t) is the infiltration rate due to rain/snowmelt


([f ]=L T-1), E(i,t) is the evapotranspiration from the rootzone
([E ]=L T-1), both of which are multiplied by the time step to
get the equivalent depth change due to the fluxes during that
time step, and qdrain is the equivalent drainage flux over the
time step (i.e. [qdrain ]=L). Writing the discretized version of
Equation (11.2.8) in this way takes into account the threshold
nature of the drainage flux and ensures that the rootzone
storage is always maintained within the correct lower and
upper bounds.

Equation (11.2.9) can be discretized in a similar way to
model the unsaturated zone storage evolution by first
computing an intermediate storage amount augmented by the

drainage flux from the rootzone:

Suz! (i,t + 1) = Suz! (i,t) + qdrain (i,t)

(11.3.7)

and then correcting for the recharge flux:

qv (i,t) = min{K 0 exp(SD(i,t) / m)dt, Suz" (i,t + 1)}

(11.3.8)

Suz! (i,t + 1) = Suz! (i,t + 1) qv (i,t)

(11.3.9)

where the recharge flux is given by gravity drainage unless


there is insucient storage to supply the full flux amount.
Equation (11.3.9) is still labeled as an intermediate state since
this neglects the possibility of saturation excess overland flow.
Physically, the upper bound is the saturation deficit (SD(i,t)).
Saturation excess runo is generated when the unsaturated
zone fills up beyond the local storage deficit. This amount of
flux can be written as:

q se! (i,t) = max{(Suz! (i,t + 1) SD(i,t)), 0}

(11.3.10)

where the saturation excess runo has depth dimensions. The


first term on the right-hand-side will only be positive when
the predicted unsaturated zone storage exceeds the saturation
deficit. Otherwise, the saturation excess overland flow will be
zero. When there is saturation excess runo it needs to be
extracted from the unsaturated zone storage term, which can
be updated via:

Suz! (i,t + 1) = Suz! (i,t + 1) q se! (i,t)

(11.3.11)

Finally, the groundwater discharge from each pixel is a


function of the local saturation deficit and other parameters
379

as given by:

qb (i,t) = T0 exp(SD(i,t) / m)tan Si

(11.3.12)

The dimensions of qb are: [qb ]=L2T-1, which implies a flow per


unit width (of stream or hillslope). Further, the basin-average
recharge flux Qv (with dimension [Qv] =L) used in Equation
(11.2.12) is given by:

1
Qv (t) =
N

q (i,t)
i=1

(11.3.13)

where N is the number of pixels in the basin. The integrated


baseflow flux used in Equation (11.2.12) is then computed via:

dt 1
Qb (t) =
Lx N

q (i,t)
i=1

(11.3.14)

where Lx is the dimension of the pixel length. From the above


set of discretized equations, Equation (11.2.12) is used to
update the basin-averaged saturation deficit and the pixelwise saturation deficit is then given by:

SD !(i,t + 1) = SD (t + 1) + m #$ (i)%&

(11.3.15)

where

!
$
a(i)
(i) = ln #
&
T
(i)tan
S(i)
" 0
%

(11.3.16)

!
$
$
a(i)
1 N !
a(i)
= ln #
& = ln #
&dA(i) (11.3.17)
T
(i)tan
S(i)
A
T
(i)tan
S(i)
" 0
%
" 0
%
i=1

where an intermediate value of the saturation deficit is


denoted in Equation (11.3.15) since there is no constraint on
non-physical values in its application.

Final updates to several variables revolve around
removing negative storage deficits that would be associated
with saturation excess flow via:

"
$ q ! (i,t) + SD !(i,t + 1) , SD !(i,t + 1) < 0
q se (i,t) = # se
q se! (i,t),
otherwise
$%

(11.3.18)

Additionally, when runo occurs, this implicitly means the


saturation deficit should be zero, which can be maintained
via:
"$ 0, SD !(i,t + 1) < 0
SD(i,t + 1) = #
(11.3.19)
$% SD !(i,t + 1), otherwise
This will keep the deficit to a minimum value of zero if
Equation (11.3.15) yields a negative value. The unsaturated
zone storage is set to zero where the saturation deficit is zero:

"$ 0, SD !(i,t + 1) < 0


Suz (i,t + 1) = #
$% Suz! (i,t + 1), otherwise

(11.3.20)

The above equations should be applied at each time step to


maintain the basin-wide soil moisture budget. Given a state
estimate at the beginning of a time step, the state estimate at
the end of the time step is determined using the equations
shown above. All fluxes are representative of the flux over the
time step.
380


Initialization of the model can be done by specifying an
initial condition of the basin-averaged saturation deficit and
then applying Equation (11.3.15). Other initial states that
need to be specified include the rootzone and unsaturated
zone storage values. A summary of the states and parameters
(in addition to the DEM and meteorological inputs) that need
to be specified for this numerical implementation of the model
are shown below in Table 11.1. Many other parmeters must
also be specified. In the default implementation of the model,
the soil parameters (roughness, hydraulic, thermal, and

T ABLE 11.1. K EY MOD-WET M ODEL P ARAMETERS

VARIABLE

DESCRIPTION (SI UNITS)

SD(t 0 )

Initial basin-average saturation deficit [m]

Suz (i,t 0 )

Initial unsaturated zone storage [m]

Srz (i,t 0 )

Initial rootzone zone storage [m]

Srz max

Maximum allowable rootzone storage [m]

Srz min

Minimum allowable rootzone storage [m]

T0

Surface saturated transmissivity [m2/s]

Conductivity decay parameter [m]

radiative properties, etc.) and snow surface characteristics


(roughness, emissivity, etc.) are the same throughout the
basin, but they could be made variable. As such, the primary
spatial variability comes in the form of topography which
directly impacts i) the redistribution of water via Equation
(11.3.15) and the routing of flow and ii) the meteorological
forcings including (among others) air temperature, which is
disaggregated via a lapse rate, shortwave radiation which
depends on elevation, slope, and aspect, surface precipitation
and phase (i.e. rain vs. snow), which depends on air
temperature.

The numerical implementation of the model is done
using a MATLAB model driver function called
MOD_WET_model_driver that is primarily used to: 1) call
various functions to initialize the model and load input files,
2) call MOD-WET functions from within a time-stepping loop
for solving mass and energy budget equations, and 3) to
collect and save outputs. The model has been tested on
several basins and should run fine provided you specify the
correct path to the MOD-WET and to the necessary input
files. The primary model function that the user must modify
is called MOD_WET_model_static_and_control_parameters
which contains the model control parameters (i.e. time step,
number of days to simulate, etc.), pointers to input filenames,
and physical parameters, including those listed in Table 11.1.
The necessary variables and their units are described in the
code. The set of parameters are stored in two structured
arrays: control_params and params to ease the passing of
variables. The files that need to be provided to the model are
381

those related to the DEM and watershed delineation (i.e.


outputs from the watershed_area_and_stream_delineation
function that must be applied as a pre-processing step) and
the meteorological inputs (which are assumed to come from a
single gage at a specified elevation within or near the basin).
The meteorological data must be at the appropriate time step
specified in the inputs file and is assumed to start at the
beginning of the water year on the Universal Time Coordinate
(UTC; or Greenwich Mean Time), i.e. at UTC 0:00 on
October 1st. Beyond the creation of these input files and the
editing of MOD_WET_model_static_and_control_parameters,
no other functions should need to be modified to run the
model.

Within the MOD_WET_model_driver function, maps of
model states and parameters and pre-allocation of variables
are initialized via the initialize_model function. The time
stepping loop updates state variables at each time step. Doing
so involves several functions including those that sequentially:
distribute the meteorological station data
(distribute_met_forcings), apply snow model physics
(snow_model) and soil energy balance physics
(soil_SEB_solver_prognostic), estimate infiltration and
infiltration excess runo using the time-compression
approximation (TCA_infiltration), estimate subsurface
processes and saturation excess runo and baseflow using the
TOPMODEL framework (TOPMODEL), and finally routing of
the runo downstream using the Muskingum-Cunge method
(routing_muskingum_cunge). The full set of functions
estimate all state variables at the end of each time step and

the fluxes that occurred over the time step. Finally, the raw
data is averaged or saved at the specified temporal resolution.
Outputs are stored in structured arrays based on variable
type, i.e.: state_maps, flux_maps, and time_series, where
each contains the key variables and the units of each.

The MOD-WET model is constructed in an attempt to
run reasonably eciently, but also in a way that makes for
easy student learning. The forward finite dierence
formulations for the energy mass/budgets are simple to
understand and allows for vectorization of the calculations.
This simply means that the change in any model state (e.g.
temperature) over the course of one time step can be
computed simultaneously for all pixels across the domain.
Alternatively, one could loop over pixels in the domain, but
loops tend to be very slow relative to vectorized calculations
and hence this saves a significant amount of time. The
primary drawback of these simple finite dierence schemes is
the need for a small time step (e.g. 15 min.), which tends to
be most necessary for the surface energy budget calculations.
More sophisticated finite dierence schemes (implicit/
iterative) may be more stable and therefore allow a larger
time step, but are generally less straightforward to vectorize.
The primary computational (CPU) expense is proportional to
the number of pixels being simulated (i.e. size of domain and
resolution of DEM) and the length of the simulation (number
of time steps required). Hence, short-duration simulations
with a small domain at coarse resolution will take the least
amount of CPU time. Long-duration simulations over a large
domain at high resolution will take the most CPU time. The
382

primary storage (RAM) expense comes from storing the


mapped arrays, which is therefore proportional to the size of
the domain and resolution of the DEM. The user can control
how often to output mapped results to control RAM
requirements. When designing simulations it is up to the user
to consider these control parameters to meet the practical (i.e.
computational) and instructional requirements.

383

S ECTION 4

Example MOD-WET
Model Applications

The model developed in the last two sections is done
generally so that it can be applied to basins of varying sizes
and characteristics. The key inputs to the model include a
DEM, soil parameters, meteorological data, etc. Here a simple
toy basin is used for qualitative demonstration of the model.
Figures 11.3 and 11.4 shows the DEM for the toy basin.
Geographically, its location (i.e. easting/northing coordinates)

F IGURE 11.4 Surface plot of the toy basin DEM showing the
varying slopes along the valley hillslope cross-section.

was chosen such that it is expected to occur in the Southern


Sierra Nevada in California (U.S.A.).

F IGURE 11.3 Toy basin DEM with elevation in meters.


The basin consists of a simple geometry with a
symmetric valley that has a stream that runs from East to
West with a specified bed slope. The hillslopes of the valley
have dierent slopes: a shallower slope in the lower part of the
valley (i.e. a flood-plain) with a steeper slope up to the ridge
lines. The basin outlet has an elevation around 3000 m with
the highest ridge line ranging up to 3050 m. The toy basin
configuration is chosen in order to highlight some of the key
points made earlier in the book including variability in
radiation, snowmelt, runo contributing areas, etc. Figure
11.5 shows the computed slope and aspect maps for the basin.
384

The slope is smallest in the valley floor and on the ridges,


with slopes on the hillsides ranging from 10-20 degrees. The
aspect map is quite simple with the pixels on the northern
slope eectively facing due south (i.e. 180 degrees) and the
southern slope eectively facing due north (i.e. 360 degrees).
These patterns will play a key role in explaining some of the
variability in hydrologic variables described below. The valley
floor and ridges have an aspect facing due West (i.e. 270
degrees) due to their shallow east-to-west slopes.

Based on the simple DEM, the topographic index
(Equation (11.2.1)) can be defined for the toy basin and is
shown in Figure 11.6. Based on the regular topography, the

F IGURE 11.5 Slope (top panel) and aspect (bottom panel)


maps for the toy basin.

F IGURE 11.6 Map of the topographic index for the toy basin.

385

topographic index varies in a predictable way with lowest


values along the ridge lines, monotonically increasing down
the hillslopes with largest values in the valley floor where the
stream pixels are expected to be located. The largest value is
at the basin outlet which by definition has the whole basin as
a contributing area.

Numerical experiments can be done to illustrate behavior
of the modeled system. Aside from the DEM, the other model
parameters used in all simulations described below are shown
in the test cases included in the MOD-WET package. First,
results are shown from a drainage experiment where the water

F IGURE 11.7 Profile of saturation deficit for the north-south

cross-section of the basin shown in Figure 11.3 at various times


over a 300 day simulation.

table is initially set to be at the surface throughout the basin


(i.e. SD(i,t0)=0) and the groundwater is allowed to drain via
baseflow. Such an experiment provides insight into the
equilibrium state the groundwater system would approach
under the case of no precipitation/infiltration/evaporation.
Figure 11.7 shows the distribution of the saturation deficit
over time through a north-south cross-section through the
middle of the basin. At the beginning of the simulation the
saturation deficit is zero throughout the cross-section.
Baseflow is largest at the beginning of the simulation causing
rapid drainage with predictable spatial patterns. Near the
stream channel in the middle of the basin, the saturation
deficit is smallest, while it is highest at the top of the
hillslopes. For the set of parameters used, after 300 days of
simulation the saturation deficit has dropped by about 0.15 m
at the stream channel and about 0.4 m at the top of the
hillslope. Over time the saturation deficit will continue to
drop, but more slowly as the baseflow is reduced.

Figure 11.8 shows the saturation deficit map at the end
of 300 days. The distribution shows the largest mode of
variability in the north-south direction, with additional
variability in the east-west direction along the stream channel.
As expected, the distribution of SD is highly correlated with
the soil-topographic index (Figure 11.6) via the application of
Equation (11.3.15). The pattern in saturation deficit provides
insight into the likely areas of saturation excess runo; namely
areas of shallow groundwater are most likely to generate
saturation excess runo during a storm that is large enough to
raise the groundwater table to the surface.
386

F IGURE 11.8 Saturation deficit map (in meters) over the toy
basin domain after 300 days of drainage.


A second experiment which is illustrated in this section
is a full-year simulation over the basin using realistic
meteorological forcing data. Figure 11.9 shows the dailyaveraged precipitation, incoming shortwave radiation, and air
temperature over a water year (WY; i.e. October 1st September 30th) corresponding to a meteorological station at
an elevation of 3300 m. It should be reiterated that the model
is run at a 15-minute time step to provide a stable solution for
the surface energy balance. The data is representative of the
Southern Sierra Nevada climate, showing the majority of

F IGURE 11.9 Key (daily-averaged) meteorological forcing

variables for full year simulation: precipitation (top panel), incoming shortwave radiation (middle panel), air temperature
(bottom panel). These forcing variables from a gage at a specified elevation and distributed spatially using topography.

387

precipitation occurring in the winter months, some of which


occurs as snow, and a strong seasonal cycle in incoming
shortwave radiation and air temperature. The model takes the
station data and distributes it across the basin using
disaggregation functions (introduced in earlier chapters) that
depend on topographic characteristics. For example, lower
elevations will receive higher air temperatures and north-

F IGURE 11.10 Basin-averaged time series of daily-averaged


surface temperature (top panel) and SWE (bottom panel).

facing slopes will receive less incident direct beam solar


radiation relative to higher elevations or south-facing slopes.

Basin-averaged results from the model are presented in
Figures 11.10-11.13. Figure 11.10 shows the daily-averaged
surface temperature and SWE over the annual cycle. Basinaveraged surface temperature decreases during the early
stages in concert with the decreasing energy input (i.e. Figure
11.9, middle panel) and reaches and stays below freezing
around 60 days after the beginning of the WY (i.e. around
December 1st). Snow begins to accumulate shortly thereafter.
Recall that the surface temperature represents soil
temperature when snow-free and corresponds to the snow
surface temperature otherwise. SWE increases until around
200 days after the beginning of the simulation and peaks at
about 0.6 m. Around the same time, the surface temperature
increases back to freezing which corresponds to the beginning
of snowmelt, where the snowmelt season lasts approximately
60 days. Once snow melts, surface temperature increases
rapidly.

Figure 11.11 (top panel) shows the daily-averaged
rootzone storage and saturation deficit. The rootzone storage
increases during the early part of the simulation in response
to rain events and decreases due to evaporation. The rootzone
storage reaches its capacity (approximately 0.05 m) around
the time that snow begins to accumulate. During that period,
there is no depletion of the rootzone moisture via soil
evaporation (i.e. all surface vapor loss is from snow
sublimation). So any infiltration due to snowmelt during that
period will quickly drain to the unsaturated zone keeping the
388


Figure 11.12 shows the daily-averaged surface energy
fluxes. Net radiation shows a decreasing trend early in the
simulation in response to decreasing incident shortwave
radiation and air temperature (which reduces incoming
longwave radiation) as shown in Figure 11.9. It is relatively
low during the snow season and then increases during the
summer. The daily-averaged sensible heat flux is generally
positive during the summer when the surface is warmer than
the overlying air and is generally negative when the surface is

F IGURE 11.11 Basin-averaged time series of daily-averaged


rootzone storage (top panel) and saturation deficit (bottom
panel).

storage fixed at its upper limit. Once snow begins to


disappear from some pixels, the rootzone storage is depleted
due to increased summer-time evaporation. The saturation
deficit (Figure 11.11, bottom panel) decreases intermittently
in response to recharge events. During such periods, some
pixels may saturate from below and yield saturation excess
runo. In between such events, saturation deficit increases due
to baseflow losses.

F IGURE 11.12 Basin-averaged time series of daily-averaged


net radiation, latent and sensible heat fluxes.

389


Figure 11.13 shows the daily-averaged runo terms
from the simulation. Saturation excess runo is shown to be
the largest contributor to runo for this basin and model
inputs. In fact, the infiltration excess runo is zero throughout
the simulation. This is because the specified saturated
hydraulic conductivity for the basin is larger than the
precipitation (or melt) rate experienced throughout the
simulation. Most of the saturation excess runo is in response
to intermittent melt events during the snow accumulation
season as well as to the sustained melt during Days 200-275 of
the Water Year.

So far only the basin-averaged results have been shown.

F IGURE 11.13 Basin-averaged time series of daily-averaged

saturation and infiltration excess runoff (top panel) and baseflow (bottom panel).

covered by snow during the winter months. The latent heat


flux is generally highest during summer and/or when rootzone
soil moisture is high. The latent heat flux is directly tied to
the moisture flux at the surface. When the surface is snowfree, the associated evaporation depletes the rootzone soil
moisture, while during the snow season it represents the
sublimation from the snowpack.

F IGURE 11.14 SWE map (in meters) over the toy basin domain on Day 260 of the Water Year.

390

The distributed model outputs these same states and fluxes at


all pixels in the basin (i.e. state/flux maps) throughout the
year. Example results are shown for SWE in Figure 11.14.
This state is chosen because, even for the simple toy basin
configuration, SWE is expected to show spatial variability.
Specifically, snowmelt should happen more quickly on south
facing slopes. Figure 11.14 shows the SWE map toward the
end of the melt season. From the figure it is clear that snow
has completely melted on the south-facing hillslope, while
snow is persistent on the north-facing hill slope. The
remaining SWE is also larger in the pixels with larger slopes,
which provides further reduction in incident shortwave
radiation due to local shading. To see the full range of

seasonal dynamics the daily-averaged SWE map is shown over


the whole year-long simulation in Movie 11.1. Keep in mind
that the precipitation is uniform over the basin so that
variability in SWE is either the result of precipitation being
classified dierently as snow vs. rain (only a function of
elevation-induced temperature dierences) or dierential melt.
The animation shows no SWE until around day of water year
(DOWY) 80 (mid-December). Around the time of peak basinaveraged SWE there is a distinct dierence between the northand south-facing slopes. The south-facing slopes have lower
SWE due to intermittent melt events during accumulation as
a result of increased radiative heating on those pixels. As
indicated in Figure 11.14, all of the snow has melted on the

F IGURE 11.15 Rootzone moisture storage map (in meters)


M OVIE 11.1 Animation of daily-average SWE (in meters) over

over the toy basin domain on Day 260 of the Water Year.

the toy basin for each day of the Water Year (DOWY).

391

south-facing slopes by DOWY 260, while the snow persists on


the north-facing slopes for a couple more weeks.

Figure 11.15 shows a similar example map of the
rootzone soil moisture storage on the same day as the SWE
map in Figure 11.14. The rootzone storage on the north-facing
hillslope is at its maximum value since the intermittent melt
water has replenished the rootzone and the snow cover is
preventing soil evaporation. For the pixels on the south-facing
slope, the rootzone storage is reduced since the melt flux is
zero and evaporation is depleting the near-surface storage
reservoir. The full range of seasonal dynamics in rootzone soil

M OVIE 11.2 Animation of daily-average rootzone soil moisture (in meters) over the toy basin for each day of the Water
Year (DOWY).

moisture is shown in Movie 11.2. While the initial condition is


the same over the whole basin, the north-facing slope
generally has a higher soil moisture prior to snow
accumulation (less radiative forcing and therefore less
evaporation). After snow cover persists (and there is some
intermittent melt), the rootzone storage fills up and is uniform
across the basin since no evaporation takes place from the soil
while snow is on the ground (any additional melt overflows
the rootzone into the unsaturated zone). It remains uniformly
distributed until the snow disappears and soil evaporation
begins. Since snow disappears from the south-facing hillslope
first, the rootzone soil moisture decreases on that portion of
the basin (around DOWY 260).

F IGURE 11.16 Saturation deficit map (in meters) over the toy
basin domain on Day 260 of the Water Year.

392

F IGURE 11.17 Saturation excess runoff map (in m3/hour)


over the toy basin domain on Day 260 of the Water Year.

M OVIE 11.3 Animation of daily-average saturation deficit (in



Figures 11.16 and 11.17 show the saturation deficit and
saturation excess runo maps on the same day as discussed
above. The SD map (Figure 11.16) shows that only the stream
channel pixels nearest the basin outlet are saturated at this
time. As expected, the saturation excess runo map (Figure
11.17) shows that only those pixels contribute runo on this
day in the model simulation. The area contributing to runo
is expected to expand and contract based on infiltration/
recharge dynamics over the course of a simulation. The
dynamics of the water table (saturation deficit) are shown in
Movie 11.3, which shows that SD is consistently smallest in
the main stream channel and nearest zero closest to the

meters) over the toy basin for each day of the Water Year
(DOWY).

outlet. The basin-wide dynamics are driven by the residual


between the drainage and baseflow fluxes (Equation
(11.2.12)). The bulk changes in SD are then distributed to the
individual pixels using the soil-topographic index via Equation
(11.2.13). As such, the pixels with highest soil-topographic
index tend to have lower SD values. This can lead to a water
table that intersects the surface and generates saturation
excess runo. In this case, it happens exclusively in the
handful of pixels upstream of the basin outlet. The full
seasonal dynamics are shown in Movie 11.4. In many instances
393

M OVIE 11.4 Animation of daily-average saturation excess runoff (in m3/h) over the toy basin for each day of water year
(DOWY).

this generates exfiltration due to the rising water table in


these pixels.

The set of results presented so far were intentionally
chosen for a synthetic basin to provide simple results for easy
interpretation. To provide an example for a more realistic case
we present analogous results for a real basin. Figure 11.18
shows a 90-m resolution DEM for the Tokopah watershed
located in the Southern Sierra Nevada (Sequoia National
Park) in California. The basin is primarily above tree-line and
has an area of 19 km2. For illustration, a simulation was
performed for Tokopah using the same meteorological forcing

F IGURE 11.18 Tokopah basin DEM (90-m resolution) in meters.

(which is representative of Tokopah elevations) and


parameters used in the toy basin simulation. The processes
and mechanisms are therefore expected to be similar in nature
to the toy basin example, so they are not explained in great
detail. The results are provided mainly to show more realistic
patterns in hydrologic response. The model was not
calibrated, but is used here solely for illustration.

The precipitation is applied uniformly over the basin and
is equivalent to that shown in Figure 11.9. As a result of the
seasonality of the precipitation and the high-elevation, it is
394

expected that Tokopah will be snow-dominated. The other


primary forcing is net radiation, which is illustrated over the
entire WY in Movie 11.5. Aside from the expected seasonal
variability and clouds, both of which are likely to have more
or less uniform impact on this small basin, there is significant
variability in the net radiation. This is primarily due to the
slope and aspect distribution throughout the basin, which
along with shading, directly impacts the incident shortwave
radiation.

Animations of the key model water budget states (SD,
SWE, and Srz) are shown in Movies 11.6-11.8. The saturation
deficit (Movie 11.6) is controlled primarily by the topographicsoil index, with low values in the stream channels and higher
values on the hillslopes. The dynamics are controlled by

M OVIE 11.5 Tokopah daily net radiation animation.

M OVIE 11.6 Tokopah daily saturation deficit animation.


drainage and baseflow, which for Tokopah is mainly driven by
snowmelt. The SWE field (Movie 11.7) is zero until around
DOWY 60 when it first starts to accumulate at high
elevations. Note that lower elevations are receiving
precipitation as well, but it is being classified as rainfall rather
than snowfall due to temperature. Near the peak of snow
accumulation, the SWE varies across the basin from
approximately 15 to 80 cm. During snow melt, the SWE
recedes from low to high elevation. By DOWY 300, the snow
has disappeared completely from the basin. The rootzone soil
moisture (Movie 11.8) is depleted by evaporation early in the
simulation prior to snowcover. During early snowcover,
intermittent rainfall/snowmelt fills the rootzone reservoir
where it remains full over most of the domain until snow
disappears and evaporation begins to again deplete the
395

reservoir. The patterns in soil moisture are primarily driven


by the impact of radiation patterns on evaporation.

M OVIE 11.7 Tokopah daily SWE animation.

M OVIE 11.8 Tokopah daily rootzone moisture animation.


The runo from Tokopah is primarily snowmelt driven.
For the precipitation and/or snowmelt rates and soil
parameters used, infiltration excess runo is zero throughout
the simulation. The saturation excess runo is shown in Movie
11.9. The results show that the saturation excess runo is zero
over most of the domain and is intermittently non-zero early
in the WY around the main channel during intermittent
rainfall/snowmelt events. The saturation excess runo is most
persistent in this same region during snowmelt. Note that
much of this predicted runo is exfiltration resulting from
upstream snowmelt infiltration that raises the water table.
The fact that most of the runo is the result of saturation

M OVIE 11.9 Tokopah daily saturation excess animation.


396

the MOD-WET model. It should be noted that it relies


heavily on previously presented MOD-WET codes and is
designed generally so that it could be used for more complex
basins. Sensitivity of hydrograph response to basin
characteristics is illustrated in the next section.

F IGURE 11.19 Tokopah outlet daily hydrograph.

excess runo is due to the choice of soil parameters in


combination with the rainfall/snowmelt rates. This is
discussed in more detail in the next section. The outlet
hydrograph for Tokopah is shown in Figure 11.19. This shows
several events early in the season (i.e. around DOWY 60, 100,
150), due to intermittent rainfall/melt events during the main
accumulation season. The bulk of the hydrograph is shown to
occur in the Spring when the primary snowmelt season occurs
before returning to near-zero baseflow late in the WY.

The results presented in this section are meant to be
illustrative of the types of simulations that can be done with
397

S ECTION 5

Sensitivity of Hydrograph
Response to Watershed
Characteristics

This section uses the MOD-WET model to illustrate how
hydrologic response (in terms of the runo hydrograph at the
outlet) is impacted by various basin characteristics. As a
starting point, the toy basin used in Section 4 is chosen as a
baseline (nominal) watershed (Figure 11.20; upper left
panel). The basin is a simple rectangular shape with the main
stream channel moving from east to west with the outlet at
the far left (west) center of the domain. The nominal domain
is 15 x 10 pixels on a DEM of 30 m resolution so that the
nominal basin area is 0.135 km2. The flow patterns are
therefore relatively straightforward with hillslope flow
primarily in the north-south direction and main stream
channel flow in the east-to-west direction. To assess the
impact of shape and size, three other basins are synthesized.
The longer basin (Figure 11.20; upper right panel) is the
same width as the nominal, but is four times longer (basin
area of 0.54 km2). The wider basin (Figure 11.20; lower left
panel) is the same length as the nominal, but is four times
wider (basin area of 0.54 km2). The larger basin (Figure
11.20; lower right panel) is both four times longer and wider
than the nominal (basin area of 2.16 km2).

F IGURE 11.20 Elevation maps (shown in meters above out-

let) of simple synthetic rectangular watersheds: nominal (upper left), longer (upper right), wider (lower left), and
larger (lower right). The outlet of each basin is at the far left/
center of each rectangular domain. White areas are outside of
each basin and all basins are plotted using the same horizontal
scale.


For simplicity, results from a Hortonian or infiltration
excess runo (Chapter 10, Section 3) event (occurring
uniformly over the basin) are used for illustration. Such events
provide a much more straightforward way to gain insight into
the sensitivity of the runo hydrograph to basin
398

characteristics. This is because all pixels in the watershed are


contributing flow. In the case of events dominated by
saturation excess runo (e.g. the results shown in Section 4),
the contributing streamflow generation areas are often isolated
to small regions around the main channel (Figure 11.17).
Hence, in such cases the size, shape, slope, roughness, etc. of
the basin may have limited or unexpected impact on the
runo hydrograph. One should be careful to note however
that, generally speaking, saturation excess runo is generally
more common and hence runo response is a complicated
function of both basin characteristics and climatology.

For infiltration excess runo to occur, soils generally
need to have very low saturated conductivity and/or
precipitation events need to be very intense (i.e. P >> Ks). To
construct such a scenario, the soil surface transmissivity
parameter (T0) was reduced significantly from that used in the
previous section (where the new value of Ks was set equal to
0.03 mm/h) and a single storm event of constant intensity of 3
mm/h was used. The storm event started six hours after the
beginning of the simulation and lasted for six hours (Figure
11.21). Additionally the saturation deficit (proportional to
average depth to water table) was increased such that all
stormflow runo was the result of infiltration excess runo.
The expected flow paths can be conceptualized as follows:
Each cross-section in the north-south direction consists of the
two hillslopes flowing to the main stream channel. One can
think of the hillslope cross-section as a unit that quickly
contributes flow to the stream and then flow will be translated
downstream. Hence the hillslopes near the outlet will quickly

F IGURE 11.21 Storm and resulting hydrographs for the basin


shapes and sizes shown in Figure 11.20.

contribute flow to the hydrograph while those upstream will


have a delayed contribution.

The nominal basin outlet hydrograph is shown in Figure
11.21 (black line). The streamflow hydrograph increases
almost immediately after the start of the storm and peaks
with a flow of approximately 0.005 m3/s before the end of the
storm. The hydrograph is relatively constant for a few hours
before receding and returning to near-zero baseflow by
approximately six hours after the end of the storm. The
399

hydrographs for the other basins are also shown in Figure


11.21.

Generally speaking, as the size of the watershed increases
it is expected that the integrated runo (i.e. area under the
curve) and the peak flow will increase. The time to peak and
shape of the hydrograph will depend on other basin
characteristics. For the wider basin (Figure 11.21; blue line),
the primary dierence is expected to be an increased area of
hillslope contributing flow to each unit of the main stream
reach. Hence the overall response time (from beginning to end
of runo response) between the wider and nominal basins are
very similar (since the basin length is the same). The peak
however is between 5 and 6 times that of the nominal case
and the flow peaks later, near the end of the storm. The
longer basin is the same area as the wider basin, but the
hydrograph shape (Figure 11.21; cyan line) diers
significantly. The longer basin hydrograph has a peak that is
about 2 times larger than the nominal, but peaks later,
plateaus for several hours and has a time base that is about
twice as long as the nominal and wider basin hydrographs.
The longer time base is simply a result of the larger travel
times associated with the longer basin. Finally, the larger
basin has a much larger peak that is shifted later in time
relative to the nominal (Figure 11.21; magenta line). The time
base for the larger basin is very similar to that of the longer
basin, illustrating that the length of the basin primarily
controls the overall duration of the stormflow response.

To illustrate how other basin characteristics impact
stormflow response, bed slope of the main channel and basin

F IGURE 11.22 Storm and resulting hydrographs for the

larger basin in Figure 11.18 under the nominal parameter set


and one involving a shallower bed slope for the main channel
and one with lower Manning roughness coefficient. The larger
basin case used here as a baseline is identical to that shown in
Figure 11.21.

roughness were also perturbed. Specifically cases where the


basin-averaged Manning roughness coecient was reduced
from 0.045 to 0.035 (less rough) and the bed slope of the
main channel was reduced by two orders of magnitude from
0.01 to 0.0001 (shallow bed) are shown in Figure 11.22. The
larger basin case was used for all three simulations since it
more easily illustrates dierences. For the case of a shallower
400

bed slope (Figure 11.22; black line) it is expected that


streamflow velocities will be smaller. As a result the peak flow
is reduced (by about 14%) and the flows are attenuated, with
low flows persisting beyond 1.5 days after the start of the
simulation. Note that this case did not change the slope of the
hillslope pixels (only the main stream channel) and hence flow
still makes its way to the channel at a comparable speed to
that of the baseline case. However once water is in the channel
it takes longer to reach the outlet. A reduction in the hillslope
gradients would be expected to further reduce the peak and
attenuate the hydrograph. For the case of a basin with
reduced roughness (Figure 11.22; blue line), the opposite
behavior is seen. The reduced roughness (which is applied over
the whole basin) speeds up flow in general. As such, the peak
is significantly increased (by about 25%).

The cases shown in this section are meant to be
illustrative and drive home some of the points from earlier in
the book in the context of model simulations. As stated
above, more realistic basins, with a dierent (or a mixture) of
runo generation mechanisms will generally have more
complicated responses to perturbations. Nevertheless a basic
understanding of how static parameters may control response
is useful. The reader is encouraged to use the MOD-WET
model to investigate basins of interest (real or synthetic) and
explore sensitivities and hypothesis tests to determine how
hydrologic response is impacted by climatology, basin
morphology, and natural or anthropogenic changes to static or
dynamic inputs.

401

S ECTION 6

MOD-WET Codes

Relevant new functions based on concepts introduced in this chapter include:
Distributed model driver function:

MOD_WET_model_driver.m
Specification of model input parameters:

MOD_WET_model_static_and_control_parameters.m
Initialization of model states and distributed parameters:

initialize_model.m
Topographic index:

topo_index.m

Computation of channel width throughout the network


based on contributing area and a power law:

channel_width.m
Wrapper to control the routing scheme:

routing_muskingum_cunge.m
Muskingum-Cunge routing scheme:

muskingum_cunge.m
Routing velocity check and adjustment of time step:

routing_celerity_check.m
Basin-averaged mass balance check at each time step:

mass_balance_check.m
Function to generate animations of mapped outputs from
a MOD-WET model simulation:

create_model_output_maps.m

Topographic-soil index:

topo_soil_index.m
TOPMODEL wrapper for evolving subsurface states and
fluxes:

TOPMODEL.m
Construction of flow network:

flow_network.m

402

S ECTION 7

Conceptual Questions
1. Provide a basic description of the topographic index and
what it represents in the TOPMODEL framework.
2. In the TOPMODEL framework, what is typically assumed
about the saturated soil hydraulic conductivity as a
function of depth?
3. In the TOPMODEL framework, what must the value of
saturation deficit (SD) be at a given pixel for it to generate
saturation excess runo?
4. In the TOPMODEL framework, what variables control the
baseflow from a given pixel?
5. Describe the meaning of the wide channel approximation
and why it is typically invoked?
6. What static channel parameters and dynamic model
variables are needed in the Muskingum-Cunge routing
equations?

403

S ECTION 8

Sample Problems

well your calibrated model does in a validation simulation for


periods outside of the calibration period.

Problem 11.1. Using the toy basin inputs provided with


MOD-WET, setup a model simulation. Run the simulation
over a full water year and explore the results (i.e. time series,
state maps, flux maps, etc.). Does the basin generate runo
primarily from saturation excess or infiltration excess
mechanisms?
Problem 11.2. Using your model constructed for Problem
11.1, perturb various static model parameters and explore
how basin runo and other variables change as a result. Use
the simulations to test your understanding and/or hypotheses
about hydrologic response to static basin characteristics.
Problem 11.3. Using your model constructed for Problem
11.1, perturb precipitation, air temperature, or other
meteorological variables to assess how hydrologic response
changes as a result of dynamic forcing inputs. To what extent
does precipitation have to increase to generate more
infiltration excess runo? In the case that your nominal
simulation generated snow accumulation, how does a change
in air temperature impact the snowmelt and resulting
hydrograph?
Problem 11.4. Use real data to construct a watershed model
for a basin of interest. Explore model calibration as a means
for fitting model predictions to measured runo. Explore how
404

Chapter 12

MATLAB
Basics


It is generally assumed that the reader has had a course
in basic MATLAB or another programming language. The idea
of this section is not to provide a full primer or reference, but
highlight some key points to help in using MOD-WET
functions and provide pointers to more general references.
First, it should be noted that MATLAB is capable of much
more than covered here, has an extremely useful built-in help
system, and Mathworks provides their own rather extensive
primer for beginners located at: http://www.mathworks.com/
help/pdf_doc/matlab/getstart.pdf. You should consult these
resources and keep them handy for reference. There are also
many textbooks designed to provide more thorough
introductions to MATLAB (e.g. Chapman, 2007; Pratap, 2009).
The best way to find a solution to a problem you have is to
find an example of how such things are done via the help
command, using the primer, or other online resources. You
should also be aware that none of the applications used in this
book need to be solved using MATLAB, it is simply chosen as a
numerical tool since it is readily available and is relatively
easy to learn.

operations. These built-in functions are m-files, which are


simply text files containing instructions that are stored
somewhere on the MATLAB search path. The search path is a
list of directories to look through for a specified m-file. For
example, when taking the average of a vector a, i.e. mean(a),
MATLAB looks through the specified path for the file mean.m
and provides it the vector a as an input and returns the mean
as an output, as obtained via the set of instructions to
perform supplied in the mean.m file. For all of the built-in
MATLAB functions and their associated toolboxes, the path to
those functions are loaded by default. Additionally, the
current directory you are working in is always on the path and
is the directory that is searched first. Hence in calling a
function, unless new directories are added to the path, the
function must exist either in the current directory or on the
default MATLAB path. It is useful to keep this in mind because
as you develop your own m-files or use those developed by
others, you will need to add the location of those functions to
your path. Along these lines it is useful to keep your functions
in an organized location that can be added easily to the path.


In its simplest form, MATLAB can be used as a calculator
to perform basic arithmetic operations on values stored in
variables, vectors, or multi-dimensional arrays. As a first step,
you need to be comfortable with how variables are stored in
arrays, array operations, indexing in arrays, etc. (see the
Language Fundamentals chapter in the MATLAB primer).


The idea of specifying a search path is particularly important when using the MOD-WET functions. Paths can be
manually added to MATLAB, but the approach described below automates this process with a simple line that you can
add to your own code. MOD-WET has several subdirectories.
Assume that you have the following directory structure:

The MATLAB search path and m-files



Intrinsic in MATLAB are many built-in functions that
perform more complicated operations built-up from simpler

/Users/username/CEE150/MOD_WET/chapter1
/Users/username/CEE150/MOD_WET/chapter2
/Users/username/CEE150/MOD_WET/chapter3
406

/Users/username/CEE150/MOD_WET/chapter5
/Users/username/CEE150/MOD_WET/chapter6
/Users/username/CEE150/MOD_WET/chapter7
/Users/username/CEE150/MOD_WET/chapter8
/Users/username/CEE150/MOD_WET/chapter11

where the individual MOD-WET functions are within each of


the directories listed above. To add all subdirectories within a
specified directory to the search path one could type:
addpath(genpath('/Users/username/CEE150/ ...
MOD_WET')).

The addition of the genpath command will allow MATLAB to


access all functions in the MOD-WET (i.e. Chapters 1 11),
since directories chapter1 chapter11 are all contained in
the MOD_WET parent directory. Adding the path to the MODWET directory or your other functions at the top of all of
your own scripts should help you avoid any problems with
MATLABs search path. The use of the genpath command is
highly recommended because you will not have to remember
which toolbox function is associated with a specific chapter or
directory.
MATLAB scripts vs. functions

In performing any analysis, it is usually required to make
use of multiple commands which can be tailored to the desired
task/s. This typically involves usage of basic programming
written in the form of scripts or functions (see the
Programming chapter in the MATLAB primer), which allow
for repeated usage and to fully document the process. Both
scripts and functions are written as m-files, with the

distinction coming via the usage of the function command at


the beginning of the file.

Scripts operate as if the commands in the file were being
typed into the command window. As such it does not have
inputs or outputs but operates on variables that either
already exist in the workspace or are created within the
script. At the end of the script, variables that were created
within it are defined in the workspace. Functions can be called
from within scripts.

Functions dier from scripts in an important way;
namely they require inputs and send back outputs. In the case
of functions, all variables are local to the function and are not
explicitly saved in the workspace when complete -- only
inputs/outputs will exist in the workspace after a function
call. A function m-file begins with a function definition line
(without this line you would have a regular script m-file), i.e.:
function [OUTPUT] = function_name (INPUT)

where INPUT is a variable provided to the function, and


OUTPUT is a variable sent back from the function, and
function_name is the name of the function. While not
required, function_name is most often also used in the
filename, i.e. function_name.m. Both the inputs and outputs
can consist of multiple variables.

The primary reason for using functions over scripts is
that they are inherently modular where only inputs/outputs
need to be dealt with. For this reason, the MOD-WET
consists entirely of functions. The value of modularity is
407

particularly apparent in the development of the MOD-WET


model, which is built up from existing MOD-WET functions.
In addition to the MOD-WET functions and those you
develop on your own, there are many community-based
functions that are provided on the MATLAB file exchange.
Visualization in MATLAB:

Visualizing imported data or results of computations are
often a crucial part of understanding and communicating your
work. MATLAB provides many options for visualizing data (see
the Graphics chapter in the MATLAB primer). Depending on
the type of data to be plotted, dierent graphics options exist
(figures can be created and controlled via the figure, clf,
close, and related commands). It is important to choose a
type of graphical representation that is well-suited to the
characteristics you are trying to communicate. When plotting
time series data or one variable vs. another, an x-y line (or
scatter) plot using the plot command is generally most
useful. Annotation of plots is crucial, so you should become
familiar with commands to control labeling and other features
of the plot (i.e. xlabel, ylabel, grid, title, legend,
colorbar, set, etc.). For time series, where date/time labels
are desired, the use of the xtick, ytick, and datetick
commands may be useful. In many cases it might also be
useful to have multiple graphics on the same figure, which can
be done via the subplot command.

In many hydrologic examples (e.g. topography) it is
useful to visualize data as mapped results (i.e. when a variable
depends on geographic x/y coordinates). The most common
functions useful for mapping are: imagesc, surf, contour,

etc. For example, in plotting a DEM one could do the


following (where comments are added for clarity):
% generate 2D map given x and y coordinates (vectors) and map (2D array) of elevations, i.e. dsm
imagesc(x,y,dsm);
% set the correct orientation for the y-axis (otherwise MATLAB flips it)
set(gca,'YDir','normal');
% impose the proper aspect ratio (i.e. each pixel
should be square not rectangular based on the implicit geographic extent if each pixel is 30m x 30m
for example)
axis image;
% add colorbar
colorbar;
% keep the map on and allows for overlaying plots
hold on


MATLAB provides several colorbar options, with jet
being the default. Other colorbars, or ones you design yourself
may be more useful in other circumstances. For example, the
hsv colormap is useful for plotting maps of aspect or wind
direction (circular/angular directions) because angles of both
0 and 360 degrees should be represented by the same color
since they are actually the same angle. High and low values
are represented with the same color in this colormap which
can be invoked via: colormap(hsv).

Finally, invariably you will need to include graphical
figures into homework assignments or lab reports being
written in a word processing program. There are multiple
ways to save figures. Saving can be done from within a script
or function using the saveas command, e.g.
408

saveas(gcf,'figure_name.png','png')

or manually via: File > Export (in the MATLAB figure


window). In saving or exporting figures it is very important
that you do so in a way that makes them readable and clear
in the document you are putting it into. You do not want all
of your hard work in generating the figure to go to waste if it
is unreadable in the word processing file! In particular you
should control the font size, line thickness, and graphical
resolution on the export. Note that the default font and line
sizes and rendering do not always make for readable figures
when pasted into a document. These and other figure
characteristics can be done from within your script (using set
or print commands) or via the Export Setup, which can be
customized for your needs.

409

Chapter 13

References

Adler, R.F. et al., 2003: The Version-2 Global Precipitation


Climatology Project (GPCP) Monthly Precipitation Analysis
(1979Present). Journal of Hydrometeorology, 4, 11471167.
Ambroise, B. et al., 1996: Towards a generalisation of the
TOPMODEL concepts: Topographic indices of hydrologic
similarity, Water Resources Research, 32, 2135-2145.
Armstrong, R.L. and E. Brun, 2010: Snow and climate:
Physical processes, surface energy exchange, and modeling,
Cambridge University Press.

Brunt, 1932: Notes on radiation in the atmosphere, Quarterly


Journal of the Royal Meteorological Society, 58, 389-420.
Brutsaert, W. H., 1975: On a derivable formula for long-wave
radiation from clear skies, Water Resources Research, 11,
742744.
Brutsaert, W. H., 1982: Evaporation into the Atmosphere:
Theory, history, and applications, Environmental Fluid
Mechanics (Book 1), Springer Publishing.

Arya, P.S., 2001: Introduction to Micrometeorology,


International Geophysics Series, Volume 79, Academic Press.

Businger, J.A. et al., 1971: Flux-profile relationships in the


atmospheric surface layer, Journal of Atmospheric Science,
28, 181-189.

Bear, J., 2007: Hydraulics of groundwater, Dover Books on


Engineering, Dover Publications.

Carle, D., 2004: Introduction to Water in California,


University of California Press.

Beven, K. and M.J. Kirby, 1979: A physically based variable


contributing area model of catchment hydrology, Hydrologic
Science Bulletin, 24, 43-69.

Chapman, Stephen J., 2007: Matlab Programming for


Engineers. Thompson, 4th ed., p. 592.

Beven, K., 1997: TOPMODEL: A Critique, Hydrological


Processes, 11, 1069-1085.
Beven, K, 2012: Rainfall-Runo Modeling: The Primer,
Wiley-Blackwell.
Bras, R.L., 1990: Hydrology: An Introduction to Hydrologic
Science, Addison-Wesley Publishing Co., Inc.
Brooks, R.H. and A.T. Corey, 1966: Properties of porous
media aecting fluid flow, Journal of Irrigation and
Drainage Division, ASCE, IR2, 61-88.

Clapp, R.B. and G.M. Hornberger, 1978: Empirical equations


for some soil hydraulic properties, Water Resources Research,
14, 601-604.
Cunge, J.A., 1969: On the subject of a flow propagation
method (Muskingum method), Journal of Hydraulics
Research, International Association of Hydraulic Research,
7(2), 205-230.
Darcy, H., 1856: Les Fontaines Publiques de la Ville de Dijon,
Dalmont, Paris.
411

Dingman, S.L., 2008: Physical Hydrology, Waveland Press.


Dunne, T., 1975: Field studies of hillslope flow processes, in:
M.J. Kirkby, ed. Hillslope Hydrology, Wiley Interscience,
227-293.
Dunne, T. and L.B. Leopold, 1978: Water in Environmental
Planning, W.H. Freeman and Co.
Eagleson, P.S., 1970: Dynamic Hydrology, McGraw-Hill Book
Co.
Filippa G. et al., 2010: Major element chemistry in inner
alpine snowpacks (Aosta Valley Region, NW Italy), Cold
Regions Science and Technology, doi:10.1016/j.coldregions.
2010.07.005.
Georgakakos, K.P. and R. Bras, 1984: A hydrologically useful
station precipitation model 1. Formulation, Water Resources
Research, 20, 1585-1596.
Graham, D.N. and M. B. Butts 2005: Flexible, integrated
watershed modelling with MIKE SHE, In Watershed Models,
Eds. V.P. Singh & D.K. Frevert, 245-272, CRC Press.
Green W.H. and G. Ampt, 1911: Studies of soil physics. Part
I: The flow of air and water through soils, Journal of
Agricultural Science, 4, 1-24.
Green, D., 2007: Managing water: Avoiding crisis in
California, University of California Press.

Gupta, S.K., 2011: Modern Hydrology and Sustainable Water


Development, Wiley-Blackwell Publishing Co.
Horton, R.E., 1933: The role of infiltration in the hydrologic
cycle, Transactions of the American Geophysical Union, 14,
446-460.
Hundley, N., 2001: The Great Thirst: Californians and water:
A History, University of California Press.
Idso, S.B., 1981: A set of equations for full spectrum and 8- to
14- micron and 10.5-12.5 micron thermal radiation from
cloudless skies, Water Resources Research, 17(2), 295-304.
Ivanov, V.Y., Vivoni, E.R., Bras, R. L. and Entekhabi, D.
2004: Catchment Hydrologic Response with a Fullydistributed Triangulated Irregular Network Model, Water
Resources Research, 40(11), W11102,
10.1029/2004WR003218.
Kustas, W., Rango, A. and Uijlenhoet, R., 1994: A simple
energy budget algorithm for the snowmelt runo model,
Water Resources Research, 30(5): doi: 10.1029/94WR00152.
Liou, K.-N., 1992: Radiation and cloud processes in the
atmosphere, Oxford Monographs on Geology and Geophysics
(Book 20), Oxford University Press.
Liou, K.-N., 2002: An introduction to Atmospheric Radiation,
International Geophysics (Book 84), Academic Press, Second
Edition.

412

Male, D.H. and D.M. Gray, 1981: Handbook of snow:


Principles, processes, management and use, The Blackburn
Press.

Penman, H.L., 1948: Natural evaporation from open water,


bare soil, and grass, Proceedings of the Royal Society
(London) A, 193, 120-145.

Marshall, J.S. and W.M. Palmer, 1948: The distribution of


raindrops with size, Journal of Meteorology, 5, 165-166.

Philip, J.R., 1960: General method of exact solution of the


concentration dependent diusion equation, Australian
Journal of Physics, 13(1), 1-12.

Marshall, J. and R.A. Plumb, 2007: Atmosphere, Ocean and


Climate Dynamics: An Introductory Text, Academic Press.
Mays, L.W., 2005: Water Resources Engineering, John Wiley
and Sons, Inc.
Muller, M.D. and D. Scherer, 2005: A grid and subgrid-scale
radiation parameterization of topographic eects for
mesoscale weather forecast models, Monthly Weather Review,
133, 1431-1442.
National Weather Service (NWS), 2002: Conceptualization of
the Sacramento Soil Moisture Accounting Model, http://
www.nws.noaa.gov/oh/hrl/nwsrfs/users_manual/part2/
_pdf/23sacsma.pdf.
Noilhan, J. and S. Planton, 1989: A simple parameterization
of land surface processes for meteorological models, Monthly
Weather Review, 117, 536-549.
Parlange, M.B. and J.W. Hopmans, 1999: Vadose Zone
Hydrology: Cutting Across Disciplines, Oxford University
Press.

Pratap, R., 2009: Getting Started With Matlab. A Quick


Introduction for Scientists and Engineers, Oxford University
Press.
Pruppacher, H.R. and J.D. Klett, 2010: Microphysics of
Clouds and Precipitation, Atmospheric and Oceanographic
Sciences Library (Book 18), Springer Publishing.
Reisner, M., 1993: Cadillac Desert, Penguin Publishing.
Richards, L.A., 1931: Capillary conduction of liquids through
porous mediums, Physics, 1(5), 318333.
Satterlund, D.R., 1979: An improved equation for estimating
long-wave radiation from the atmosphere, Water Resources
Research, 15, 1649-1650.
Sivapalan, M. et al., 1987: On hydrologic similarity. 2. A
scaled model of storm runo production, Water Resources
Research, 23, 2266-2278.
Sturm et al., 1995: A seasonal snow cover classification system
for local to global applications, Journal of Climate, 8,
1261-1283.
413

Takeuchi, K. et al., 1999: Introduction of block-wise use of


TOPMODEL and Muskingum-Cunge method for the
hydroenvironmental simulation of a large ungauged basin,
Hydrological Sciences Journal, 44(4), 633-646.
U.S. Army Corps of Engineers (USACE), 1956: Snow
hydrology, Washington D.C., U.S. Department of Commerce
Oce of Technical Services, PB 151660.
U.S. Army Corps of Engineers (USACE), 2010: Hydrologic
Modeling System, User Manual, http://
www.hec.usace.army.mil/software/hec-hms/documentation/
HEC-HMS_Users_Manual_3.5.pdf.
U.S. National Research Council (NRC), 1991: Opportunities
in the Hydrologic Sciences, Committee on Opportunities in
the Hydrologic Sciences, Water Science and Technology
Board, National Academy Press, Washington, D.C.
World Meteorological Organization (WMO), 2010: Manual on
stream gauging, Volume I - Fieldwork, WMO-No. 1044,
http://www.nws.noaa.gov/os/water/RFC_support/
resources/WMO%20Manual%20on%20Stream%20Gauging
%20-%20Field%20Work%20%201044_Vol_I_en.pdf.
Yau, M.K. and R.R. Rogers, 1984: A short course in cloud
physics, International Series in Natural Philosophy,
Butterworth-Heinemann Publishers.

414

Potrebbero piacerti anche