You are on page 1of 94

4.

Management of Groundwater Systems


Daene C. McKinney and Andre Savitsky
(Chapter 4 of Analysis of Water Resource Systems)
Revised January 19, 2014

Contents
Section Page
4. Management of Groundwater Systems ....................................................................................... 1  
4.1. BASIC CONCEPTS AND PRINCIPLES ..................................................................................................................... 3  
4.1.1 Moisture Distribution in the Subsurface .................................................................................................... 3  
4.1.2. Flow of Immiscible Fluids ......................................................................................................................... 4  
4.1.3 Continuum Approach to Porous Media ...................................................................................................... 4  
4.1.4 Porosity ....................................................................................................................................................... 6  
4.1.5. Interfacial Tension and Wettability ........................................................................................................... 7  
4.1.6. Capillary Pressure ..................................................................................................................................... 8  
4.1.7. Drainage .................................................................................................................................................. 10  
4.1.8. Energy Relationships in Fluids ................................................................................................................ 10  
4.1.8.1. Hydraulic Head.................................................................................................................................................... 10  
4.1.8.2. Piezometric Head ................................................................................................................................................ 11  
4.1.8.3. Piezometric Head in Unsaturated Flow ............................................................................................................... 12  
4.1.9. Soil Water Characteristic Curves ............................................................................................................ 12  
4.1.10 Aquifer Types .......................................................................................................................................... 13  
4.1.11 Aquifer Properties .................................................................................................................................. 14  
4.1.11.1 Hydraulic Conductivity ....................................................................................................................................... 14  
4.1.11.2 Storativity ........................................................................................................................................................... 15  
4.2. DARCY’S LAW .................................................................................................................................................. 16  
4.3. HETEROGENEITY AND ANISOTROPY ................................................................................................................. 20  
4.3.1. General .................................................................................................................................................... 20  
4.3.2 Equivalent conductivity in layered aquifer formations ............................................................................ 21  
4.3.3 Anisotropic Porous Media ........................................................................................................................ 23  
4.4. POTENTIOMETRIC SURFACES ............................................................................................................................. 23  
4.5. GENERAL GROUNDWATER FLOW EQUATIONS .................................................................................................. 25  
4.6. HORIZONTAL FLOW IN CONFINED AQUIFERS ................................................................................................... 28  
4.6.1 Introduction .............................................................................................................................................. 28  
4.6.2 Confined Groundwater Flow Equations ................................................................................................... 28  
4.6.3. Examples .................................................................................................................................................. 30  
4.6.3.1. Flow in a confined aquifer................................................................................................................................... 30  
4.6.3.2. Flow in a leaky confined aquifer ......................................................................................................................... 31  
4.6.3.3. Steady Radial Flow in a Confined Aquifer ......................................................................................................... 33  
4.6.3.4. Theim Equation Starting from Darcy’s Law ....................................................................................................... 36  
4.6.3.5. Drawdown in two monitoring wells due to pumping in a confined aquifer......................................................... 37  
4.6.3.6. Steady Radial Flow in a Leaky (Semi-Confined) Aquifer .................................................................................. 37  
4.6.3.7. Transient Radial Flow in a Confined Aquifer ..................................................................................................... 39  
4.6.3.8. Theis’ method and Jacobs’ approximation .......................................................................................................... 41  
4.6.4 Some Potential Theory .............................................................................................................................. 41  
4.6.4.1 Introduction .......................................................................................................................................................... 41  
4.6.4.2 Potentials for Different Aquifer Types ................................................................................................................. 46  
4.6.4.3. Fundamental Singularities ................................................................................................................................... 47  
4.6.4.4. Superposition ....................................................................................................................................................... 50  
4.6.4.5. Head distribution from Two Pumping Wells in a Confined Aquifer .................................................................. 51  
4.6.4.6. Capture Zone of a Well in Uniform Flow ........................................................................................................... 51  
4.6.4.7. Steady Drawdown in a Confined Aquifer ........................................................................................................... 53  
4.6.4.8. Transient Radial Flow in a Confined Aquifer ..................................................................................................... 55  
4.6.4.9. Minimum drawdown in a confined aquifer ......................................................................................................... 57  

1
4.7. HORIZONTAL FLOW IN UNCONFINED AQUIFERS............................................................................................... 59  
4.7.1 Introduction .............................................................................................................................................. 59  
4.7.2 Unconfined Groundwater Flow Equations ............................................................................................... 62  
4.7.3 Examples ................................................................................................................................................... 65  
4.7.3.1. Flow in an Unconfined Aquifer........................................................................................................................... 65  
4.7.3.2. Drains .................................................................................................................................................................. 67  
4.3.3.3 Steady Radial Flow in a Unconfined Aquifer ...................................................................................................... 68  
4.8 MODELING GROUNDWATER SYSTEMS ............................................................................................................... 70  
4.8.1 Methods for Solving Groundwater Problems ........................................................................................... 70  
4.8.2 Classification of Partial Differential Equations ....................................................................................... 71  
4.8.3 Numerical Differentiation ......................................................................................................................... 72  
4.8.4 Grids and Discretization .......................................................................................................................... 74  
4.8.4.1. Introduction .......................................................................................................................................................... 74  
4.8.4.2. Consistency .......................................................................................................................................................... 75  
4.8.4.3. Order of Approximation ....................................................................................................................................... 75  
4.8.5 One-Dimensional Flow ............................................................................................................................. 75  
4.8.5.1 Steady-state flow .................................................................................................................................................. 75  
4.8.5.2. Simulation of Steady Flow in a 1-D Homogeneous System ................................................................................ 76  
4.8.5.3. Simulation of Transient Flow in a 1-D Homogeneous System ............................................................................ 76  
4.8.5.4. Optimization of Steady Flow in a 1-D Homogeneous System ............................................................................ 82  
4.8.6 Two-Dimensional Flow ............................................................................................................................ 83  
4.8.6.1 Steady-State flow ................................................................................................................................................. 83  
4.8.6.2 Boundary Conditions ............................................................................................................................................ 85  
4.8.6.3. Optimization of Steady Flow in 2-D Homogeneous System ............................................................................... 87  
4.8.7 Heterogeneous Systems ............................................................................................................................ 88  
4.8.7.1. Introduction .......................................................................................................................................................... 88  
4.8.7.2. Optimization of Steady Flow in a 2-D Heterogeneous System........................................................................... 90  
4.8.8 Transient Problems ................................................................................................................................... 91  
4.8.8.1. Introduction ......................................................................................................................................................... 91  
4.8.8.2. Optimization of Transient Flow in 2-D Homogeneous System .......................................................................... 92  
4.9 MULTIPHASE SYSTEMS ...................................................................................................................................... 94  

2
4.1. Basic Concepts and Principles
4.1.1 Moisture Distribution in the Subsurface
Groundwater consists of all waters found beneath the ground surface. Groundwater occupies the
pore space (i.e., voids or interstices) of subsurface formations. The pore space is the space in a
formation not occupied by solid matter. Porous media consist of numerous pores of small size.
Large openings such as fractures or caverns are typical in some regions, such as the Edwards
aquifer in Central Texas. The voids of a porous medium contain fluids, typically water and air,
sometimes contaminants either dissolved in the aqueous phase or as separate phases (e.g.,
nonaqueous phase liquids--NAPLs). The connected pores act as conduits for the flow of fluids.

Voids are the receptacles that hold the water (Meinzer, 1923). The kinds of rocks present in a
formation and their (1) number, (2) size, (3) and arrangement of the voids affect the properties of
the water storage and flow through a formation. Interstices are generally irregular in shape.
Different irregularities are characteristic of different of different kinds of rocks. These
differences result mainly form the differences in the minerals making up the rocks and the
geologic processes experienced by them.

Different zones of saturation (% of void space occupied by water) exist in the subsurface. The
zone of aeration is the region where some pores contain water and air. The soil water zone is the
first level encountered below the surface and depends on surface conditions. In this zone, water
moves down (up) during infiltration (evaporation). Field capacity is the content of water in the
soil remaining after gravity drainage, that is, it is held in place by capillary action. The wilting
point is the water content remaining after gravity drainage and evapotranspiration. The vadose
zone lies below the soil water zone and is where water is held in place by capillary forces but is
relatively unaffected by overlying vegetation. The saturation is at or near field capacity in this
zone except during infiltration. The capillary zone lies just above the water table and is
completely saturated at the base and near field capacity at the top. In this zone, water is pulled
up from the water table by capillary forces.

•  Different(zones(
–  depend%on%%%of%pore%
space%filled%with%water%
•  Unsaturated(Zone(
–  Water%held%by%capillary%
forces,%water%content%
near%field%capacity%
except%during%
infiltra:on%
•  Soil(zone(
–  Water%moves%down%
(up)%during%infiltra:on%
(evapora:on)%
•  Capillary(fringe(
–  Saturated%ar%base%
–  Field%capacity%at%top%
•  Saturated(Zone(
–  Fully%saturated%pores%

Figure 4.1. Distribution of Water in the Subsurface (adapted from Charbeneau 2000)

3
The saturated zone is the region of the subsurface where water fills all of the void space. In this
zone, the porosity provides a measure of the water per unit volume. Water can be removed from
this zone by either drainage or pumping. The saturated zone is bounded above by the water table
or phreatic surface and below by impermeable or semi-pervious layers.

4.1.2. Flow of Immiscible Fluids

Consider the flow of two or more fluids through a porous medium. We can have: (1) miscible
displacement where the fluids are completely soluble in each other, the interfacial tension
between the fluids is zero, the fluids dissolve in each other, and a distinct fluid-fluid interface
does not exist; or (2) immiscible displacement where there is simultaneous flow of immiscible
fluids or phases in the porous medium. The interfacial tension between the fluids is not aero,
distinct fluid-fluid interfaces exist and separate the phases in each pore. Unsaturated flow is the
flow of water and water vapor through a porous medium where the void space is partially filled
with air. This represents the flow of two immiscible fluids (water and air), except that the air is
practically immobile.

When void space in a porous medium is filled by two or more immiscible fluids, the saturation
with respect to a particular fluid is the fraction of the void space filled by that fluid within a pore
around a point
!! !"#$%&  !"  !"#$%  !
𝑆! = !!
= !"#$%&  !"  !"#$%
, ! 𝑆! =1 (4.1.1)

Water content is defined as


! !"#$%&  !"  !"#$%
𝜃! = !! = !"#$  !"#$%&  
, 0 ≤ 𝜃! ≤ 𝜙, 𝜃! = 𝜙𝑆!     (4.1.2)  
!

Water Saturation is defined as


! !"#$%&  !"  !"#$%
𝑆! = !! = !"#$%&  !"  !"#$% , 0 ≤ 𝑆! ≤ 1     (4.1.3)  
!

4.1.3 Continuum Approach to Porous Media

In fluid mechanics, variables such as pressure, density and piezometric head apply to fluid
elements that are large relative to molecular dimensions, but small relative to the size of the flow
problem under consideration. The Navier-Stokes equations describe the fluid motion and they
are valid for flow in the pore spaces of a porous medium. However, two problems arise: (1) it is
impossible to describe the complicated geometry of the porous medium through which the flow
takes place in order to solve the Navier-Stokes equations at the pore scale; and (2) it is extremely
difficult to measure flow at the pore scale.

One solution to this problem is to adopt the Representative Elementary Volume (REV) approach

4
in which we consider average values of the variables so that points become associated with some
volume of the porous medium (the REV) over which the averaging takes place. The REV must
be sufficiently large to contain enough pores to define the average value of the variable in the
fluid phase and to ensure that the pore-to-pore fluctuations are smoothed out. On the other hand,
the REV must be sufficiently small that larger scale heterogeneities do not get averaged out
(layering, etc.). The REV is the smallest volume for which gradual or smooth changes in the
average values of the variables are observed (Bear, 1972).

Figure 4.7 Types of rocks making up typical porous media and aquifers.
Source: Freeze and Cherry, 1979 after Meinzer, 1923

Consider porosity for example. If we consider porosity as being defined at every point in space,
then

!
𝜙=! !(!)
𝜒 𝑥 − 𝑥 ! 𝑑𝑥′ (4.1.4)

where V is the REV volume, and

0  𝑖𝑓  𝑥  𝑖𝑠  𝑖𝑛  𝑠𝑜𝑙𝑖𝑑  𝑚𝑎𝑡𝑒𝑟𝑖𝑎𝑙


𝜒 𝑥 = (4.1.5)
1 𝑖𝑓  𝑥  𝑖𝑠  𝑖𝑛  𝑓𝑙𝑢𝑖𝑑

We can plot this as shown in Figure 4.8.

5
Figure 4.8. Fluctuation of porosity with size of REV.
Source: Bear 1972

In a homogeneous porous medium, the REV will exhibit the same porosity everywhere.
Whereas, in a heterogeneous porous medium, the REV will show smooth variations of porosity
as x is changed. Once the REV has been established, it is used to derive macroscopic
(continuum) properties of the flow by averaging the microscopic (pore scale) values of variables.

4.1.4 Porosity
Consider a porous medium comprised of two regions, solid and voids (s,i). Each region occupies
a volume and Vs + Vi = V. Porosity is a property of the interstices of the porous medium. It is
the percent of total pore volume occupied by void space or the percent of total pore volume not
occupied by solid material

Volume voids Vi V −Vi
φ= = = (4.1.6)
Volume total V V

One can also show that in terms of the densities of the particles (ρm) and the bulk material (ρb)

ρ m − ρ di ρ
φ= = 1− d (4.1.7)
ρm ρm

In many cases, especially mass transport of contaminants in aquifers, only the interconnected
pores through which fluid can actually flow are of interest. In this case, the property of interest
is the effective porosity, defined as the volume of connected pores divided by the total volume.

The porosity of a porous medium depends on many factors, the most important of which are the
quantity of fine particles in the medium, the packing, shape, arrangement and size distribution of
the grains, and any possible cementation of the particles of the porous matrix (Meinzer, 1923).

6
Table 4.1. Typical Porosity Values of Natural Sedimentary Materials.
Sedimentary Material Porosity (%)
Peat Soil 60-80
Soils 50-60
Clay 45-55
Silt 40-50
Med. to Coarse Sand 35-40
Uniform Sand 30-40
Fine to Med Sand 30-35
Gravel 30-40
Gravel and Sand 30-35
Sandstone 10-20
Shale 1-10
Limestone 1-10
Source: R.A. Freeze and J. Cherry, Groundwater, Prentice-Hall Publishers, Englewood Cliffs, 1979]

4.1.5. Interfacial Tension and Wettability


Liquids in contact with another substance experience an interfacial energy arising from the
inward attraction of the molecules of each phase and those at the interface. Interfacial energy is
manifested as an interfacial tension σik defined as the quantity of energy required to separate a
surface of unit area of the substances i and k. Surface tension is the interfacial tension between a
substance and its own vapor. Two immiscible fluids in contact with an impervious surface form
an angle b measured from 0 to 180 degrees in the denser fluid.

σgl
gas
liquid
σsg β#

solid σsl
Figure 4.2. Surface tension between two fluids and a solid.

Equilibrium requires that

𝜎!" 𝑐𝑜𝑠 𝛽 = 𝜎!" −   𝜎!"   (4.1.8)  

or
!!" !  !!"
𝑐𝑜𝑠 𝛽 = !!"
    (4.1.9)  

and this is Young’s Equation. When β < 90o the liquid is wetting the solid (gas is non-wetting)
and when β > 90o the liquid is non-wetting for the solid surface (gas is wetting). For

7
complicated geometries, a mean curvature is used to arrive at a more general expression.

air
air

solid
solid
β"
Hg
β"
water

(a) Mercury nonwetting solid (b) Water wetting solid

Figure 4.3. Nonwetting (a) and wetting (b) relations between fluids and solid surfaces

4.1.6. Capillary Pressure


Capillary pressure is inversely proportional to curvature and dominated by the smallest local
curvature of the interface. A real porous medium has many capillaries (pores) that differ in size,
shape and internal geometry. A capillary pressure exists in an unsaturated system - the
difference between the pressure in the air phase and the water phase. The air phase is at
atmospheric pressure (pair = patm = 0) and the water phase is at less than the atmospheric pressure
(at negative gage pressure) (pw < patm < 0).

In a saturated system, the water is at a pressure greater than atmospheric pressure (pw > patm) and
the energy is greater than some reference state. Water will move spontaneously from soil into
the reference state and the energy (potential) is considered positive. In an unsaturated system the
water is at a pressure less than atmospheric pressure (pw < patm < 0) and the energy level is less
than the reference state. Water will move spontaneously form the reference state into the soil
and the energy level (potential) is considered negative.

Two immiscible fluids in contact exhibit a discontinuity in pressure across the interface
separating them. This pressure difference is capillary pressure pc and it depends on the curvature
of the interface.

𝑝! = 𝑝!" − 𝑝!     (4.1.10)  

where pnw is the pressure in the nonwetting fluid (air, say) and pw is the pressure in the wetting
fluid (water, say). Recall the rise of water in a capillary tube. The capillary forces must balance
the weight of water.

2𝜋𝑟𝜎 𝑐𝑜𝑠 𝛽 =  𝛾𝜋𝑟 ! 𝜓     (4.1.11)  

8
or

!! !"# !
𝜓 =   !"
    (4.1.12)  

where capillary pressure head is defined as


!
𝜓 =   − !! > 0     (4.1.13)  
!

Air
pair

z

pair

€ Negative pw
pressure Solid Solid
ψ

p
Water

Positive pc r
pressure €


Figure 4.4. Capillary rise in a tube.

Consider points on each side of the interface: (A) below the water level

!! !"# !
𝑝! =   −𝛾𝜓 =   − !
<0 (4.1.14)

and (B) above the water level

𝑝!" =  0  (𝑎𝑖𝑟) (4.1.15)

or the difference in pressure across the interface is

!! !"# !
𝑝! = 𝑝!" − 𝑝! =   !
(4.1.16)

This expression relates capillary pressure across an interface to curvature, surface tension and
contact angle. What happens to capillary pressure as curvature is reduced? What happens to
capillary pressure as curvature is increased? What happens to capillary pressure as surface
tension goes to zero?

9
4.1.7. Drainage
Drainage is a process that occurs when the water pressure in the pores becomes less than the air
pressure. Water adheres to solid surfaces better than air does and the air-water interface becomes
curved and the interfacial forces are opposing the pressure forces. Equilibrium occurs when the
interfacial forces balance the pressure forces.

solid

Pore water r Pore air


press. = -p press. = 0

solid

Figure 4.5. Drainage of a pore

Interfacial tension prevents displacement of water in the left pore. A force balance can be
written

2𝜋𝑟𝜎 = 𝜋𝑟 ! 𝑝!    (4.1.17)  

or
!!
𝑟 =   !     (4.1.18)  
!

If pc increases, then the radius must decrease, or water occupies smaller pores. Water drains
from the large pores first. Water recedes into pores small enough to support the interface with a
radius required to balance the capillary force.

4.1.8. Energy Relationships in Fluids

4.1.8.1. Hydraulic Head


The hydraulic grade line (HGL) is the height that water rises to in a vertical tube connected to a
piezometer opening
!
𝐻𝐺𝐿 = ! + 𝑧   (4.1.19)  

10
The energy grade line (EGL) is the height water rises to in pitot tubes inserted into the flow field
of a pipe

!! !
𝐸𝐺𝐿 = !! + ! + 𝑧   (4.1.20)  

where z (m) is the elevation head (potential energy per unit weight of fluid), p/γ (m) is the
pressure head (pressure energy per unit weight of fluid), and v2/2g (m) is the velocity head
(kinetic energy per unit weight of fluid). The velocity in groundwater systems is generally very
slow and the velocity head is almost always neglected. The hydraulic gradient is

! !
𝐻𝑦𝑑𝑟𝑎𝑢𝑙𝑖𝑐  𝐺𝑟𝑎𝑑𝑖𝑒𝑛𝑡 = !" !
+ 𝑧   (4.1.21)  
and the energy gradient is

! !! !
𝐸𝑛𝑒𝑟𝑔𝑦  𝐺𝑟𝑎𝑑𝑖𝑒𝑛𝑡 = !" !!
+ ! + 𝑧   (4.1.22)  

4.1.8.2. Piezometric Head


The state of a flow system is described by its energy and can be characterized by the head.
Relative difference in energy give rise to water movement from regions of high energy to areas
of lower energy. The rate of decrease of energy per unit length of flow is the driving force for
flow, i.e.,

!! ! !
− !" = − !" !
+ 𝑧   (4.1.23)  

the minus sign indicates that the driving force is in the direction of decreasing energy. In
groundwater problems, we neglect the velocity head and use
!
ℎ = ! + 𝑧   (4.1.24)  

for an incompressible fluid. In a compressible fluid, the density r varies with pressure and
temperature ρ = ρ (p, T), so actually we have to consider
!"
− !" = 𝜌𝑔   (4.1.25)  

where we must introduce an equation of state, say ρ = ρ(p), or


! !"
ℎ=𝑧+ ! ! ! !
  (4.1.26)  

11
p
"

!
z

Figure 4.4. Definition of Piezometric


! Head.
Source: USGS (1999)

4.1.8.3. Piezometric Head in Unsaturated Flow


We can also define piezometric head in unsaturated flow
! !
ℎ = 𝑧 + !! = 𝑧 − ! ! (4.1.27)
! !
or

ℎ =𝑧−𝜓 (4.1.28)

Saturated Zone Water Table Unsaturated Zone


θ=φ θ<φ
ψ<0 ψ=0 ψ>0
pw > 0 pw = 0 pw < 0

4.1.9. Soil Water Characteristic Curves

Capillary pressure depends on (a) pore size distribution, (b) contact angle, and (c) moisture
content. In natural porous media, the geometry of the pore space is so complicated that we
cannot describe the pore size distribution or the contact angle analytically, but we can determine
the relationship between capillary pressure head and moisture content

𝑝! = 𝑝! 𝜃    𝑜𝑟      𝜓 =  𝜓(𝜃) (4.1.29)

This must be done experimentally in the laboratory.

12
ψ# Vadose
Zone Porosity

ψ = ψ (θ )

Capillary
Zone
ψb! €
Critacal
Head
(Bubbling
Press.) θo φ#
Irreducible Porosity
Water content
Figure 4.6. Soil water characteristic curve

If we start with a saturated system and drain it, the large pores empty first at low suction and the
small pores don’t empty until we get to a large suction (capillary pressure head). ψb is the
critical head (or bubbling pressure) where almost no water will drain until this level of head is
reached. θo is the irreducible water content that will remain even at extremely high heads.

4.1.10 Aquifer Types


An aquifer (from the Greek, aqua - water, fer - bearing) is a geologic formation capable of
storing (porosity) and transmitting (permeability) a significant quantity of water. Aquifers are
typically composed of (1) unconsolidated deposits of sand and gravel, (2) sandstones and
conglomerates, or (3) carbonate rocks. An aquiclude is a geologic formation capable of storing
significant quantities but incapable of transmitting them in any significant quantity. Typically
composed of clays and shales. These formations often form the impervious boundaries of
aquifers. An aquitard, on the other hand, is a geologic formations which cannot transmit
significant quantities of water. Typically composed of clays and shales. These formations often
form the semi-pervious, leaky confining layers of aquifers.

A confined aquifer is under pressure greater than the sum of hydrostatic plus atmospheric
pressure and is bounded above and below by impervious (or at least semi-pervious) layers.
Water in a well (or piezometer) in a confined aquifer will rise above the confining layer. If the

13
water in the well reaches the ground surface the aquifer is commonly referred to as artesian. A
piezometer (or observation well) indicates the piezometric head at the point of the screen of the
piezometer.

Table 4.2. Water - Where is it?


Share of World Water (%)
Form of Water 3 Total Water Fresh Water
Volume (km )
Ocean 1,338,000,000 96.5 -
Groundwater 23,400,000 1.7 -
Fresh Groundwater 10,530,000 0.76 30.1
Soil Moisture 16,500 0.001 0.05
Glaciers 24,064,100 1.74 68.7
Permafrost 300,000 0.022 0.86
Lakes-fresh 91,000 0.007 0.26
Lakes-saline 85,400 0.006 -
Marsh 11,470 0.0008 0.03
Rivers 2,120 0.0002 0.006
Biological 1,120 0.0001 0.003
Atmospheric 12,900 0.001 0.04
Total Water 1,385,984,610 100 -
Fresh Water 35,029,210 2.53 100.

Source: World Water Balance and Water Resources of the Earth, UNESCO, Paris, 1978

Table 4.3. Usage of Groundwater in United States (1980).


Usage Type Usage
Total 88x109 Gal. per day
Public Supply 14%
Rural Supply 5%
Irrigation 68%
Industry 13%
Source: National Water Summary, U.S. Geological Survey,
Washington, D.C., 1987

An unconfined aquifer (also known as a phreatic or water table aquifer) is bounded above by a
water table, or fictitious surface where p = patm = 0 (gage). A capillary fringe of moisture exists
in the pore space above the water table (phreatic surface). Unconfined aquifers are capable of
being recharged directly by infiltration from the ground surface.

4.1.11 Aquifer Properties

4.1.11.1 Hydraulic Conductivity


Darcy’s law, in fact, any law describing a physical phenomenon, is meaningless unless we can
determine the numerical value of the coefficient or coefficients appearing in it. Here it is the
hydraulic conductivity K (Bear, 1972). Hydraulic conductivity (K [L/T]) represents the ease

14
with which a subsurface formation can transmit water under a hydraulic gradient. It is the
specific discharge (discharge per unit area) per unit hydraulic gradient. It depends on both the
matrix and the fluid properties of the system, in particular, the fluid properties of density ρ, and
viscosity µ, and the matrix properties of pore size distribution, pore shape, tortuosity, specific
surface area, and porosity. Hydraulic conductivity can be written as

ρg
K =k
µ (4.1.30)

where k is the intrinsic permeability. Kozeny and Carman developed the following equation to
estimate the intrinsic permeability from the properties of porosity and grain size (McCabe and
Smith 2005):

⎛ n3 ⎞ 2
k = cd 2 = ⎜ ⎟d
⎜ 180(1 − n ) 2 ⎟
⎝ ⎠ (4.1.31)

where d is the mean particle size.

4.1.11.2 Storativity
Storativity (S) represents the ability of a formation to store water, or the change in the volume of
water stored in a formation due to changes in the piezometric head in the formation. It is the
volume of water released (taken up) from a vertical column of aquifer with a unit cross-sectional
area per unit decline (rise) in piezometric head. In a confined aquifer, the compressibility of the
water in the pore space and the compressibility of the solid matrix determine the storage
capabilities of the aquifer. In an unconfined aquifer, on the other hand, pore space drainage is
the dominant factor determining the storage characteristics of the aquifer. In this case, a certain
amount of water will be retained in the pores due to surface tension (capillary action) and
molecular forces between the water molecules and the solid matrix. Thus, in an unconfined
aquifer, the storativity is less than the porosity by an amount known as the specific retention.
The storativity of an unconfined aquifer is often called the specific yield. The elastic storativity
of a confined aquifer (S ≅ 10-5) is much less than the specific yield of an unconfined aquifer (Sy ≅
10-1 or 10 - 25%).

The specific yield of an unconfined aquifer is the water obtained by gravity drainage when the
water table declines y a unit amount

Volume  from gravity drainage
Sy =
Volume total (4.1.32)

Volume drained

𝑉!"#$%&! = 𝑆! ∆ℎ𝐴 (4.1.33)

15
The Specific Retention of the unconfined aquifer is defined as the water retained after gravity
drainage (sometimes called field capacity)

Sr = φ − Sy
(4.1.34)

The storativity of a confined aquifer depends on both the compressibility of the water (β) and the
compressibility of the porous medium itself (α).

S = ρ g(α + φβ ) (4.1.35)

Unconfined Aquifer Confined Aquifer

Ss = S y S s = ρg (α + nβ )
"V = S y A"h "V = S s A"h
Figure 4.3. Storage relations of confined and unconfined aquifers.
Source: Heath (1983)

! !
4.2. Darcy’s Law
Henri Darcy published a study of the design of sand filters for the city of Lyon, France in 1856
(Darcy, 1856). In this study he found that the flow rate, Q [L3/T], through a vertical sand filter
was proportional to (1) the cross-sectional area of the filter, A [L2], and (2) the head loss through
the filter, Δh [L], and inversely proportional to the length of the flow path, L [L], or

Q h −h h −h dh
q= = K 1 2 = −K 2 1 = −K
A L L dl (4.2.1)

where K, the hydraulic conductivity, is the constant of proportionality, and q is the specific
discharge [L/T]. Note that the average velocity through the area of porous medium is V = q/n
where n is the porosity.

16
In general
∆!
lim∆!→! ∆!
= ∇ℎ (4.2.2)

where Δl is positive and directed normal to surface of constant h.

!! !! !!
𝛻ℎ = 𝑔𝑟𝑎𝑑 ℎ = !" 𝑒! + !" 𝑒! + !" 𝑒!   (4.2.3)  

is a vector in the direction of increasing h.

Darcy’s Law is typically generalized to three-dimensions as

𝑞 = −𝐾𝛻ℎ   (4.2.4)  
 
!! !! !!
𝑞 = 𝑞! 𝑒! + 𝑞! 𝑒! + 𝑞! 𝑒! = −𝐾𝛻ℎ = −𝐾 !" 𝑒! − 𝐾 !" 𝑒! − 𝐾 !! 𝑒!   (4.2.5)  

The units of hydraulic conductivity are typically reported in m/s or gal/(day-ft2), where 1
gal/(day-ft2) = 4.72x10-7 m/s. For typical values of conductivity for various types of porous
media, refer to Freeze and Cherry (1979) or Bear (1972).

hL
P1/γ"

P2/γ"
v
Q
,A

z1 Sand
ea
Ar

column

Datum z2
plane
Q

Figure 4.5. Darcy’s Experimental Apparatus.

17
Figure 4.6. Experimental results obtained by Darcy.

y !ℎ
Circular hydraulic !
head contours !! !
Δh"

!!
q"

Well, Q x
!!
h1
h2 h3

h1 < h2 < h3

Figure 4.7. Plan view of an aquifer being pumped by a single well. Conductivity is constant
through the aquifer.

The specific discharge (or Darcy velocity) is a vector which represents the volume flow rate or
discharge from a unit cross-section area of aquifer formation normal to the direction of flow.
The average velocity (or seepage velocity) is

q
v= (4.2.6)
φ
Darcy’s Law is a macroscopic equation of conservation of momentum, expressing the
relationship between head gradient and drag force resulting from viscous stress on the matrix-
fluid interface. The drag force is the average force exerted by the fluid on the surface per unit
volume and is a linear function of specific discharge. Darcy’s Law can be derived by solving the

18
Navier-Stokes equations for various porous medium models, e.g., a bundle of capillary tubes, a
block of fractures, hydraulic radius approach, resistance to flow approach, creeping flow
approach, or averaging the Navier-Stokes equations. All of these approaches result in some
expression for the specific discharge like
!
𝒒 = −𝑐𝑙 ! ! ∇ℎ (4.2.7)

Darcy’s Law specifies a linear relationship between specific discharge and hydraulic gradient.
At some upper limit of specific discharge, this linear relations is no longer valid. Darcy’s Law is
valid only for laminar flow through porous media at low Reynolds numbers. Darcy’s Law is
valid for flow conditions when the Reynolds number characterizing the flow (specific discharge,
effective grain size diameter, and viscosity) is not excessive (Bear 1972)

ρ qd10
NR = <1 (4.2.8)
µ

where d10 is the effective grain size diameter.

−∇ℎ

Experiment
Re = 10 shows this

Re = 1 Darcy Law
predicts this

α"
q
tan-1(α)= (1/K)
Figure 4.8. Relationship between specific discharge and head gradient illustrating the
region of validity for Darcy’s Law.

Hydraulic conductivity has dimensions of velocity [L/T] and is comprised of porous medium
properties and fluid properties. It is a combined property of the medium and the fluid and
represents the ease with which fluid moves through the medium.
!
𝑲 = 𝑘! (4.2.9)

where k = cd2 is intrinsic permeability, ρ is density, µ is dynamic viscosity, and g is specific


weight.

19
4.3. Heterogeneity and Anisotropy
4.3.1. General
Most subsurface formations are extremely heterogeneous and consequently, aquifer properties
are as well (Freeze and Cherry, 1979; Dagan, 1989; Gelhar, 1994). Homogeneous aquifer
properties are ones whose values are independent of spatial location, whereas the values of
heterogeneous properties vary in space. Isotropic aquifer properties are ones whose value is
independent of direction, otherwise they are anisotropic. Anisotropy often results from
stratification resulting from different particle shapes and sizes during sedimentation.

Aquifers are typically composed of unconsolidated geologic deposits of complex arrays of lenses
or strata of essentially unknown geometry and variable hydraulic properties. The degree of
heterogeneity and the spatial structure of the hydraulic properties has a large effect on the flow
and mass transport characteristics of the aquifer. The conventional approach to groundwater
modeling is to treat an aquifer a an idealized homogeneous formation with a constant value for
each physical parameter so that the idealized aquifer behaves like the actual formation
(Vomvoris and Gelhar, 1986). In cases of predictable heterogeneity, one can subdivide the
formation into a number of homogeneous subzone, each of which with a different equivalent
value for the parameter of interest.

The complexity of most common groundwater systems has lead to the consideration of the
physical properties of the porous medium as stochastic (probabilistic) processes or spatial
random functions where the random hydraulic parameters are represented statistically (Dagan,
1989; Gelhar, 1994). Then average macroscopic or “equivalent” parameters are derived for
making flow calculations on a large scale. The macroscopic parameters are related to the
statistical parameters describing the complex three-dimensional heterogeneity of the porous
medium’s hydraulic properties.

From the available data, we expect groundwater velocity in a particularly formation will vary
irregularly in three-dimensions over scales of approximately 1-10 cm in the vertical direction and
1-2 m in the horizontal direction. The fundamental question to be answered using the data:
“How do we deal with the heterogeneity in order to make predictions of flow and mass transport
over scales ranging up to 100’s of meters?” It is impractical to make these detailed
measurements over the scale of 100’s of meters and construct a deterministic description of the
site.

This problem can be handled by conceiving of the hydraulic conductivity distribution as a spatial
random field of which the measured data at hand are a representative sample with spatial
persistence characterized by a covariance function estimated from the data. From this stochastic
representation (i.e., the covariance function of the logarithm of hydraulic conductivity)
“effective” hydraulic conductivity values can be estimated which are applicable to the large scale
problem.

20
4.3.2 Equivalent conductivity in layered aquifer formations
In natural porous media systems, we often run into layered formations where each layer has a
distinct conductivity. That is, we see many depositional systems where the layers are deposited
in a manner that confers a preference for horizontal rather than vertical flow. Thus, flow parallel
to layers experiences less resistance than flow perpendicular to the layers. In many cases we can
replace the individual conductivities of the layers with a single, equivalent hydraulic
conductivity. Here we distinguish between horizontal and vertical flow systems (or equivalently,
parallel and perpendicular flow systems). In a parallel flow system the flow is parallel to the
layering as in the following figure.

Piezometric surface

Δh
h1
h2

datum
b1 K1 Q1

Q2 Q
b" b2 K2

b3 K3 Q3

W"
Figure 4.9. Parallel flow in a layered system.

! ! !!
𝑄= !!! 𝑄! = !!! −𝑏! 𝐾! !" (4.3.1)

!! !!!
𝑄 = − !!!! 𝑏! 𝐾! !
  (4.3.2)  
 
 
!! !!!
𝑄 = − 𝑏𝐾 !
  (4.3.3)  

21
 
!
!! ! !
𝐾= !!!
!
  (4.3.4)  

In a perpendicular flow system the flow is normal to the layering as in the following figure.

Piezometric surface
Δh1"
Δh2" Δh"
Δh3"

K1 K2 K3

Q"
b"
Q"

L1" L2" L3"


L"
Figure 4.10. Perpendicular flow in a layered system.

∆!!
𝑄 = 𝐾! (𝑏𝑥1) !!
  (4.3.5)  
 
! !!
∆ℎ! = ! !!
  (4.3.6)  

22
 
! ! ! !!
∆ℎ = !!! ∆ℎ! = ! !!! !   (4.3.7)  
!
 
∆! ∆!
𝑄=𝑏 ! !! = 𝑏𝐾 !
  (4.3.8)  
!!! !
!
 
!
𝐾= ! !!
  (4.3.9)  
!!! !!

4.3.3 Anisotropic Porous Media


In isotropic porous media K = K(x,y,z) is a scalar (constant) and the specific discharge and
hydraulic gradient are collinear.

𝑞 = −𝐾𝛻ℎ   (4.3.10)  

Layered porous media results in an equivalent porous medium with directional conductivity
greater along the layers than across the layers. Specific discharge and hydraulic gradient are no
longer collinear, but they are linearly related. This motivates a generalization of Darcy’s Law
for anisotropic porous media. The most general relationship between specific discharge and
hydraulic gradient is

!! !! !!
𝑞! = −𝐾!! !" −𝐾!" !" −𝐾!" !"   (4.3.11a)  
!! !! !!
𝑞! = −𝐾!" !" −𝐾!! !" −𝐾!" !"   (4.3.11b)  
!! !! !!
𝑞! = −𝐾!" !" −𝐾!" !" −𝐾!! !"   (4.3.11c)  

which we write in matrix form as

!!
𝑞! 𝐾!! 𝐾!" 𝐾!" !"
!!
𝑞! = − 𝐾!" 𝐾!! 𝐾!"   (4.3.12)  
!"
𝑞! 𝐾!" 𝐾!" 𝐾!! !!
!"
or
𝑞 = −𝐾𝛻ℎ   (4.3.13)  

K is symmetric, i.e., Kij = Kji. Note that flow in the x direction is influenced not only by the
!! !! !!
gradient 𝐽! = − !" , but also by the other gradients 𝐽! = − !" and 𝐽! = − !" .

4.4. Potentiometric Surfaces


We can use flow field data to determine the direction of groundwater movement and the
hydraulic gradient. The data required from boreholes, piezometers and wells include: surface

23
elevation, total depth, water levels, and screen or perforation levels. The water level data can be
plotted and contoured on a map and groundwater flow lines drawn at right angles to the water
level elevation contours. Then we can calculate the hydraulic gradient

!"#$"%&"  !"  !"#$%  !"#"!


𝐽 = 𝛻ℎ = !"#$  !!"#$!%&'(  !"#$%&'(   (4.4.1)  

for different points on the map. Caution: Potentiometric surface is not static nor is it a simple
linear surface so measurements should be form the same time and the same screened interval
depth. In multi-aquifer systems, each unit has its own piezometric surface, direction of flow and
hydraulic gradient.

The magnitude of the hydraulic gradient is

𝐽 = 𝐽!! + 𝐽!! + 𝐽!!   (4.4.2)  

and the direction of the hydraulic gradient is (in 2 dimensions)


!!
𝜃 = 𝑡𝑎𝑛!! !!
  (4.4.3)  

For three closely spaced wells (piezometers), we can fit a plane through the head field

ℎ = 𝑎 + 𝑏𝑥 + 𝑐𝑦   (4.4.4)  

where a, b and c are determined form the data, i.e.,

𝑎 + 𝑏𝑥! + 𝑐𝑦! = ℎ!   (4.4.5a)  


𝑎 + 𝑏𝑥! + 𝑐𝑦! = ℎ!   (4.4.5b)  
𝑎 + 𝑏𝑥! + 𝑐𝑦! = ℎ!   (4.4.5c)  

and this set of linear equations can be solved for a, b and c given (xi, hi, i=1, 2, 3). Note:

!!
𝐽! = − !" = −𝑏   (4.4.6a)  

!!
𝐽! = − !" = −𝑐   (4.4.6b)  
Then

𝐽 = 𝑏 ! + 𝑐 !   (4.4.7)  

and
!
𝜃 = 𝑡𝑎𝑛!! !
  (4.4.8)  

24
4.5. General Groundwater Flow Equations
Darcy’s law (conservation of momentum) relates the specific discharge (q) and piezometric head
(h) in an aquifer. This vector expression contains four unknowns (qx, qy, qz, and h) but only
three equations:

∂h ∂h ∂h
q x =− K , q y =− K , and q z =− K (4.5.1)
∂x ∂y ∂z

Conservation of mass (continuity) provides the additional equation required to close the system.
Consider the elemental control volume or representative elementary volume (REV) of porous
media shown in Figure 4.11. The Control Volume has dimensions Δx, Δy, Δz, is completely
saturated with a fluid of density ρ, and the point P(x, y, z) is located at the center. The fluid may
be compressible, but its density depends only on the fluid pressure.

The principle of conservation of mass for the Control Volume requires that for a small time
period:

mass flux in - mass flux out + mass generation = change of mass storage

Mass!flux!in! Mass!flux!out!
∂ ( ρq x ) Δx ∂ ( ρq x ) Δx
ρq x − Δz ρq x +
∂x 2 Control ∂x 2
volume

z! y!
Δy
Δx
Δx Δx
x!
x−
2
x x+
2

Figure 4.11. Control volume for continuity calculations.

The mass flux per unit time at the point P is ρq. The mass flux through the sides of the REV can
be found from a Taylor series expansion of the mass flux at point P. For example, the mass flux
per unit area through the side at x+Δx/2 is

( )
∂ ρq x 2
( )
Δx 1 ∂ ρq x ⎛ Δx ⎞
2
(4.5.2)
ρq x = ρq x + + ⎜ ⎟ + !
x + Δx / 2 x ∂x 2 2 ∂x 2 ⎝ 2 ⎠
x x

where variables and derivatives on the right hand side are evaluated at the point (x,y,z). The
Taylor series is truncated after the second (linear) term and higher order terms (involving Δx2)
are neglected, i.e., a local approximation.

25
∂ (ρq x ) Δx
ρq x ≈ ρq x x + (4.5.3)
x + Δx / 2 ∂x x 2

Similar expressions can be developed for the mass flux across the side at x-Δx/2 and in the y and
z directions.

Contributions to the mass flux into and out of the Control Volume over a short time interval Δt
are:

x direction: ⎡ ∂ (ρq x ) Δx ⎤ ⎡ ∂ ( ρq x ) Δx ⎤
⎢⎣ ρq x − ΔyΔzΔt ρq x + ΔyΔzΔt
∂x 2 ⎥⎦ ⎢⎣ ∂x 2 ⎥⎦
y direction: ⎡ ( )
∂ ρq y Δy ⎤ ⎡ ∂( ρq y ) Δy ⎤
⎢ ρq y − ⎥ ΔxΔzΔt ⎢ ρq y + ⎥ ΔxΔzΔt
⎣ ∂ y 2 ⎦ ⎣ ∂ y 2 ⎦
z direction: ⎡ ∂ ( ρ q z ) Δz ⎤ ⎡ ∂ ( ρq z ) Δz ⎤
⎢⎣ ρq z − ∂z 2 ⎥⎦
ΔxΔyΔt ⎢⎣ ρq z + ∂z 2 ⎥⎦
ΔxΔyΔt

Adding all of the contributions to mass flux into the Control Volume and subtracting all of the
contributions to mass flux out of the Control Volume results in

⎡ ∂( ρq x ) ∂( ρq y ) ∂( ρq z ) ⎤
⎢ + + ⎥ ΔVΔt (4.5.4)
⎣ ∂x ∂y ∂z ⎦

where ΔV = ΔxΔyΔz is the volume of the Control Volume.

If n is the porosity of the porous media, then the mass of fluid, Δm, stored in the REV is
Δm = φρΔV. The rate of change of mass in the Control Volume over Δt is

∂(φρΔV ) ∂(ρ qx )
=− ΔV (4.5.5)
∂t ∂x

Combining and dividing by Δt, we have

# ∂ ( ρ q ) ∂ ( ρ qy ) ∂ ( ρ q ) & ∂ (φρΔV )
x z
−% + + ( ΔV = (4.5.6)
%$ ∂x ∂y ∂z (' ∂t

or

! !"
−∇. (𝜌𝒒) = (4.5.7)
!"

where we can expand the right hand side as

26
! !" !" !" !" !" !" !"
=𝜌 +𝜙 =𝜌 +𝜙 (4.5.8)
!" !" !" !" !" !" !"

The compressibility of water is

! !"
𝛽=   (4.5.9)  
! !"
or  
!"
= 𝛽𝜌   (4.5.10)  
!"
 
and  the  compressibility  of  the  porous  medium  (coefficient  of  consolidation)  is    
 
!"
= 𝛼   (4.5.11)  
!"

Then we have

! !" !" !!
!"
= 𝜌 𝛼 + 𝛽𝜙 !"
= 𝜌 𝛼 + 𝛽𝜙 𝜌𝑔 !"   (4.5.12)  

or

! !" !!
= 𝜌! 𝑔 𝛼 + 𝛽𝜙   (4.5.13)  
!" !"
 
!!
−∇. (𝜌𝒒) = 𝜌! 𝑔 𝛼 + 𝛽𝜙 !"
  (4.5.14)  
 
!!
−∇. (𝜌𝒒) = 𝜌𝑔 𝛼 + 𝛽𝜙 !"
  (4.5.15)  
 
where  the  aquifer  storativity  is  𝑆 = 𝜌𝑔 𝛼 + 𝛽𝜙  so  
 
!!
−∇. (𝜌𝒒) = 𝑆 !"   (4.5.16)  
 
Recall Darcy's law

𝒒 = −𝐾∇ℎ  

This can be substituted into the expression for the conservation of mass to yield

!!
∇. 𝐾∇ℎ = 𝑆 !"   (4.5.19)  

27
4.6. Horizontal Flow in Confined Aquifers
4.6.1 Introduction
In general, flow in aquifers is three-dimensional with the specific discharge, q, having
components qx, qy, and qz, in all three coordinate directions and piezometric head h = h(x,y,z,t),
a function of all three coordinates and time. Since most aquifers are thin compared to their
horizontal extent, we can often assume that flow is essentially horizontal or two-dimensional,
i.e., we can neglect vertical flow components, qz, and variations in head over the thickness of the
aquifer, so that h(x, y, z, t) = h(x, y, t). This is a good approximation when variations in the
thickness of the aquifer are small compared with average aquifer thickness. See Bear [1979 pp.
26 - 28, 69, 103] for a discussion of situations where this "hydraulic" approach is justified.
Consider flow in the leaky, semiconfined aquifer shown in Figure 1 where K >> K’. Flow in an
unconfined or phreatic aquifer will be considered later.

4.6.2 Confined Groundwater Flow Equations


The assumption of horizontal flow implies that the piezometric head, h(x,y,z,t), can be replaced
by an average head, h(x,y,t) , with the average taken over the thickness b of the aquifer along a
line from the bottom to the top of the aquifer

1 b
h (x,y,t)= ∫ h(x,y,z,t)dz (4.6.1)
b 0

The transmissivity is defined as the discharge through the entire thickness of the aquifer per unit
width per unit head gradient. Transmissivity is equal to the product of the average hydraulic
conductivity and the aquifer thickness, T =K b , where

1 b
K (x,y)= ∫ K(x,y,z)dz (4.6.2)
b 0

Aquifer storativity, S, defined as the volume of water, Vw, released from (added to) storage per
unit area of aquifer per unit decline (or rise) in average head, h , in the aquifer, is equal to the
product of the specific storativity and the aquifer thickness

Vw
S=S s b= (4.6.3)
AΔ h

where Ss is the volume of water released from storage per unit volume of porous media during a
unit decline in head, and A is the area under consideration. The average specific discharge
components are

28
1 b 1 b
q x(x,y,t)= ∫ q x(x,y,z,t)dz , and q y (x,y,t)= ∫ q y (x,y,z,t)dz (4.6.4)
b 0 b 0

Using Darcy's law, we can write Q = bq − −T∇h , where Q is the discharge per unit width of
aquifer and T is the transmissivity tensor. The components of discharge per unit aquifer width in
the x and y directions are

∂h ∂h
Q x =bq x =− Tx , and Q y =bq y =− T y (4.6.5)
∂x ∂y

where we have assumed that x and y are in the direction of the principal components of the
hydraulic conductivity (and transmssivity) tensor.

ground surface
head in unconfined aquifer, i.e., water
head in confined table
aquifer -
unconfined h
aquitard 0
aquifer
K' b'
-
q (b) h
K z Q' z b
confined x
y aquifer
x
bedrock control volume
Δx q (0)
z

Figure 4.12. Leaky confined aquifer.

The two-dimensional conservation of mass (continuity) equation for a semiconfined (leaky)


aquifer can be derived by averaging the three-dimensional continuity equation over the vertical
extent of the aquifer. If there is no leakage from below and flow in the confining layer is
essentially vertical, and we have another aquifer above the semipervious layer where the head is
h0 , we obtain the equation of continuity for unsteady two-dimensional flow in a heterogeneous,
anisotropic semiconfined aquifer (Bear, 1979)

K ʹ′ ∂h
∇ ⋅ T ∇h + (h0 −h ) = S −W (4.6.6)
bʹ′ ∂t

where W is the source or sink function averaged over the aquifer thickness.

29
4.6.3. Examples

4.6.3.1. Flow in a confined aquifer


Consider the steady flow from left to right in the confined aquifer shown in the Figure.
Determine the head in the aquifer.

Figure 4.13. Flow in a confined aquifer.

The governing equation for flow in the aquifer can be written as

∇ ⋅ K∇h = 0 (4.6.7)

If the aquifer is considered to be homogeneous and isotropic (i.e., K is a constant), we can write
this equation as

∇2 h = 0 (4.6.8)

Since the flow is one-dimensional, we have

d 2h
=0 (4.6.9)
dx 2

The solution of this equation is (integrate twice and apply the boundary conditions shown in the
Figure)

30
h − hA
h( x ) = h A + B x (4.6.10)
L

Consider the values L = 1000 m, HA = 100 m, HB = 80 m, K = 20 m/day (clean sand), b = 50 m, n


= 0.35. Using these values, we have

h( x) = 100 − 0.02 x (4.6.11)

The specific discharge is

dh
q=−K
dx
hB − h A
=−K (4.6.12)
L
= 0.4 m / day

The average velocity is

q
v= = 1.14 m / day (4.6.13)
n

4.6.3.2. Flow in a leaky confined aquifer


Consider the steady flow from left to right in the semi-confined aquifer shown in Figure 4. The
aquitard is leaky. Determine the head in the aquifer. The governing equation for flow in the
aquifer can be written as

K ʹ′
∇ ⋅ T∇h + (h − h) = 0 (4.6.14)
bʹ′ 0

Or

∂ ⎛ ∂h ⎞ ∂ ⎛ ∂h ⎞ K ʹ′
⎜ T ⎟ + ⎜ T ⎟ + (h − h ) = 0 (4.6.15)
∂x ⎝ x ∂x ⎠ ∂y ⎜⎝ y ∂y ⎟⎠ bʹ′ 0

If the aquifer is considered to be homogeneous and isotropic, we can write this equation as

h −h
∇2 h + 0 =0 (4.6.16)
λ2

where λ2 = bKb’/K’.

31
Figure 4.14. Flow in a leaky-confined aquifer.

Now for one-dimensional flow, we have

2 d 2h
λ − h = − h0 (4.6.17)
dx 2

a second-order ordinary differential equation with constant coefficients which has the solution

L−x x
(h A − h0 ) sinh( ) + (hB − h0 ) sinh( )
h( x) = h0 − λ λ (4.6.18)
L
sinh( )
λ

Consider the values L = 1000 m, HA = 100 m, HB = 80 m, K = 20 m/day (clean sand), b = 50 m,


b’ = 2 m, K’ = 0.10 m/day (silt), n = 0.35. The head distributions for values of h0 (head in the
overlying aquifer) are shown in the Table and plotted in the Figure 5.

32
110

105

100

He 95
ad
(m 90
) H0=110
85
H0=105
80 H0=100
H0=95
75
H0=90
70
0 200 400 600 800 1000
Distance (m)

Figure 4.15. Semi-confined aquifer head values for various overlying aquifer head levels.

Table 4.4. Semi-confined aquifer head values for various overlying aquifer head levels.

X h0=110 h0=105 h0=100 h0=95 h0=90


0.00 100.00 100.00 100.00 100.00 100.00
100.00 105.03 102.50 99.97 97.45 94.92
200.00 107.47 103.70 99.93 96.17 92.40
300.00 108.59 104.23 99.86 95.49 91.13
400.00 108.98 104.35 99.71 95.08 90.45
500.00 108.84 104.13 99.42 94.71 90.00
600.00 108.08 103.45 98.82 94.19 89.55
700.00 106.33 101.97 97.60 93.24 88.87
800.00 102.67 98.91 95.14 91.37 87.60
900.00 95.19 92.67 90.14 87.61 85.08
1000.00 80.00 80.00 80.00 80.00 80.00

4.6.3.3. Steady Radial Flow in a Confined Aquifer


Consider a pumping well in a homogeneous, isotropic aquifer of infinite extent shown in the
following Figure. The governing equation for flow in this aquifer is

∂h
∇ ⋅ T∇h = S (4.6.19)
∂t

Which for steady state becomes

∇2 h = 0 (4.6.20)

33
Q
Ground surface
Prepumping
head
Pumping
Drawdown curve well

Observation
wells h1
h2
h0 r1
hw
Confining Layer r2

Confined b Q
aquifer

Bedrock
2rw
Figure 4.14.

Let’s write this in radial coordinates

1 ∂ ∂h 1 ∂ 2 h
(r ) + =0 (4.6.21)
r ∂r ∂r r ∂θ 2

But for radial symmetry (isotropic medium), the second term is zero, so

d dh
(r ) = 0 (4.6.22)
dr dr

Integrating, we have

dh
r = C1 (4.6.23)
dr

Now, by Darcy’s Law, we have

dh
Q = Aq = (2πrb) K (4.6.24)
dr

So

34
dh Q
r = = C1 (4.6.25)
dr 2πT

Therefore

Q dr
dh = (4.6.26)
2πT r

Finally, we have Theim’s equation

Q r
h2 = h1 + ln( 2 ) (4.6.27)
2πT r1

We define the drawdown in an aquifer as the amount that the piezometric surface is lowered due
to pumping

s( r ) = h0 − h( r ) (4.6.28)

And the “cone of depression” is the resulting piezometric surface near a well, which is
logarithmic in shape and extends radially outward from the well. The extent and shape of the
cone of depression depend ultimately on the hydraulic conductivity of the aquifer, pumping rate
and the length of time the aquifer has been pumped.

We can write the Theim Equation in terms of drawdown as

Q r
s( r ) = h0 = h( r ) = ln( 0 ) (4.6.29)
2πT r

35
Figure 4.15. Drawdown due to pumping in a confined aquifer.

4.6.3.4. Theim Equation Starting from Darcy’s Law


The flow through a cylinder of radius r, where the head is equal to h, is equal to the discharge
from the well

∂h
Q = Aq= 2πrbK (4.6.30)
∂r

where flow is in the direction toward the well, opposite the radial direction r. If h = h0 at r = R,
then this equation can be integrated to yield the Theim equation

Q ⎛ r ⎞
h(r ) = h0 + ln⎜ ⎟ (4.6.31)
2πT ⎝ R ⎠

where T is the transmissivity of the aquifer. The drawdown due to a pumping well is the vertical
distance between the initial piezometric surface and the piezometric surface at some later time at
the same point. In this case, the drawdown is

Q ⎛ R ⎞
s(r ) = ln⎜ ⎟ (4.6.32)
2πT ⎝ r ⎠

The distance R is called the radius of influence if it is the distance from the well to points at
which practically no drawdown can be observed.

36
4.6.3.5. Drawdown in two monitoring wells due to pumping in a confined aquifer
Consider a well pumping at a rate Q = 400 m3/hr from a homogeneous, isotropic confined
aquifer of constant thickness b = 40 m. Two observation wells, one at a distance r1 = 25 m from
the pumping well shows a steady water level at +85.3 m the other at r2 = 75 m shows a steady
water level at +89.6 m. Determine the aquifer’s transmissivity and hydraulic conductivity (Bear,
1979, prob. 8.1).

Rearranging the Theim equation, we have

Q r 400 25
T= ln( 1 ) = ln( ) = 16.3 m 2 / hr
2π ( h1 − h2 ) r2 2π (85.3 − 89.6) 75
(4.6.33)
T 16.3
K= = m / hr = 1.13 x10 − 4 m / s
B 40

Figure 4.16. Drawdown in two monitoring wells due to pumping in a confined aquifer.

4.6.3.6. Steady Radial Flow in a Leaky (Semi-Confined) Aquifer


Consider the steady flow from left to right in the semi-confined aquifer shown in Figure 4. The
aquitard is leaky. Determine the head in the aquifer. The governing equation for flow in the
aquifer can be written as

37
K ʹ′
∇ ⋅ T∇h + (h − h) = 0 (4.6.34)
bʹ′ 0

If the aquifer is considered to be homogeneous and isotropic, we can write this equation as

h −h
∇2 h + 0 =0 (4.6.35)
λ2

where λ2 = bKb’/K’. Written in radial coordinates, this is

1 ∂ ∂h h0 − h
(r ) + =0 (4.6.36)
r ∂r ∂r λ2

Or

∂ 2h 1 ∂h h h
+ − =− 0 (4.6.37)
∂r 2 r ∂r λ2 λ2

Let η = r / λ . Then

∂ 2h 1 ∂ 2h
= (4.6.38)
∂r 2 λ2 ∂η 2
So
d 2h 1 ∂h
+ − h = −h0 (4.6.39)
dη 2 η ∂η

An ordinary differential equation with solution

⎛ r ⎞ ⎛ r ⎞
h = h0 + AI 0 ⎜ ⎟ + BK 0 ⎜ ⎟ (4.6.40)
⎝ λ ⎠ ⎝ λ ⎠

Where I0 is a modified Bessel function, order aero, of the first kind, and K1 is a modified Bessel
function, order aero, of the second kind, and A and B are constants determined by boundary
conditions. I (η ) → ∞ asη → ∞ , so A = 0. Given pumping Q from the well, we have
0

Q 1
B=− (4.6.41)
2π ( rw / λ ) ⎛ r ⎞
K1 ⎜⎜ w ⎟⎟
⎝ λ ⎠

So that

38
Q ⎛ r ⎞
h = h0 − K 0 ⎜ ⎟ (4.6.42)
(
2π ( rw / λ ) K1 rw / λ ) ⎝ λ ⎠

w 1 w
(w
) w 1 w
( )
In many cases r / λ is small and K r / λ ≈ λ / r and r / λK r / λ → 1, so we have

Q ⎛ r ⎞
h = h0 − K 0 ⎜ ⎟ (4.6.43)
2π ⎝ λ ⎠

Or

Q ⎛ r ⎞
s= K 0 ⎜ ⎟ (4.6.44)
2πT ⎝ λ ⎠

For r << λ

Q ⎛ 1.123λ ⎞
s= ln⎜ ⎟ (4.6.45)
2πT ⎝ r ⎠

4.6.3.7. Transient Radial Flow in a Confined Aquifer


The transient, or time dependent, drawdown resulting from a well pumping in a confined aquifer
of infinite extent is given by the Theis equation. This equation can be derived by considering the
governing equation

∂h
∇ ⋅ T∇h = S (4.6.46)
∂t

Let’s write this in radial coordinates

1 ∂ ∂h 1 ∂ 2 h S ∂h
(r ) + = (4.6.47)
r ∂r ∂r r ∂θ 2 T ∂t

Which, if we consider radial symmetry and write the equation in terms of drawdown

1 ∂ ∂s S ∂s
(r ) = (4.6.48)
r ∂r ∂r T ∂t

Now, we apply a Boltzman transformation of variables u = r 2 S 4Tt , the equation becomes

39
S ⎛⎜ ∂ 2 s ∂s ⎞⎟ S ∂s
u + = (4.6.49)
Tt ⎜⎝ ∂u 2 ∂u ⎟⎠ T ∂t
Or
d 2s 1 ds
+ (1 + ) =0 (4.6.50)
2 u du
du

Another transformation

ds
p= (4.6.51)
du

So
dp 1
+ (1 + ) p = 0 (4.6.52)
du u

Or

e −u
p=c (4.6.53)
u

s(∞) ∞ e −u
∫ ds = c ∫ du (4.6.54)
s(u ) u u

Q
s(u ) = W (u ) (4.6.55)
4πT

where

∞ e −η
W (u ) = ∫ dη (4.6.56)
u η

is the well function and u = r 2 S 4Tt . For small values of u (e.g., at large times or close to the
well) the Jacob approximation to the well function can be used. This results in the Cooper-Jacob
equation for drawdown

Q ⎛ 2.25 ⎞
s(r , t ) = ln⎜ ⎟ (4.6.57)
4πT ⎝ 4u ⎠

40
4.6.3.8. Theis’ method and Jacobs’ approximation
Compare Theis’ method and Jacobs’ approximation to calculate the drawdown of the
piezometric surface in a confined aquifer at distances of r = 100 m and 200 m from a well
pumping Q = 1000 m3/day for 10 days. T = 1000 m2/day and S = 10-4 for the aquifer.

Table 4.5. Drawdown in a confined aquifer using the Theis’s and Jacobs’ methods.

r = 100 r = 200
Time Theis Jacob Theis Jacob
(days) U W(u) s (m) s (m) u W(u) s (m) s (m)
0.001 0.25 1.0443 0.08310 0.06439 1 0.2194 0.01746 -0.0459
0.005 0.05 2.4679 0.20439 0.19246 0.2 1.2227 0.0973 0.08214
0.01 0.025 3.1365 0.24959 0.24762 0.1 1.8229 0.14506 0.1373
0.05 0.005 4.7261 0.37609 0.37569 0.02 3.3547 0.26696 0.26538
0.1 0.0025 5.4167 0.43105 0.43085 0.01 4.0379 0.32133 0.32054
0.5 0.0005 7.0242 0.55897 0.55893 0.002 5.6394 0.44877 0.44861
1 0.00025 7.7172 0.61412 0.61409 0.001 6.3315 0.50384 0.50377
5 0.00005 9.3263 0.74216 0.74216 0.0002 7.9402 0.63186 0.63184
10 2.5E-05 10.0194 0.79732 0.79732 0.0001 8.6332 0.68701 0.687

Time (days)

s100_Theis
0.25
s200_Theis
Dr
aw
do
wn 0.5
(m)

0.75

1
0 1 2 3 4 5 6 7 8 9 10

Figure 4.17. Drawdown in a confined aquifer at two locations using Theis’s method.

4.6.4 Some Potential Theory

4.6.4.1 Introduction
Recall the definition of piezometric head

p
h= +z
γ (4.6.58)

41
Which can be thought of as the energy or potential per unit weight of fluid. We can define a
“discharge potential” as

φ = Kh (4.6.59)

So that
!
q = −∇φ (4.6.60)

Or in components

∂φ ∂φ ∂φ
qx = − ; qy = − ; qz = −
∂x ∂y ∂z (4.6.61)
∂φ 1 ∂φ
qr = − ; qθ = −
∂r r ∂θ (4.6.62)

For steady, incompressible flow with no sources or sinks, we have


!
∇⋅q = 0 (4.6.63)

or

∇2φ = 0 Laplace' s Equation (4.6.64)

Functions satisfying Laplace’s equation are called harmonic functions.

Equipotentials are curves where φ is constant. The component of specific discharge in any
! !
direction l can be found by differentiating φ in the direction l

∂φ
ql = −
∂l (4.6.65)

If the specific discharge is parallel to an equipotential, then it can have no component tangent to
the equipotnetial. Therefore, the specific discharge must be normal to the equipotential lines,
which allows us to visualize the flow patterns. We can describe the flow patterns by a family of
curves which are tangent to the specific discharge called stream functions, ψ. Streamfunctions
are a family of curves which are everywhere tangent to the specific discharge. Streamlines are
the locus of all points for which the streamfunction has a constant value, i.e., ψ = Constant on a
streamline.

We can draw vectors at every point representing the discharge in the flow field. These vectors
are tangent to the streamlines. Streamlines are the locus of all points for which ψ is constant.

42
Let A be a fixed point and P be an arbitrary point. Suppose no sources or sinks are in the region
near A and P. Then the flux across any set of lines ACP depends only on P and time.

flux ACP = ψ ( x, y, t ) (4.6.66)

P
C
Flux

A
Figure 4.18. Flux in a region.

For two points P1 and P2 with streamfunction values ψ 1 and ψ 2

flux AP2 = flux AP1 + fluxP1 P2


(4.6.67)

hence

fluxP1P2 = ψ 2 −ψ 1
(4.6.68)

P2
P1

Flux

A
Figure 4.19. Flux across streamlines

Now, let P1 P2 = Δs, a small arc of a curve. Specific discharge across this curve is broken into
normal and tangential components; the tangential component contributes nothing to the flux
across Δs. The normal component is

flux across Δs ψ 2 −ψ 2 Δψ
qn = = =
Δs Δs Δs (4.6.69)

Or

43
∂ψ
qn =
∂s (4.6.70)

P2
qn
qs
P1
Δs
Figure 4.20. Normal and tangential components of specific discharge

P1
qn
Δ y Δs
qx qs
A
qy P2
Δx
x
Figure 4.21. Specific discharge in (x,y) and (s,n) coordinate systems.

So in Cartesian coordinates, we have

∂ψ
qx = −
∂y
∂ψ
qy =
∂x (4.6.71)

In radial coordinates

1 ∂ψ
qr = −
r ∂θ
∂ψ
qθ =
∂r (4.6.72)

Comparing the two definitions of specific discharge, we have

44
∂φ ∂ψ
qx = − =−
∂x ∂y
∂φ ∂ψ
qy = =
∂y ∂x (4.6.73)
Or
∂φ ∂ψ
− =−
∂x ∂y
∂φ ∂ψ
=
∂y ∂x (4.6.74)

These last equations are known as the Cauchy-Reimann Conditions.

Streamlines and equipotentials are orthogonal. At any point on an equipotential, φ = constant, so

∂φ ∂φ
dφ = dx + dy = 0
∂x ∂y (4.6.75)

− ∂φ
dy
= ∂x = − q x
dx ∂φ qy
φ =constant ∂y (4.6.76)

At any point on a streamline ψ = constant, so

− ∂ψ
dy
= ∂x = q y
dx ∂ψ qx
ψ =constant ∂y (4.6.77)

So that the slopes of the equipotential and streamlines are negative reciprocals of each other,
proving orthogonality.

45
s

! ψ2
t
t

y
! ds θ2
n dy
θ1
ψ1 dx x

n
Figure 4.22. Flow between two streamlines

! ! dy dx ˆ
dQ = q ⋅ nds = ( q x iˆ + q y ˆj ) ⋅ ( iˆ − j )ds
ds ds
= q x dy − q y dx
∂ψ ∂ψ
=− dy − dx
∂y ∂x
= − dψ (4.6.78)

2 2
∫ dQ = − ∫ dψ
1 1 (4.6.79)

Q1− 2 = ψ1 −ψ 2 (4.6.80)

4.6.4.2 Potentials for Different Aquifer Types


For a confined aquifer with homogeneous and isotropic properties experiencing steady flow, we
can write the continuity equation as

∇ ⋅ T ⋅ ∇h = 0 (4.6.81)

where

Q = −T ⋅ ∇h (4.6.82)

We can define a potential in this case as

46
φ = Th (4.6.83)

Then

Q = −∇φ and ∇ 2φ = 0 (4.6.84)

For an unconfined aquifer with homogeneous and isotropic properties experiencing steady flow,
we can write the continuity equation as

∇ ⋅ Kh ⋅ ∇h = 0 (4.6.85)

where

K
Q = − Kh ⋅ ∇h = − ∇h 2
2 (4.6.86)

We can define a potential in this case as

K 2
φ= h
2 (4.6.87)

Then

Q = −∇φ and ∇ 2φ = 0 (4.6.88)

4.6.4.3. Fundamental Singularities


Analytical solutions to potential flow problems can be obtained as combinations of 3
fundamental solutions. In two cases, the solutions are singular, that is Laplace’s equations is not
satisfied at a point called a “singular” point.

Linear flow: Streamlines are straight and parallel, and specific discharge is constant. In one-
dimension, we have

d 2φ
=0
dx 2 (4.6.89)


=A
dx (4.6.90)


qx = − ; A = −q x
dx (4.6.91)

47
φ ( x ) = −q x + B
(4.6.92)

where the constant B is determined from boundary conditions. The Cauchy-Reimann conditions
tell us that

dφ dψ dψ
= ; −q x =
dx dy dy (4.6.93)

ψ ( x ) = −q x y + C
(4.6.94)

where the constant C is determined from boundary conditions.

Radial flow: Streamlines are straight and radial to the origin, and specific discharge varies with
radial distance. In radial coordinates, we have

1 d ⎛ dφ ⎞
⎜ r ⎟ = 0
r dr ⎝ dr ⎠ (4.6.95)


r =A
dr (4.6.96)

φ ( r,θ ) = A ln( r ) + B = A ln x 2 + y 2 + B (4.6.97)

The Cauchy-Reimann conditions tell us that

∂φ 1 ∂ψ
=
∂r r ∂θ (4.6.98)


[A ln(r ) + B] = 1 ∂ψ
∂r r ∂θ

A 1 ∂ψ
=
r r ∂θ (4.6.99)

⎛ y ⎞
ψ ( r,θ ) = Aθ + C = A tan −1 ⎜ ⎟ + C
⎝ x ⎠ (4.6.100)

48
y

r y
θ
x
Well  pumping  
Equipotential
Q m3/s  at  the  
origin
Streamline

Figure 4.23. Well pumping at the origin.

Discharge vectors in the aquifer are everywhere directed toward the origin (-r direction).

Qr < 0
(4.6.101)

So that

Q = 2πr ( −Q r )
Q dφ A
Qr = − =− =−
2πr dr r
Q
A=
2π (4.6.102)

In a confined aquifer, we have

Q Q
φ = Th = ln r + B; ψ = θ +C
2π 2π
Q
h= ln r + B'
2πT
(4.6.103)
If h = h0 @ r = r0, then

Q
B' = h0 − ln r0
2πT (4.6.104)
And

49
Q ⎛ r ⎞
h − h0 = ln⎜ ⎟
2πT ⎜⎝ r0 ⎟⎠
(4.6.105)

Q ⎛ r0 ⎞
s( r ) = h0 − h = ln⎜ ⎟
2πT ⎜⎝ r ⎟⎠
(4.6.106)

In an unconfined aquifer, we have

K 2 Q Q
φ= h = ln r + B; ψ= θ +C
2 2π 2π
Q
h2 = ln r + B'
πK (4.6.107)

If h = h0 @ r = r0, then

Q
B' = h02 − ln r
πK 0 (4.6.108)
And

Q ⎛ r0 ⎞
h02 − h 2 = ln⎜ ⎟
πK ⎜⎝ r ⎟⎠
(4.6.109)

These equations still hold when the center of the well is located at the point (xo, y0) of the
coordinate system, but

r 2 = ( x − x0 ) 2 + ( y − y 0 ) 2
(4.6.110)

4.6.4.4. Superposition
If the governing equations and boundary conditions of a problem are linear, then solutions of the
equations can be superposed (added) to obtain new solutions. In this way, elementary solutions
can be put together to form other solutions which satisfy the boundary conditions of a particular
problem. Let φ1 and φ2 both satisfy a linear partial differential equation L(φ) = 0 (e.g., Laplace’s
equation or Poisson’s equation), then another solution of the problem is φ3 = Aφ1 + Bφ2, where A
and B are constants to be determined from the boundary conditions.

In a confined aquifer, if several wells are pumping simultaneously, we can extend the previous
equations to include any number of wells i = 1, 2, …, n located at points (xi, yi). Let hi be the
head at a point (x, y) due to pumping Qi from the ith well located at (xi, yi), then

50
Qi
hi =
2πT
ln (x − xi )2 + (y − yi )2 + Ai
(4.6.111)

The head h(x, y) due to the pumping of all n wells will be

n n Q
[( ) (
h( x, y ) = ∑ hi = ∑ i ln x − xi 2 + y − yi 2 + C
i =1 i =1 4πT
)]
(4.6.112)

Where C is a constant that must be determined form the boundary conditions.

Let s1 and s2 be the drawdowns due to two wells pumping independently, then the total
drawdown due to both wells pumping is the sum of the drawdowns due to the individual wells,
or s = s1 + s2. Superposition and symmetry can be used to calculate the effect of nearby wells,
aquifer property discontinuities or boundaries, recharge from streams and lakes, and intermittent
pumping.

4.6.4.5. Head distribution from Two Pumping Wells in a Confined Aquifer


[See De Weist p. 244]

[ADD Figure]
Figure 4.24. Two wells pumping in a confined aquifer

4.6.4.6. Capture Zone of a Well in Uniform Flow


The zone from which a well is drawing water in an aquifer is known as its “capture zone”. When
a well is in an aquifer with a uniform regional flow, we can calculate the extent of the capture
zone by superposing the solution for uniform flow with that for well flow in the aquifer. IN this
case, we have the equipotentials and streamlines as:

Q
φ = −Ux + ln( r ) + C
2πB (4.6.113)
Q
ψ = −Uy + θ + C'
2πB (4.6.114)

∂φ Q x
qx = − =U −
∂x 2π r 2 (4.6.115)

At the stagnation point, we have

51
q x = 0, x = x s , y = 0, r 2 = x 2
(4.6.116)

2πU 2 Q
xs = r ⇒ xs =
Q s 2πU (4.6.117)

Q
ψ = −U * 0 + * 0 + C' = 0 @ ( x s , y − 0)
2π (4.6.118)

So ψ = 0 on the groundwater divide

Q
ψ = 0 = −Uy + ϑ
2π (4.6.119)

Q
y= θ on the divide
2πU (4.6.120)

Q
as x → ∞,θ → π , y=
2U (4.6.121)

y
Groundwater  Divide
U
Q
Capture  Zone
2πU ( xs ,0)

Q x
Well

Stagnation  
Groundwater  Divide point

Figure 4.25. Capture Zone of a Well

Often when groundwater contamination problems are discovered at a site, some kind of
groundwater extraction and treatment system is designed to remediate the problem. Various
techniques can be used for containing and remediating the problem including source control –
the prevention of continued release of contaminants to the groundwater. Sources of
contamination often include landfills and leaks of chemicals from old tanks or spills. A feasible
goal of the remediation may be to halt any further spread of the contamination that has already

52
entered the aquifer. Methods that are used to achieve this goal revolve around containment of
contaminants to prevent spreading. Hydrodynamic isolation (or hydraulic control) is used to
isolate the zone where groundwater has become contaminated. Capture zones are useful in this
regard. After a sufficient time of pumping, steady-state flow conditions will exist in the aquifer.
A drawdown cone will have developed and the regional flow lines will have been changed to
accommodate the pumping regime. Water will be drawn into the well form the “capture zone”
or that portion of the aquifer that contains groundwater that will eventually be captured by the
well.

Design a pumping well network in a homogeneous, isotropic, horizontal, confined aquifer in


which the preoperational, natural flow field is uniform and steady. Note that the potential and
stream function derived for a single well can be extended to multiple wells in a remediation
system:

n Qi
φ ( x, y ) = −Ux + ∑ ln ( x − xi ) 2 + ( y − yi ) 2
i =1 2πB (4.6.122)
n Q ⎛ y − yi ⎞
ψ ( x, y ) = −Uy + ∑ tan −1 ⎜ ⎟
i =1 2πB ⎜ x − x
⎝ i
⎟
⎠ (4.6.123)

These functions can be used to map the head and flow lines for a large number of points for any
set of n well locations and pumping rates. The overall discharge will be

∂φ n Q x
qx = − =U − ∑ i
∂x 2 2
i =1 2π ( x − xi ) ( y − yi )
(4.6.124)

4.6.4.7. Steady Drawdown in a Confined Aquifer


The following example is adapted from Mays and Tung (1992, prob. 8.4.3). Consider an
undeveloped, homogeneous, isotropic, confined aquifer of large areal extent in which there are L
potential production wells located at points [( xℓ , yℓ ),ℓ = 1,..., L] and K points at which water
levels are to be observed [(xk , yk ),k = 1,..., K ] . For steady-state flow in the aquifer, the Theim
equation can be used to predict the drawdown sk at the control points due to the pumping Qℓ at
the production wells

L L
s k = ∑ s kℓ = ∑ akℓ Qℓ k = 1,!, K (4.6.125)
ℓ =1 ℓ =1

where skℓ is the drawdown at point k due to pumping at well ℓ alone, and

1 ⎛ R ⎞
akℓ = ln⎜⎜ ℓ ⎟⎟ k = 1, !, K ; ℓ = 1, !, L (4.6.126)
2πT ⎝ rkℓ ⎠

53
is the response function or unit drawdown at control point k due to unit pumping at production
well ℓ, T is the transmissivity of the aquifer, Rℓ is the radius of influence of the ℓ -th production
well, and the distance from the ℓ-th well to the k-th “control” point is

rkℓ = (xk − xℓ )2 + ( yk − yℓ )2 (4.6.127)

A management model for the optimal development of the aquifer can be formulated. The
objective of the model is to maximize the total withdrawal from the aquifer while meeting
specified upper bound constraints on the drawdown at the control points and upper and lower
bounds on the pumping rates. The management model is expressed as

L
Maximize ∑ Qℓ
ℓ =1
Subject to
L
* (4.6.128)
∑ a kℓ Qℓ = s k ≤ s k k = 1, ! , K
ℓ =1

Qℓlow ≤ Qℓ ≤ Qℓup ℓ = 1, ! , L

This management model is a linear program where the decision variables are the unknown
production rates, Qℓ , ℓ = 1,!, L at each well. Consider the situation (shown in the Figure) of
three production wells (L=3) and five control points (K=5) a transmissivity of 5,000 gal/day/ft2,
and a radius of influence for each well of 700 ft. Additional data related to the problem are listed
in the Table and the response coefficients akℓ are listed in the Table.

200 ft 800 ft

s1 s2
200 ft
Q1 Q2

s3 800 ft

Q3

s4 s5

Control point Well

Figure 4.26. Aquifer for steady-state pumping optimization example.

54
Table 4.6. Distance (ft.) between well ℓ and control point k.

rk,ℓ Production
Control Point Capacity,
Well k=1 2 3 4 5 Qℓup
(gal/day)
ℓ= 1 282.8 632.5 282.8 848.5 632.5 200,000
2 632.5 282.8 282.8 632.5 848.5 200,000
3 721.1 721.1 200 565.7 565.7 200,000
Max. Allowable
Drawdown, sk* 7 7 15 7 7
(ft)

Table 4.7. Response matrix A (x10-5) for 5 control points and 3 production wells.

ak,ℓ Well
Control Pt. ℓ=1 2 3
k= 1 2.885x10-5 3.228 x10-6 0
2 3.228 x10-6 2.885x10-5 0
3 2.885x10-5 2.885x10-5 3.988 x10-5
4 0 3.228 x10-6 6.781 x10-6
5 3.228 x10-6 0 6.781 x10-6

The optimization model is written as

Maximize Q1 + Q2 + Q3
Subject to
2.885x10 − 5 Q1 + 3.228 x10 − 6 Q2 ≤7
3.228 x10 − 6 Q1 + 2.885x10 − 5 Q2 ≤7
2.885x10 − 5 Q1 + 2.885x10 − 5 Q2 + 3.988 x10 − 5 Q3 ≤ 15 (4.6.129)
+3.228 x10 − 6 Q2 + 6.781x10 − 6 Q3 ≤7
3.228 x10 − 6 Q1 + 6.781x10 − 6 Q3 ≤7
0≤Q1 ≤ 200,000
0≤Q2 ≤ 200,000
0≤Q3 ≤ 200,000

4.6.4.8. Transient Radial Flow in a Confined Aquifer


The drawdown at a point in an aquifer resulting from multiple wells pumping is predicted by
summing the drawdowns at that point due to the individual wells pumping alone. Consider the
problem of trying to predict the drawdown at a control point ( xk , y k ) in a homogeneous,

55
isotropic aquifer of large extent due to a well located at the point ( xℓ , yℓ ) the pumping
schedule shown in the following Figure.

Pumping rate

Q2

Q1
Q3

t1 t2 t3 Time

Figure 4.27. Pumping schedule.

For the time period 0 ≤ t ≤ t1 , we have

Qℓ1
sk ( t ) = W (u1 ) (4.6.130)
4πT

where
rk2ℓ S
u1 = (4.6.131)
4Tt

and
rkℓ = (xk − xℓ )2 + ( yk − yℓ )2 (4.6.132)

For t1 ≤ t ≤ t 2 we have

Qℓ1 Q 2 − Qℓ1
s k (t ) = W (u1 ) + ℓ W (u 2 ) (4.6.133)
4πT 4πT

where

rk2ℓ S
u2 = (4.6.134)
4T (t − t1 )

Finally, for t 2 ≤ t ≤ t 3

56
Qℓ1 Q 2 − Qℓ1 Q 3 − Qℓ2
s k (t ) = W (u1 ) + ℓ W (u 2 ) + ℓ W (u3 ) (4.6.135)
4πT 4πT 4πT

where
rk2ℓ S
u3 = (4.6.136)
4T (t − t 2 )

When a well pumps Qℓi during the period from ti-1 to ti i=1,…,n where ti-1 - ti = Δt, we can write
the drawdown at point ( x k , y k ) at time tn as (Bear, 1979, p. 325)

⎡
1 n i ⎢ ⎛⎜ Srk2ℓ ⎞ ⎛
⎟ ⎜ Srk2ℓ ⎞⎤
⎟⎥
s kℓ (t n ) = ∑ Qℓ W ⎜ −W
4πT i =1 ⎢ 4T (n − i + 1)Δt ⎟ ⎜ 4T (n − i )Δt ⎟⎥
⎣ ⎝ ⎠ ⎝ ⎠⎦ (4.6.137)
n
= ∑ Qℓi a kℓ (n − i )
i =1

where
⎡
1 ⎢ ⎛⎜ Srk2ℓ ⎞ ⎛
⎟ ⎜ Srk2ℓ ⎞⎤
⎟⎥
akℓ (n − i) = W −W (4.6.138)
⎜
4πT ⎢ 4T (n − i + 1)Δt ⎟ ⎜ 4T (n − i)Δt ⎟⎥
⎣ ⎝ ⎠ ⎝ ⎠⎦

These coefficients are variously termed influence coefficients, algebraic technological function,
response coefficients and response functions. When L wells pump at rates Qℓi (ℓ=1,…,L) during
the period from ti-1 to ti and ti-1 - ti = Δt, we can write the drawdown at point ( x k , y k ) at time tn
as (Bear, 1979, p. 499)

L L n
s k (n) = ∑ s kℓ (t n )= ∑ ∑ Qℓi akℓ (n − i ) (4.6.139)
ℓ =1 ℓ =1 i =1

and the distance from the ℓ-th well to the k-th “control” point is

rkℓ = (xk − xℓ )2 + ( yk − yℓ )2 (4.6.140)

4.6.4.9. Minimum drawdown in a confined aquifer


For transient management the Theis equation can be used to predict the drawdown sk (n) at the
control point ( x k , y k ) during time period n due to the pumping Qℓi at production well ℓ during
time period i. Based on the expression for drawdown at various control points in the aquifer, a

57
management model for the optimal operation of the aquifer over the planning horizon (i =
1,…,N) can be formulated. Consider the objective of the model to minimize the drawdown in
the aquifer at the control points (k = 1,…,K) while satisfying a time variable schedule of future
demands (Di, i = 1,…,N). The management model is expressed as

N K
Minimize ∑ ∑ ski
i =1k =1
Subject to
L n
i * (4.6.141)
∑ ∑ akℓ ( n − i )Qℓ =sk ( n )≤sk k = 1,! K; n = 1,! ,N
ℓ =1 i =1
L
i
∑ Qℓ ≥ Di i = 1,! ,N
ℓ =1
Qℓlow ≤ Qℓi ≤ Qℓup ℓ = 1,! ,L; i = 1,! ,N

This management model is a linear program where the decision variables are the unknown
production rates Qℓi at each well for each time period i = 1,…,N.

Recall the example from the previous section with three production wells (L=3) and five
control points (K=5). Consider now that the aquifer is to be managed for three periods of fifty
days each. The storage coefficient for the aquifer is 0.002, and the maximum permissible
drawdowns at each of the control points are given in the following Table. The transmissivity in
this case is 5,000 gal/day/ft2, and the radius of influence of each well is 700 ft. Additional data
related to the problem are listed in the following Table and the response coefficients are listed in
the following Table for the different management periods.

Table 4.8. Permissible drawdowns (feet) at 5 control points in 3 management periods.

sk* (n) Control point k


Period (days) 1 2 3 4 5
1 (0 - 50) 5 5 8 5 5
2 (51 - 100) 8 8 10 8 8
3 (101 - 150) 10 10 15 10 10

58
Table 4.9. Response coefficients between 5 control points and 3 production wells for 3
management periods of fifty days each.
r
(a) Response matrix a ni (x 10-4) for n-i=0 [(n,i) = (1,1), (2,2), (3,3)]

akℓ (n − i) Well (ℓ)

Control Pt. (k) 1 2 3


1 1.485 1.109 1.154
2 1.205 1.332 1.154
3 1.485 1.290 1.410
4 1.332 1.135 1.410
5 1.166 1.485 1.410
r
(b) Response matrix a ni (x 10-5) for n-i=1 [(n,i) = (2,1), (3,2)]

akℓ (n − i) Well (ℓ)

Control Pt. (k) 1 2 3


1 1.103 1.103 1.103
2 1.103 1.103 1.103
3 1.103 1.103 1.103
4 1.103 1.103 1.103
5 1.103 1.103 1.103
r
(c) Response matrix a ni (x 10-6) for n-i=2 [(n,i) = (3,1)]

akℓ (n − i) Well (ℓ)

Control Pt. (k) 1 2 3


1 6.453 6.453 6.453
2 6.453 6.453 6.453
3 6.453 6.453 6.453
4 6.453 6.453 6.453
5 6.453 6.453 6.453

4.7. Horizontal Flow in Unconfined Aquifers


4.7.1 Introduction
Unconfined aquifers are bounded above by a phreatic surface where p = patm = 0 (gage). The
solution to problems of flow in phreatic aquifers is complicated by the presence of a free surface
boundary condition (the water table). The position of this nonlinear surface is unknown and

59
must be solved for by some means. The Dupuit approximations allows us to derive solutions to
unconfined flow problems where the flow domain is bounded above by a phreatic surface (water
table) and the capillary fringe is much smaller than the saturated thickness of the aquifer below.
Dupuit (1863) made some simplifying assumptions based on the observation that the slope of the
phreatic surface is very small in most groundwater problems. Referring to Figure 3 below, the
specific discharge along the phreatic surface (coordinate direction s) is given by Darcy’s law as

∂h
qs = −K
∂s
∂ ⎛ p ⎞
= − K ⎜⎜ + z ⎟⎟
∂s ⎝ γ ⎠ (4.7.1)
∂z
= −K
∂s
= − K sin θ

since h = p/γ+z, p = 0 on the phreatic surface, θ is the angle whose tangent is the slope of the
phreatic surface and ∂z/∂s = sinθ. For small angles we can use the approximation sinθ ≈ tanθ
=∂z/∂x. Thus the specific discharge in the x direction can be expressed as

∂z ∂h
q x = −K = −K
∂x ∂x (4.7.2)

These assumptions are equivalent to:

(1) flowlines are horizontal and equipotentials (lines of constant head) are vertical, (h =
h(x,y) independent of z);

∂h
(2) the hydraulic gradient, , is equal to the slope of the phreatic surface, tan(θ), and
∂x
invariant with depth (not a function of z); and

(3) velocities are horizontal (no vertical flow component) and parallel to each other along
the same vertical line.

60
θ Δz
dx
θ
dz =
ds
dh s
h
Δx

ground
surface
water
table
unconfine
aquife
d
r
h
Qx K
z
y
x
bedroc
k Δx
Figure 4.28. Unconfined aquifer and free surface.

Dupuit approximation
constant head line Actual constant
head line Actual water table

ground surface
Dupuit approximation
water table

seepage face
hL
h
hR

bedrock
Dupuit approximation Actual velocity
velocity distribution distribution

Figure 4.29. Dupuit assumptions for an unconfined aquifer.

Using these assumptions, the discharge per unit width across any vertical section through the
point x where the saturated thickness is h is

61
∂h ∂(h 2 / 2)
Q x = (− K )h = − K (4.7.3)
∂x ∂x

and
∂h ∂ ( h 2 / 2)
Q y = (− K )h = − K (4.7.4)
∂y ∂y

or, in vector notation

Q = − Kh∇h = − K∇(h 2 / 2) (4.7.5)

The important thing to notice here is that the number of independent variables has been reduced
by one, from (x,y,z,t) to (x,y,t), since z no longer appears as an independent variable. The Dupuit
assumptions actually amount to neglecting the vertical flow component qz, see Figure 4, and are
generally valid wherever the lengths of interest in the direction of flow are much larger (1.5 to 2
times) than the saturated thickness of the aquifer [Bear, 1972, sec 8.1.1].

4.7.2 Unconfined Groundwater Flow Equations


Consider a control volume or REV of area ΔxΔy and water table height h centered around point
P of unconfined aquifer shown in Figure 5. Let Q be the discharge per unit width of aquifer at
the point P. The principle of conservation of mass for the REV requires that for a small time
period:

mass flux in - mass flux out = change of mass storage

The mass flux per unit time at the point P is ρQ. The mass flux through the sides of the REV
can be found from a Taylor series expansion of the mass flux at point P. For example, the mass
flux per unit area through the side at x+Δx/2 is

∂ ( ρQx ) Δx 1 ∂ 2 (ρQx ) $ Δx & 2


% 2 ' +!
ρQx x +Δx / 2 = ρ Qx x + + (4.7.6)
∂x x 2 2 ∂x 2
x

where variables and derivatives on the right hand side are evaluated at the point (x,y,z).

62
Figure 4.30. Control volume for an unconfined aquifer.

The Taylor series is truncated after the second (linear) term and higher order terms (involving
Δx2) are neglected, i.e., a local approximation.

∂ (ρQ x ) Δx
ρQ x x + Δx / 2
≈ ρQ x x + (4.7.8)
∂x x 2

Similar expressions can be developed for the mass flux across the side at x-Δx/2 and in the y and
z directions. Contributions to the mass flux into the REV over a short time interval Δt are:

⎡ ∂ (ρQ x ) Δx ⎤
x direction: ⎢ ρQ x − ∂x ΔyΔt
⎣ 2 ⎥⎦

y direction:
⎡ ( )
∂ ρQ y Δy ⎤
⎢ ρQ y − ⎥ ΔxΔt
⎣ ∂y 2 ⎦
z direction: [R + N ]ΔxΔyΔt
(4.7.9)

where R and N are the contributions from artificial and natural recharge, respectively.

63
Contributions to the mass flux out of the REV over a short time interval Δt are:

⎡ ∂Qx Δx⎤
x direction: ⎢Qx + ⎥ ΔyΔt
⎣ ∂x 2 ⎦

⎡ ∂Qy Δy⎤
y direction: ⎢Qy + ⎥ ΔxΔt
⎣ ∂y 2 ⎦

z direction: [ - P + qz] ΔxΔyΔt


(4.7.10)

where P and qz are the contributions from pumping from the aquifer and leakage through the
bottom of the aquifer, respectively. Adding all of the contributions to mass flux into and
subtracting all of the contributions to mass flux out of the REV results in

net mass flux = mass flux in - mass flux out

⎡ ∂Q x ∂Q y ⎤
= ⎢− − + R + N − P + q z ⎥ ΔxΔyΔt (4.7.11)
⎣ ∂x ∂y ⎦

The storativity of unconfined aquifers (also known as the specific yield) is defined as

Vw
Sy =
AΔh (4.7.12)

where Δh = h(x,y,t+Δt) - h(x,y,t) is a change (decline or rise) in the water table over the time
period Δt. Water is either drained from or added to the volume of pore space as a result of the
change Δh. The excess of inflow over outflow in the control volume goes into storage in the
unconfined aquifer as:

(1) a rise in the phreatic surface by Δh so that a volume of porous medium ΔxΔyΔh stores
or releases an amount of water SyΔxΔyΔh;

(2) a rise in the pressure in the aquifer due to compressibility of the porous medium and
water. This effect is negligible in unconfined aquifers due to the lack of a confining
stratum so that release from storage is due primarily to drainage from the pore space.

The change in the mass of water in the REV over a time Δt is

∂h
Sy dxdydt (4.7.13)
∂t

Combining these equations, we have

64
∂Q x ∂Q y ∂h
− − _ R + N − P + qz = S y (4.7.14)
∂x ∂y ∂t

Introducing the expressions for the discharge, we obtain Boussinesque’s equation for two-
dimensional unsteady flow in an unconfined aquifer

∂ ⎛ ∂h ⎞ ∂ ⎛ ∂h ⎞ ∂h
⎜ K x h ⎟ + ⎜ K x h ⎟ + R + N − P + q z = S y (4.7.15)
∂x ⎝ ∂x ⎠ ∂x ⎝ ∂x ⎠ ∂t

or

∂h
∇K∇h + R + N − P + q z = S y (4.7.16)
∂t

For a homogeneous, isotropic porous medium the equation becomes

K ⎡⎛ ∂ 2 h 2 ⎞ ⎛ ∂ 2 h 2 ⎞⎤
⎢⎜ ⎟ + ⎜ ⎟⎥ + R + N − P + q z = S y ∂h (4.7.17)
2 2 ⎟ ⎜ 2
⎢⎣⎝ ∂x ⎠ ⎝ ∂y ⎟⎠⎥⎦
⎜ ∂t

4.7.3 Examples

4.7.3.1. Flow in an Unconfined Aquifer


Consider the steady-state Boussinesque equation for flow in an unconfined aquifer with
infiltration shown in Figure 8. The governing equation is

∇ ⋅ Kh∇h + N = 0 (4.7.17)

For one-dimensional flow in a homogeneous, isotropic aquifer this equation becomes

K d 2 (h 2 )
+ N =0 (4.7.18)
2 dx 2

Integrating this equation twice and applying the boundary conditions shown in Figure 8 results in
an expression for the head at any position x between the two streams

N h2 − h2 N
h 2 ( x) = hL2 + ( L + R L )x − x2 (4.7.19)
K L K

The discharge at any point x is

65
K d (h 2 )
Q( x) = −
2 dx (4.7.20)
K N hR2 − hL2
=− ( L+ ) + Nx
2 K L

The groundwater divide occurs at the point of no flow or where

d (h 2 )
=0 (4.7.21)
dx

or
L K
xdivide = + (hR2 − hL2 ) (4.7.22)
2 2 NL

Thus if hL > hR the divide moves to the left, and if hL < hR then the divide shifts right.

N, infiltration

ground surface

water table

h h m ax Flow
hL hR

bedrock
L
x

Figure 4.31. Uniform infiltration and drainage from an unconfined aquifer to streams.

Consider the case when L = 3000 m, K = 20 m/day, hL = 30 m, hR = 20 m and N = 500 mm/yr.


Determine the rates of flow to the two streams and the shape of the water table. The
groundwater divide occurs at x = 300 m, and the discharge to the left-hand stream is QL = -0.39
m3/day and the discharge to the right-hand stream is QR = 3.72 m3/day. The head values are
listed in Table 10.

66
Table 4.10. Head distribution for flow between two streams in an unconfined aquifer.

X h x h
0.00 30.00 1600.00 28.05
100.00 30.05 1700.00 27.71
200.00 30.08 1800.00 27.35
300.00 30.09 1900.00 26.95
400.00 30.08 2000.00 26.53
500.00 30.04 2100.00 26.07
600.00 29.98 2200.00 25.57
700.00 29.89 2300.00 25.04
800.00 29.79 2400.00 24.47
900.00 29.66 2500.00 23.85
1000.00 29.50 2600.00 23.19
1100.00 29.32 2700.00 22.48
1200.00 29.12 2800.00 21.72
1300.00 28.89 2900.00 20.89
1400.00 28.64 3000.00 20.00
1500.00 28.36

Figure 4.32. Flow between two streams in an unconfined aquifer.

4.7.3.2. Drains
[Check McWhorter and Sunada for a numerical example]
The figure below illustrates the placement of two drains in the subsurface. Several design
variables are evident in the figure: the height of the drains above the lower confining layer (hd),
the distance between the drains (L), and the maximum allowable height of the water above the
drains (hMax). In some cases, the heights may be referenced to the ground surface rather than the

67
distance above the hardpan.
N, Infiltration

Ground Surface
Water Table

hMax

hd

Lower
L
x Confining
Layer
Figure 4.33. Drains in the subsurface.

Assuming steady flow and homogeneous, isotropic porous media properties, the flow equation
for unconfined aquifers is

! !! !!
!!
Kh !! + N = S! !! (4.7.23)

or for steady-state flow

!! ! ! !
!!!
= −2 ! (4.7.24)

Integrating twice and applying boundary conditions, we have


!"
ℎ! = !
𝐿 − 𝑥 + ℎ!! (4.7.25)

Thus the maximum height of water above the lower confining layer is in the middle between the
drains, or

!!! !
!
ℎ!"# = ℎ! = !!
+ ℎ!!      𝑎𝑡      𝑥 = ! (4.7.26)

4.3.3.3 Steady Radial Flow in a Unconfined Aquifer


Consider a pumping well in a homogeneous, isotropic, unconfined aquifer of infinite extent
shown in the following Figure. The governing equation for flow in this aquifer is

68
!"
ℎ! = !
𝐿 − 𝑥 + ℎ!! (4.7.27)

Which for steady state becomes

NL2
2
hMax = [h ( L / 2)]2 = + hd2
4K
NL2
hMax = + hd2
4K  
(4.7.28)  

Figure 34. Drawdown due to pumping in a unconfined aquifer.

Let’s write this in radial coordinates using radial symmetry (isotropic medium)

! !! !
!"
𝑟 !"
=0 (4.7.29)

Integrating, we have

!! !
𝑟 !"
= C! (4.7.30)

Now, by Darcy’s Law, we have

!! !! !
𝑄 = 𝐴𝑞 = 2𝜋𝑟ℎ 𝐾 !" = 𝜋𝑟𝐾 !"
(4.7.31)

69
So
! !"
𝑑ℎ! = !" !
(4.7.32)

and we have

!
ℎ! = !" ln 𝑟 + 𝐶! (4.7.33)

So if h = h0 at r = R, we have

! !
ℎ!! − ℎ! = !" ln !
(4.7.34)

[Add figure]
Figure 4.35. Drawdown due to pumping in a unconfined aquifer.

4.8 Modeling Groundwater Systems


4.8.1 Methods for Solving Groundwater Problems
Various tools have been used to solve groundwater problems for decades. Physical and analog
methods were some of the first methods used. These methods, while very accurate and useful for
visualization, are somewhat cumbersome to adapt to various and different boundary conditions
and formation layouts. Analytical methods are the technique we have been discussing so far.

Analytical solutions are difficult for situations where you have irregular boundaries, various
different boundary conditions in a region, heterogeneous and anisotropic porous media
properties, or multiple phases. Often the aquifers and subsurface environments we want to
model exhibit all of these characteristics. So what can we do? Numerical solutions to the
governing equations of flow and mass transport in porous media provide a useful and convenient
method of handling these complications. In this section, we will discuss the most common
technique used to model groundwater flow and mass transport, the finite-difference method. In
many cases, we can derive the analytical solution to the equations governing the flow of
groundwater. However, in many cases, this becomes impractical due to the size of the region
being modeled, the desired detail of the solution, the heterogeneity of the aquifer properties, or
the nonlinearity of the equations. In this case, we can try to use numerical methods.

Typical approaches to determining the response of a multi-dimensional, transient groundwater


system involves transferring the Partial Differential Equations governing the flow of
groundwater into a system of Ordinary differential Equations or algebraic equations.

The dependent variables in the governing PDEs are known as state variables. For groundwater
flow problems, these consist of head and specific discharge. A solution to the governing PDE
determines the state variables at a set of discrete note points within the aquifer.

70
Finite difference methods are based on a Taylor series representation of the derivatives appearing
in the governing equations. Consider the equation governing horizontal flow in a leaky confined
aquifer.

! !! ! !! ! !! !! ! !!
!"
𝑇! !" + !" 𝑇! !" + !" 𝑇! !" + !! ℎ! − ℎ ± !!! 𝑄! 𝛿 𝑥 − 𝑥! = 𝑆 !" (4.8.1)

We want to solve this equation numerically, subject to appropriate initial and boundary
conditions.

4.8.2 Classification of Partial Differential Equations


Mathematicians divide PDEs into three classes: Elliptical, Parabolic and Hyperbolic PDEs.

Elliptical PDEs – Elliptical PDEs are steady state equations in two or three dimensions,
including such famous examples as Poisson’s equation

!! ! !! !
!! !
+ !! ! = 𝑓 (4.8.2)

or Laplace’s equation (Poisson’s equation with f = 0). Properties of elliptical PDEs include
closed solution regions; boundary condition information is needed for each point on the
boundary of the solution region:

• Dirichlet boundary condition: known dependent variable, e.g., h


• Neuman boundary condition: known derivative of dependent variable

A change in the conditions interior to the solution region cause changes at all parts of the region.
Examples include: membrane stretching, steady state temperature distribution, subsonic fluid
flow.

Parabolic PDEs – Parabolic PDEs are time dependent equations in two or three dimensions,
including such famous examples as heat diffusion equation

!! !! !
!"
= !! ! + 𝑓 (4.8.3)

Properties of parabolic PDEs include an open solution region (due to the time dimension); initial
condition information is needed for the entire solution region at the initial time; boundary
condition information is needed for each point on the boundary of the solution region:

• Dirichlet boundary condition: known dependent variable, e.g., h


• Neuman boundary condition: known derivative of dependent variable ∂h / ∂n

A change in the conditions interior to the solution region causes changes in the region only at

71
later times. Examples include: unsteady heat flow, molecular diffusion, boundary layers in fluid
flow.

Hyperbolic PDEs – Hyperbolic PDEs are time dependent equations in two or three dimensions,
including such famous examples as wave equation

!! ! !! !
𝑐 ! !! ! = !! !
+𝑓 (4.8.4)

Properties of hyperbolic PDEs include an open solution region (due to the time dimension); two
sets of initial condition information are needed for the entire solution region at the initial time;
boundary condition information is needed for each point on the boundary of the solution region:

• Dirichlet boundary condition: known dependent variable, e.g., h


• Neuman boundary condition: known derivative of dependent variable ∂h / ∂n

A change in the conditions interior to the solution region causes changes in the region only at
later times. Examples include: vibrations, wave motion, unsteady or supersonic fluid flow.

4.8.3 Numerical Differentiation


The method of solution for groundwater problems that will be illustrated here is the finite-
difference method. Other methods that might be considered include the finite element method
and the finite volume method. The finite difference method is based on replacing partial
derivatives in the governing equations with numerical approximations and solving a resulting set
of linear or nonlinear algebraic equations. As an introduction to this method, we will review the
development of finite-difference formulas for derivatives.

The forward Taylor series expansion of f(x) in the neighborhood of a point x is

!" !! !! ! !! !! ! !! !! !
𝑓 𝑥 + ℎ = 𝑓 𝑥 + ℎ !" + !! !! !
+ !! !! !
+ !! !! !
+⋯ (4.8.5)

where h is a small number. The backward Taylor series expansion of f(x) about x can be written
as

!" !! !! ! !! !! ! !! !! !
𝑓 𝑥 − ℎ = 𝑓 𝑥 − ℎ !" + !! !! !
− !! !! !
+ !! !! !
−⋯ (4.8.6)

Solving the forward equation for the first derivative, f’(x), we have

!" ! !!! !! ! ! !!! !! !! !


!"
= !
− !! !! ! − !! !! !
−⋯ (4.8.7)

or

72
!" ! !!! !! !
!"
= !
+ 𝜗(ℎ) (4.8.8)

!" !!!! !!!


!" !!
= !
+ 𝜗(ℎ) (4.8.9)

h(x)

hi−1 hi+1

hi

€ Δx € Δx
x

i −1 i i +1
Figure 4.36. Forward, backward and central finite-difference approximations.

The error in this forward finite-difference approximation of the first derivative is


€ € €
! !!!
𝜗 ℎ = − !! !! ! (4.8.10)

where. Similarly, we can solve the backward equation for the first derivative and we get

!" ! ! !! !!! !! !! ! !! !! ! !! !! !
!"
= !
+ !! !! !
− !! !! !
+ !! !! !
−⋯ (4.8.11)
or
!" !! !!!!!
!" !!
= !
+ 𝜗(ℎ) (4.8.12)

which is a backward finite-difference approximation of the first derivative and it has the same
(first-order) error as the forward approximation.

An expression for the second derivative can be obtained by adding the forward and backward
Taylor series expansions

!! !! ! !! !! !
𝑓 𝑥 + ℎ + 𝑓 𝑥 − ℎ = 2𝑓 𝑥 + 2 !! !! ! + 2 !! !! ! + ⋯ (4.8.13)

or, solving for f”(x)

!!! ! !!! !!! ! !! !!! !! !! !


!! !
= !!
− 2 !! !! ! − ⋯ (4.8.14)
!!! !!!! !!!! !!!!!
!! !
= !!
+ 𝜗(ℎ! ) (4.8.15)

73
which is the central finite-difference approximation for the second derivative.

4.8.4 Grids and Discretization

4.8.4.1. Introduction
The finite-difference method is used to approximate the governing equations (PDEs) of
groundwater flow. With the finite-difference method the governing partial differential equation
(PDE) is replaced by a numerical approximation. In this process the continuous derivatives of
the PDE are replaced by discrete approximations. The result of this discretization is a set of
simultaneous algebraic equations that must be solved for the values of the unknown variables at
discrete locations in the modeled domain. In order to accomplish the discretization process a
mesh or grid must be defined that covers the domain. The grid consists of a series of
intersecting, orthogonal, straight lines. The unknown state variable(s) are then solved for at
either the intersections of the gridlines (mesh-centered) or at the centers of the gridblocks (block-
centered).

y, j

Mesh

Domain

i,j+1

Δy
i-1,j i,j i+1,j

i,j-1 Node point

x, i
Δx
Δy

Δx
Grid cell€
Figure 4.37. Finite-difference mesh.

Aquifer properties and state variables (e.g., head,


€ velocity, concentration, etc.) are considered to

74
be constant within each grid cell in the mesh. The goal of the finite-difference method is to
predict the values of the state variables at the node points of the mesh, i.e., so as to determine the
effects of pumping, the flow from a river due to increased pumping, etc. Finite-difference
methods are popular due to simplicity of the approximation process and the ease of solution.
They are attractive for problems with simple geometry where node points can be aligned with a
regular grid.

Consider the grid with uniform spacing Δx and Δy indexed in two-dimensions by the integers i
and j, respectively. Intersections define a node point at which a dependent variable f is defined.
Let f(xi, yj) = fi,j where xi, yj represent the x and y coordinates of the note point (i,j), xi = iΔx and yj
= jΔy.

4.8.4.2. Consistency
In finite-difference methods, we must require that as the node spacing if the grid becomes small,
the finite difference expression becomes arbitrarily close to the derivative expressions. That is,
in the limit of small spacing, the truncation error vanishes. This consistency requirement defines
a relationship between the continuous PDE and the discrete, finite-difference approximation.

4.8.4.3. Order of Approximation


The rate at which the truncation error approached zero as the grid spacing is decreased is very
important in finite-difference modeling. For an rth-order scheme: if Δx is decreased by a factor of
2, then the truncation error decreases by a factor of 2r.

4.8.5 One-Dimensional Flow

4.8.5.1 Steady-state flow

Consider steady, one-dimensional flow in a confined aquifer shown in Figure 4.36. The
governing equation for one-dimensional flow is

! !! ! !!
!"
𝑇! !" ± !!! 𝑄! 𝛿 𝑥 − 𝑥! = 𝑆 !" (4.8.16)

or for steady-state flow

! !! !
!"
𝑇! !" ± !!! 𝑄! 𝛿 𝑥 − 𝑥! = 0 (4.8.17)

where T [L2/T] is the transmissivity of the aquifer and Q [L3/T/L2 = L/T] is the pumping rate per
unit area of aquifer.

75
Ground surface

Confining Layer
hA

Node Aquifer Δx
hB
b
z! y!
x!

i= 0 1 2 3 4 5 6 7 8 9 10
Grid Cell
Figure 4.38. One-dimensional example aquifer.

When the transmissivity is homogeneous, and we apply a finite-difference approximation to the


second derivative, we have
!! !!
!! ! !!
!" !!!/! !" !!!/!
∆!
− 𝑄! = 0 (4.8.18)

!!!/! !!!! !!! !!!/! !! !!!!!


!! !!!
∆! ∆!
∆!
− 𝑄! = 0 (4.8.19)

where Qi is the source/sink rate for a particular node or cell. This finite-difference equation must be written for each
node i where the head is unknown (nodes 1 – 4). The boundary conditions are: h(x = 0,t) and h(x = L,t). We can
write this as
 
𝐴! ℎ!!! − ℎ! + 𝐵! ℎ! − ℎ!!! − 𝑄! = 0 (4.8.20)  

where
! !!
𝐴! = ∆! ! , 𝐵! = ∆! ! (4.8.21)  

4.8.5.2. Simulation of Steady Flow in a 1-D Homogeneous System


(ADD THIS)

4.8.5.3. Simulation of Transient Flow in a 1-D Homogeneous System


Consider the situation shown below where the conductivity of the homogeneous, isotropic,

76
confined aquifer is (K = 0.5 m/day, and the thickness is b = 1.5 m (transmissivity T = 0.5*1.5 =
0.75 m2/day). The heads on the left- and right-hand boundaries are constant values of hA = 6.1 m
and hB = 1.5 m, respectively. The mesh spading is Δx = 1 m and the initial head across the
aquifer is h(x, 0) = 6.1 m. The storage coefficient is S = 0.02.

z
hB

Ground surface

y
7 8 9 1
Aquifer

Confining
b Layer
hA Δx i= 0 1 2 3 4 5 6

Figure 4.39. Vertical cross-section through example aquifer.

t, l
i, l + 1

Δt
i − 1, l i + 1, l
i, l
x, i
Δx

i, l − 1

Figure 4.40. Space (horizontal axis) and time (vertical axis) domain for one-dimensional
transient flow.

The governing equation for flow in the aquifer is

77
! !! ! !!
!"
𝑇! !" ± !!! 𝑄! 𝛿 𝑥 − 𝑥! = 𝑆 !" (4.8.20)

A numerical approximation of the left-hand-side of this equation is

!!!/! !!!! !!! !!!/! !! !!!!!


!! !!!
∆! ∆!
∆!
− 𝑄! (4.8.21)

and for the right-hand-side, we have two choices

!! ! !!!!! !!!!
𝑆 !" ≈𝑆 ∆!
(forward FD) (4.8.22a)
!
!! ! !!! !!!!!!
𝑆 !" ≈𝑆 ∆!
(backward FD) (4.8.22b)
!

Explicit method of solution -- In the “explicit” method of solution, we use all the information at
the previous time step to compute the value at this time step. The method proceeds by
calculating for the unknown head point-by-point across the domain (i = 1, …, n). This method
can be unstable for large time steps. The finite-difference equation for the calculation is

l
hi+1 − 2hil + hi−1
l
S hil+1 − hil
= ; i = 1,...,n; l = 0,1,2,...
Δx 2 T Δt (4.8.23)  

T Δt l
2
(hi-1 − 2hil + hil+1 ) = hil +1 − hil
€ S Δx (4.8.24)

t − axis (index l)

i, l + 1

Δt
i − 1, l i, l i + 1, l
x − axis 
Δx (index  i)

i, l − 1

Figure 41. Grid points in backward in time (explicit) finite-difference approximation

78
 
Define the parameter of the calculation

T Δt
r=
S Δx 2 (4.8.25)  

so that

(
hil +1 = hil + r hi-1
l
)
− 2hil + hil+1
(4.8.26)
 
where the unknown head at time level l+1 is on the left-hand-side and all of the known
information form the l time level is on the right-hand-side.

Consider two slightly different values for the parameter r: r = 0.48, so that

S
Δt = rΔx 2 = 0.0128d ≈ 18.5min
T (4.8.27)  

and r = 0.52, so that


€ Δt = 0.0139d ≈ 20min (4.8.28)
 
When r = 0.48, we obtain the following results for the odd nodes in the mesh; however, when r =
0.52, we obtain the following results for the odd nodes in the mesh.

What’s going on here? At the initial time there is no flow across the aquifer, but at time t > 0,
flow occurs. The water released from storage in a grid cell over time Δt is

ΔV = SΔx(1)Δh (4.8.29)  

Water flowing out of the cell over the time interval


Δh
€ ΔV = T Δt
Δx /2 (4.8.30)
 
So, this needs to be limited to be less than the total water available in the cell, or

Δh
2T Δt ≤ SΔxΔh
Δx (4.8.31)  
or
T Δt 1
r= 2

€ S Δx 2 (4.8.32)  
 
When r > 0.5 the time interval is too large and the cell doesn’t contain enough water. This
causes an instable solution.

79
(a) r = 0.48

(b) r = 0.52
Figure 4.42. Results from solving the explicit approximation: (a) stable, and (b) unstable
solutions.

Implicit method of solution -- How can we avoid the instability of the explicit method? We can
use information from one point at the previous time step to compute the unknown values at all
points of the next time step. To do this we have to solve for all points in domain simultaneously,
and the method is inherently stable.

80
t, l
i − 1, l + 1 i, l + 1 i + 1, l + 1

Δt
i − 1, l i, l i + 1, l
x, i
Δx
i − 1, l − 1 i, l − 1 i + 1, l − 1

Figure 4.43. Grid points used in forward in time (implicit) finite-difference approximation.

The method proceeds by calculating for the unknown head point-by-point across the domain (i =
1, …, n). This method can be unstable for large time steps. The finite-difference equation for the
calculation is

l+1
hi+1 − 2hil+1 + hi−1
l+1
S hil+1 − hil
= ; i = 1,...,n; l = 0,1,2,...
Δx 2 T Δt (4.8.33)  

l +1
hi-1 −2hil +1 + hil+1
+1
S hil +1 −hil
=
€ Δx 2 T Δt (4.8.34)  

l +1
hi-1 −2hil +1 + hil+1
+1
S hil +1 −hil
=
€ Δx 2 T Δt (4.8.35)  

T Δt
r=
€ S Δx 2 (4.8.36)  

l+1
−rhi-1 + (1+ 2r)hil+1 − rhi+1
l+1
= hil (4.8.37)
€  
where the unknown head at time level l+1 is on the left-hand-side and all of the known
information form the l time level is on the right-hand-side. The set of equations to be solved at

each time step looks like the following for the first time step (h0 = hA = 6.1 m and h10 = hB = 1.5
m are the known boundary conditions on the left and right of the domain):

81
−rh10 + (1+ 2r)h11 − rh12 = h10  
−rh11 + (1+ 2r)h12 − rh13 = h20  
−rh12 + (1+ 2r)h13 − rh14 = h30  
€ −rh13 + (1+ 2r)h14 − rh15 = h40 (4.8.38)  
€ −rh14 + (1+ 2r)h15 − rh16 = h50  
€ −rh15 + (1+ 2r)h16 − rh17 = h60  
€ −rh16 + (1+ 2r)h17 − rh81 = h70  
€ −rh17 + (1+ 2r)h81 − rh19 = h80  
€ −rh81 + (1+ 2r)h19 − rh10
1
= h90  
€  
€ #1+ 2r −r & # h11 & # h10 + rh10 & #6.1+ r * 6.1&
% (% 1 ( % 0
( % (
€ % −r 1+ 2r −r ( %h2 ( % h2 ( % 6.1 (
% −r 1+ 2r −r %
( h3 1 ( % h30 ( % 6.1 (
% (% ( % ( % (
% −r 1+ 2r −r (%  ( %  ( %  (
% −r 1+ 2r −r (% ( = % (=% (
% (% ( % ( % (
−r 1+ 2r −r
% (% ( % ( % (
% −r 1+ 2r −r (% ( % ( % (
% ( % 1( % 0 ( % (
−r 1+ 2r −r % h8 ( % h8 ( % 6.1
% ( (
%$ −r 1+ 2r (' %$h19 (' %$ h90 + rh10
1 ( %$6.1+ r *1.5 ('
'
 
(4.8.39)  

4.8.5.4. Optimization of Steady Flow in a 1-D Homogeneous System
(adapted from Mays and Tung, Problem 8.3.1)
Consider the situation where the transmissivity of the aquifer is T = 1000 m2/day, the wells are
Δx = 80m apart, the heads on the left- and right-hand boundaries are constant values of 40m and
35m, respectively. The objective of the optimization model is to maximize the head distribution
in the aquifer while supplying a minimum pumping rate of Demand = 500 m3/d/m2. The
optimization model can be written as

82
4
Maximize ∑ hi
i =1
Subject to
4
∑ Qi ≥ Demand
i =1
Ai, j (hi+1,j − hi,j ) + Bi, j (hi,j − hi-1,j ) − Qi,j = 0 i = 1,...,4
hi = 40 i=0
hi = 35 i=5 (4.8.40)

The GAMS code for solving this model is listed in the Appendix.

The solution of the model is:

Node Head Pumping


0 40.00
1 11.00 3.96
2 7.33 0.00
3 3.67 0.00
4 0.00 6.04
5 35.00

4.8.6 Two-Dimensional Flow

4.8.6.1 Steady-State flow


Consider the steady, two-dimensional flow in a confined aquifer shown in Figure 4.5.1. The
governing equation is

∂ # ∂h & ∂ # ∂h & W ∂h
%Tx ( + %Ty ( ± ∑ Qw = S
∂x $ ∂x ' ∂y $ ∂y ' w=1 ∂t (4.8.41)

where T [L2/T] is the transmissivity of the aquifer and Q [L3/T/L2 = L/T] is the source/sink flow
€ rate per unit area of aquifer.

83
y, j Node No. (1,5) Unknown heads
(5,4) Known heads
(1,5) (2,5) (3,5) (4,5)

(0,4) (1,4) (2,4) (3,4) (4,4) (5,4)

(0,3) (1,3) (2,3) (3,3) (4,3) (5,3)


Δy
(0,2) (1,2) (2,2) (3,2) (4,2) (5,2)

(0,1) (1,1) (2,1) (3,1) (4,1) (5,1)

(1,0) (2,0) (3,0) (4,0)


x, i
Δx
Figure 4.44. Two-dimensional example finite-difference grid.

Considering the cell (i,j) (see Figure 4.5.2) we can see that the flows in and out of the cell are
represented by Darcy’s Law written for aquifer flow. When the transmissivity is homogeneous,
and we apply a finite-difference approximation to the second derivatives, we have

!! !! !! !!
! ! ! ! ! !
!" !!!/!,! !" !!!/!,! !" !,!!!/! !" !,!!!/!

∆!
+ ∆!
− 𝑄!,! = 0 (4.8.42)

!!!!,! !!!,! !!,! !!!!!,! !!,!!! !!!,! !!,! !!!,!!!


! !! ! !!
∆! ∆! ∆! ∆!
∆!
+ ∆!
− 𝑄!,! = 0 (4.8.43)

𝐴!,! ℎ!!!,! − ℎ!,! + 𝐵!,! ℎ!,! − ℎ!!!,! + 𝐶!,! ℎ!,!!! − ℎ!,! + 𝐷!,! ℎ!,! − ℎ!,!!! − 𝑄!,! = 0
(4.8.44)  

where
! !! ! !!
𝐴!,! = ∆! ! , 𝐵!,! = ∆! ! , 𝐶!,! = ∆! ! , 𝐷!,! = ∆! ! (4.8.45)  

This equation must be solved for each node (i,j) on the interior of the solution domain.

84
y
i-1/2 i+1/2 node (i,j)
Δx
j+1 cell (i,j)

Qy,j+1/2
j+1/2

Qx,i-1/2 Qx,i+1/2 Δy
Figurej 4.45. Finite difference grid for two-dimensional flow.

j-1/2
4.8.6.2 Boundary Conditions
Two types of boundary conditions are often encountered in groundwater flow problems:

Qy,j-1/2 h = f1(x,y,t) on C1.


Constant or prescribed head (Dirichlet) conditions:
j-1
In the previous example, if we write out the finite-difference equation for each node (i,j) where
the head is unknown (nodes (1,1) - (4,4)), wei have, e.g., at node (1,1)
i-1 i+1

𝐴!,! ℎ!,! − ℎ!,! + 𝐵!,! ℎ!,! − ℎ!,! + 𝐶!,! ℎ!,! − ℎ!,! + 𝐷!,! ℎ!,! − ℎ!,! − 𝑄!,! = 0
  (4.8.46)  
x
where the heads h0,1 and h1,0 are known from the boundary conditions. For node (2,3), we have

𝐴!,! ℎ!,! − ℎ!,! + 𝐵!,! ℎ!,! − ℎ!,! + 𝐶!,! ℎ!,! − ℎ!,! + 𝐷!,! ℎ!,! − ℎ!,! − 𝑄!,! = 0  
  (4.8.47)  

Prescribed flux (Neuman) conditions: qn = f2(x,y,t) on C2 or

∂h
= f 2 ( x, y, t ) on C 2
∂n (4.8.48)  

Consider the case where we have constant head boundary conditions on the west and east sides
of an aquifer and no-flow boundary conditions on the north and south sides.

85
y,j

(0,5) (1,5) (2,5) (3,5) (4,5) (5,5)


No flow
(1,5)

Constant (0,4) (1,4) (2,4) (3,4) (4,4) (5,4)


head Constant
head
(0,3) (1,3) (2,3) (3,3) (4,3) (5,3)
Flow
(0,2) (1,2) (2,2) (3,2) (4,2) (5,2)

(0,1) (1,1) (2,1) (3,1) (4,1) (5,1)


No flow

(0,0) (1,0) (2,0) (3,0) (4,0) (5,0)


x,i

Figure 4.46. Two-dimensional example with no-flow boundary conditions.

The north and south side no-flow conditions can be written as

∂h ∂h
=0 and =0
dy i, j =1 dy i, j = 4
(4.8.49)  

These conditions can easily be approximated by finite-differences, using a central difference


approximation as

hi,0 − hi,1 hi,5 − hi,4


= 0 and =0
Δy Δy (4.8.50)  
or
hi,0 = hi,1 and hi,5 = hi,4
(4.8.51)  

The resulting finite-difference equations for nodes along these no-flow boundaries are:

for j = 1 (south boundary):


Ai,1 (hi+1,1 − hi,1 ) + Bi,1 (hi,1 − hi-1,1 ) + Ci,1 (hi,2 − hi,1 ) + Di,1 (hi,1 − hi,0 ) − Qi,1 = 0
(4.8.52)  
 
for j = 4 (north boundary):

86
Ai,4 (hi+1,4 − hi,4 ) + Bi,4 (hi,4 − hi-1,4 ) + Ci,4 (hi,5 − hi,4 ) + Di,4 (hi,4 − hi,3 ) − Qi,4 = 0
(4.8.53)  
 
Substituting in the boundary information, we have

for j = 1 (south boundary):


Ai,1 (hi+1,1 − hi,1 ) + Bi,1 (hi,1 − hi-1,1 ) + Ci,1 (hi,2 − hi,1 ) + 0 − Qi,1 = 0
(4.8.54)  

for j = 4 (north boundary):


Ai,4 (hi+1,4 − hi,4 ) + Bi,4 (hi,4 − hi-1,4 ) + 0 + Di,4 (hi,4 − hi,3 ) − Qi,4 = 0
(4.8.55)  

4.8.6.3. Optimization of Steady Flow in 2-D Homogeneous System


Consider the flow in the example aquifer shown in Figure 4.5.2, where the mesh size is 10m,
western and eastern boundaries are held fixed at 100m and 50m respectively, the transmissivity
is 200 m2/day, and the maximum pumping is limited to 40 m3/day/m2. The optimization model
for this problem can be written as

9 8
Minimize ∑ ∑ Qi, j
i = 0 j =1
Subject to
9 8
∑ ∑ Qi, j ≥ Demand
i = 0 j =1
Ai, j (hi+1,j − hi,j ) + Bi, j (hi,j − hi-1,j ) +
Ci, j (hi,j+1 − hi,j ) + Di, j (hi,j − hi,j-1) − Qi,j = 0 i = 1,...,8; j = 1,...8
Ai,1(hi+1,1 − hi,1) + Bi,1(hi,1 − hi-1,1) + Ci,1(hi,2 − hi,1) = 0 i = 0; j = 1,...8
Ai,4 (hi+1,4 − hi,4 ) + Bi,4 (hi,4 − hi-1,4 ) + Di,4 (hi,4 − hi,3 ) = 0 i = 9; j = 1,...,8
hi, j = 100 i = 0,...,9; j = 0
hi, j = 50 i = 0....,9; j = 9
(4.8.56)

The GAMS code to solve this linear program is shown in the Appendix. The reader should note
that the grid orientation is rotated so that the origin is in the upper left-hand corner and the x axis
runs down the page. The results of solving this model are shown in the following table and
figure.

Objective value = 11566.11


Pumping
J0 J1 J2 J3 J4 J5 J6 J7 J8 J9 J10
I0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
I1 0.0 10.0 10.0 10.0 10.0 10.0 10.0 10.0 10.0 10.0 0.0
I2 0.0 10.0 10.0 0.0 0.0 0.0 0.0 0.0 10.0 10.0 0.0

87
I3 0.0 10.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 10.0 0.0
I4 0.0 10.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 10.0 0.0
I5 0.0 10.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 10.0 0.0
I6 0.0 10.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 10.0 0.0
I7 0.0 10.0 10.0 0.0 0.0 0.0 0.0 0.0 10.0 10.0 0.0
I8 0.0 10.0 10.0 10.0 0.0 0.0 0.0 10.0 10.0 10.0 0.0
I9 0.0 10.0 10.0 10.0 10.0 10.0 10.0 10.0 10.0 10.0 0.0
I10 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

Head
J0 J1 J2 J3 J4 J5 J6 J7 J8 J9 J10
I0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0
I1 100.0 96.0 94.6 94.5 94.5 94.5 94.5 94.5 94.6 96.0 100.0
I2 100.0 94.6 92.9 93.7 94.0 94.0 94.0 93.7 92.9 94.6 100.0
I3 100.0 94.4 93.7 93.6 93.7 93.7 93.7 93.6 93.7 94.4 100.0
I4 100.0 94.3 93.7 93.5 93.4 93.4 93.4 93.5 93.7 94.3 100.0
I5 100.0 94.1 93.4 93.1 93.0 93.0 93.0 93.1 93.4 94.1 100.0
I6 100.0 93.8 92.7 92.5 92.6 92.7 92.6 92.5 92.7 93.8 100.0
I7 100.0 93.4 90.9 91.7 92.3 92.5 92.3 91.7 90.9 93.4 100.0
I8 100.0 93.8 91.1 90.9 92.4 92.8 92.4 90.9 91.1 93.8 100.0
I9 100.0 95.6 93.8 93.3 93.6 93.7 93.6 93.3 93.8 95.6 100.0
I10 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0

4.8.7 Heterogeneous Systems

4.8.7.1. Introduction
Consider the steady, two-dimensional flow in a heterogeneous, anisotropic, confined aquifer.
The governing equation is (see Figure 4.5.3)

( ) (
∂ − Qx ∂ − Q y
+ =
)
∂x ∂y
∂ ⎛ x ∂h ⎞ ∂ ⎛ y ∂h ⎞
⎜ T ⎟ + ⎜ T ⎟ = Q (4.8.57)
∂x ⎝ ∂x ⎠ ∂y ⎜⎝ ∂y ⎟⎠

where Tx and Ty [L2/T] are transmissivities in the x and y directions, respectively. When we apply
a finite-difference approximation to the derivatives, we have

⎛ x ∂h ⎞ ∂h ⎞ ⎛ y ∂h ⎞ ⎛ ∂h ⎞
⎜ T ⎟
⎛
− ⎜ T x ⎟ ⎜⎜ T ⎟⎟ − ⎜⎜ T y ⎟⎟
⎝ ∂x ⎠ i +1/ 2.j ⎝ ∂x ⎠ i −1/ 2.j ⎝ ∂y ⎠ i.j +1/ 2 ⎝ ∂y ⎠ i.j −1/ 2
+ − Qi,j = 0
Δx Δy (4.8.58)

Applying a finite difference approximation to the remaining derivatives results in

88
⎛ x hi+1,j − hi,j ⎞ ⎛ x hi,j − hi-1,j ⎞
⎜ T ⎟ − ⎜ T ⎟
⎜ i+1/ 2 ,j Δx ⎟ ⎜ i-1/ 2 ,j Δ x ⎟
⎝ ⎠ ⎝ ⎠
Δx

⎛ y hi,j+1 − hi,j ⎞ ⎛ y hi,j − hi,j-1 ⎞


⎜ T ⎟ − ⎜ T ⎟
⎜ i,j+1/ 2 Δy ⎟ ⎜ i,j-1/ 2 Δy ⎟
+ ⎝ ⎠ ⎝ ⎠ − Q = 0
i,j
Δy (4.8.59)
where
x x y y
Ti+ 1,jTi,j y
Ti,j+1Ti,j
x Ti,j+
Ti+ 1/ 2 ,j =2
x x and 1/ 2 = 2 y y (4.8.60)
Ti+ 1,j+Ti,j
Ti,j+1+Ti,j

are the harmonic averages of the x- and y-direction transmissivities, respectively, between cells
(i,j) and (i+1,j) and between cells (i,j) and (i,j+1). We can write this equation as

Ai, j (hi+1,j − hi,j ) + Bi, j (hi,j − hi-1,j ) + Ci, j (hi,j+1 − hi,j ) + Di, j (hi,j − hi,j-1 ) − Qi,j = 0
(4.8.61)

where
2 Ti x+1,jTi,jx x x
− 2 Ti −1,jTi,j
Ai,j = , Bi,j = ,
Δx 2 Ti x+1,j + Ti,jx Δx 2 Ti x−1,j + Ti,jx
(4.8.62)
2 Ti,jy +1Ti,jy y y
− 2 Ti,j11Ti,j
Ci,j = , Di,j =
Δy 2 Ti,jy +1 + Ti,jy Δy 2 Ti,jy11 + Ti,jy

If no flow boundaries exist in the system, then appropriate modifications, as described in the
previous section, must be made to this system of equations.

89
Aside: Note that for an unequal grid spacing ( Δxi −1 ≠ Δxi ≠ Δxi +1 and Δyi −1 ≠ Δyi ≠ Δyi +1),
then we write

⎛ x hi+1,j − hi,j ⎞ ⎛ x hi,j − hi-1,j ⎞


⎜ T ⎟ − ⎜ T ⎟
⎜ i+1/ 2 ,j Δx ⎟ ⎜ i-1/ 2 ,j Δx ⎟
⎝ i +1 / 2 ⎠ ⎝ i −1 / 2 ⎠
Δxi

⎛ y hi,j+1 − hi,j ⎞ ⎛ y hi,j − hi,j-1 ⎞


⎜ T ⎟ − ⎜ T ⎟
⎜ i,j+1/ 2 Δy j +1 / 2 ⎟ ⎜ i,j-1/ 2 Δy j −1 / 2 ⎟
+ ⎝ ⎠ ⎝ ⎠ − Q = 0
i,j
Δy j

where Δxi −1 / 2 is the distance between the centers of cells (i-1, j) and (i,j), and

Ti,jx Ti+
x
1,j
x
Ti+ 1/ 2 ,j =2 Δxi +1 / 2
Ti,jx Δxi +1 + Ti+
x
1,j Δxi

4.8.7.2. Optimization of Steady Flow in a 2-D Heterogeneous System


[Add model description]

The GAMS code for this model is shown in the Appendix. The results of running this model are:
Objective value = 7014.44
Pumping
J0 J1 J2 J3 J4 J5 J6 J7 J8 J9 J10
I0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
I1 0.0 240.4 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
I2 0.0 318.3 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
I3 0.0 441.3 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
I4 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
I5 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
I6 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
I7 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
I8 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
I9 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
I10 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

Head
J0 J1 J2 J3 J4 J5 J6 J7 J8 J9 J10
I0 100.0 50.8 37.5 36.3 38.6 41.5 43.9 45.8 47.4 48.7 50.0
I1 100.0 12.4 25.7 32.8 38.0 41.8 44.4 46.2 47.7 48.9 50.0
I2 100.0 0.0 20.3 31.8 38.6 43.0 45.5 47.2 48.3 49.2 50.0
I3 100.0 0.0 25.4 36.3 42.2 45.8 47.6 48.6 49.3 49.7 50.0
I4 100.0 54.0 47.1 47.4 48.9 49.9 50.4 50.5 50.5 50.3 50.0
I5 100.0 74.7 63.3 58.5 56.1 54.5 53.5 52.6 51.8 50.9 50.0

90
I6 100.0 83.3 73.2 66.8 62.1 58.7 56.4 54.7 53.1 51.5 50.0
I7 100.0 87.6 78.6 72.1 66.5 62.0 58.9 56.5 54.3 52.1 50.0
I8 100.0 90.1 82.1 75.5 69.6 64.5 60.8 57.8 55.2 52.6 50.0
I9 100.0 91.3 84.1 77.6 71.5 66.1 62.1 58.8 55.8 52.9 50.0
I10 100.0 91.8 84.9 78.6 72.4 66.9 62.7 59.3 56.1 53.0 50.0

4.8.8 Transient Problems

4.8.8.1. Introduction
In transient problems, time is added as an independent variable. Therefore, we need to consider
the rate of release or uptake of water from or to storage in an aquifer. We also need to specify
initial head conditions in the aquifer in order to find a solution of a problem.

If we consider the transient, two-dimensional flow in a confined aquifer. The governing


equation is

∂ ⎛ x ∂h ⎞ ∂ ⎛ y ∂h ⎞ ∂h
⎜ T ⎟ + ⎜⎜ T ⎟⎟ − Q = S (4.8.63)
∂x ⎝ ∂x ⎠ ∂y ⎝ ∂y ⎠ ∂t

where S is the aquifer storativity. When the transmissivity and storativity are heterogeneous, and
we apply a finite-difference approximation to the second derivatives on the left-hand-side of this
equation, we have

LHS i, j = Ai, j (hi+1,j − hi,j ) + Bi, j (hi,j − hi-1,j ) + Ci, j (hi,j+1 − hi,j ) + Di, j (hi,j − hi,j-1 ) − Qi,j
(4.8.64)

Consider a backward finite difference approximation of the time difference on the right-hand-
side

t +1 t
t +1 hi,j − hi,j
RHS i,j =S i,j (4.8.65)
Δt

where hit, j and hit,+j1 are the heads in cell (i,j) at the time levels t, and t+Δt, respectively. Now,
we have not specified the time level where we are evaluating the LHS. Using a weighted average
of the LHS at the t and t+Δt levels, we have

S i,j
θLHS it,+j1 + (1 − θ ) LHS it, j =
Δt
( t +1
hi,j t
− hi,j ) (4.8.66)

where 0 ≤ θ ≤ 1 .

Various time-stepping schemes result from different selections of θ: e.g., Explicit Scheme:
θ = 0; and Crank-Nicolson: θ = 0.5.

91
One of the most common is the Fully Implicit Scheme: θ = 1.

Ai, j (hit++11, j − hit,+j1 ) + Bi, j (hit,+j1 − hit−+11, j )

+Ci, j (hit,+j1+1 − hit,+j1 ) + Di, j (hit,+j1 − hit,+j1−1 ) − Qit,+j1 (4.8.67)

= Fi, j (hit,+j1 − hit, j )

where
S i,j
Fi,j =
Δt (4.8.68)

4.8.8.2. Optimization of Transient Flow in 2-D Homogeneous System


[Add model description]

The GAMS code for this model is shown in the Appendix. The results of running this model are:
Objective = 9100.00
Pumping
Time = L0 Demand = 100.00
J0 J1 J2 J3 J4 J5 J6 J7 J8 J9 J10
I0 0 40 40 20 0 0 0 0 0 0 0
I1 0 0 0 0 0 0 0 0 0 0 0
I2 0 0 0 0 0 0 0 0 0 0 0
I3 0 0 0 0 0 0 0 0 0 0 0
I4 0 0 0 0 0 0 0 0 0 0 0
I5 0 0 0 0 0 0 0 0 0 0 0
I6 0 0 0 0 0 0 0 0 0 0 0
I7 0 0 0 0 0 0 0 0 0 0 0
I8 0 0 0 0 0 0 0 0 0 0 0
I9 0 0 0 0 0 0 0 0 0 0 0
I10 0 0 0 0 0 0 0 0 0 0 0
Time = L1 Demand = 200.00
J0 J1 J2 J3 J4 J5 J6 J7 J8 J9 J10
I0 0 0 0 0 40 40 0 0 0 0 0
I1 0 0 0 0 0 0 0 40 0 0 0
I2 0 0 0 0 0 0 0 0 0 0 0
I3 0 0 0 0 40 0 0 0 0 0 0
I4 0 0 0 0 0 0 0 0 0 0 0
I5 0 0 0 0 0 0 0 0 0 0 0
I6 0 0 0 0 0 0 0 0 0 0 0
I7 0 0 0 0 0 0 0 0 0 0 0
I8 0 0 0 0 0 0 0 0 0 0 0
I9 0 0 0 0 0 0 0 0 0 0 0
I10 0 0 0 40 0 0 0 0 0 0 0
Time = L2 Demand = 300.00
J0 J1 J2 J3 J4 J5 J6 J7 J8 J9 J10
I0 0 0 0 0 0 0 0 0 0 0 0
I1 0 0 0 0 0 0 0 0 0 0 0
I2 0 0 0 0 0 0 0 0 0 0 0
I3 0 0 0 0 0 0 0 0 0 0 0
I4 0 0 0 0 0 0 0 0 0 0 0

92
I5 0 0 0 0 0 20 0 0 0 0 0
I6 0 0 0 0 0 40 0 0 0 0 0
I7 0 0 0 40 0 0 0 40 0 0 0
I8 0 0 0 0 0 40 0 0 0 0 0
I9 0 0 0 0 0 0 0 0 0 0 0
I10 0 0 0 0 40 40 0 40 0 0 0
Time = L3 Demand = 400.00
J0 J1 J2 J3 J4 J5 J6 J7 J8 J9 J10
I0 0 0 0 40 0 40 0 0 0 0 0
I1 0 0 0 0 0 0 0 0 0 0 0
I2 0 0 0 0 0 40 0 0 0 0 0
I3 0 0 0 0 0 40 0 0 0 0 0
I4 0 0 0 0 40 0 0 40 0 0 0
I5 0 0 0 0 0 40 0 0 0 0 0
I6 0 0 0 0 0 40 0 0 0 0 0
I7 0 0 0 40 0 40 0 0 0 0 0
I8 0 0 0 0 0 0 0 0 0 0 0
I9 0 0 0 0 0 0 0 0 0 0 0
I10 0 0 0 0 0 0 0 0 0 0 0
Time = L12 Demand = 1.3E+3
J0 J1 J2 J3 J4 J5 J6 J7 J8 J9 J10
I0 0 0 0 40 15 0 0 28 0 0 0
I1 0 0 40 0 21 0 0 27 0 0 0
I2 0 0 40 40 15 0 0 25 0 0 0
I3 0 0 40 40 15 0 0 21 0 25 0
I4 0 0 40 0 23 0 0 17 0 40 0
I5 0 0 40 0 24 0 0 16 0 40 0
I6 0 0 40 0 24 0 0 15 0 40 0
I7 0 0 40 40 16 0 0 15 0 40 0
I8 0 0 40 0 23 0 0 14 0 40 0
I9 0 0 40 0 26 0 0 14 0 40 0
I10 0 0 40 0 27 0 0 14 0 40 0
Head
J0 J1 J2 J3 J4 J5 J6 J7 J8 J9 J10
Time = L0
J0 J1 J2 J3 J4 J5 J6 J7 J8 J9 J10
I0 100 100 100 100 100 100 100 100 100 100 50
I1 100 100 100 100 100 100 100 100 100 100 50
I2 100 100 100 100 100 100 100 100 100 100 50
I3 100 100 100 100 100 100 100 100 100 100 50
I4 100 100 100 100 100 100 100 100 100 100 50
I5 100 100 100 100 100 100 100 100 100 100 50
I6 100 100 100 100 100 100 100 100 100 100 50
I7 100 100 100 100 100 100 100 100 100 100 50
I8 100 100 100 100 100 100 100 100 100 100 50
I9 100 100 100 100 100 100 100 100 100 100 50
I10 100 100 100 100 100 100 100 100 100 100 50
Time = L1
J0 J1 J2 J3 J4 J5 J6 J7 J8 J9 J10
I0 100 88 76 61 40 34 38 39 43 47 50
I1 100 89 77 65 51 44 42 37 43 47 50
I2 100 90 79 68 56 50 47 45 46 48 50
I3 100 91 81 70 55 54 52 50 49 50 50
I4 100 91 82 73 64 59 55 53 52 51 50
I5 100 91 83 76 68 62 59 56 53 52 50
I6 100 91 84 77 70 65 61 57 55 52 50
I7 100 91 84 78 71 66 62 58 56 53 50
I8 100 91 84 77 71 66 62 59 56 53 50
I9 100 91 82 74 69 65 62 59 56 53 50
I10 100 90 80 67 67 65 62 59 56 53 50
Time = L2
J0 J1 J2 J3 J4 J5 J6 J7 J8 J9 J10
I0 100 93 86 79 72 66 62 58 55 52 50

93
I1 100 93 86 79 72 65 61 57 55 52 50
I2 100 93 85 77 70 64 59 56 53 52 50
I3 100 92 84 76 67 60 56 53 52 51 50
I4 100 90 81 72 63 54 51 50 49 49 50
I5 100 89 78 67 56 45 45 45 46 48 50
I6 100 86 74 61 50 37 38 39 42 46 50
I7 100 85 70 54 45 35 33 30 38 44 50
I8 100 84 70 55 41 28 30 31 36 43 50
I9 100 84 69 54 38 27 26 28 34 42 50
I10 100 83 68 51 27 18 21 20 31 41 50
Time = L3
J0 J1 J2 J3 J4 J5 J6 J7 J8 J9 J10
I0 100 83 65 44 36 24 30 34 39 44 50
I1 100 84 68 52 39 30 31 34 39 44 50
I2 100 85 70 55 40 25 30 33 38 44 50
I3 100 86 71 56 39 25 29 32 37 44 50
I4 100 85 71 55 34 28 29 27 36 44 50
I5 100 85 71 56 40 26 30 33 39 44 50
I6 100 85 71 56 43 28 33 37 41 45 50
I7 100 85 71 54 46 33 38 40 43 47 50
I8 100 86 74 62 53 45 44 44 45 48 50
I9 100 88 77 66 58 51 48 47 47 48 50
I10 100 88 78 68 60 53 50 48 48 49 50
Time = L12
J0 J1 J2 J3 J4 J5 J6 J7 J8 J9 J10
I0 100 68 36 9 0 0 0 0 15 32 50
I1 100 67 31 12 0 0 0 0 15 31 50
I2 100 66 30 8 0 0 0 0 14 30 50
I3 100 66 30 9 0 0 0 0 11 23 50
I4 100 65 30 14 0 0 0 0 9 19 50
I5 100 64 29 14 0 0 0 0 9 17 50
I6 100 62 28 13 0 0 0 0 8 17 50
I7 100 61 27 9 0 0 0 0 8 17 50
I8 100 62 28 13 0 0 0 0 8 17 50
I9 100 63 29 15 0 0 0 0 8 17 50
I10 100 63 30 16 0 0 0 0 8 17 50

4.9 Multiphase Systems


[NEED TO ADD THIS SECTION]

94

You might also like