You are on page 1of 26

Source: HYDRAULIC DESIGN HANDBOOK

CHAPTER 4
SUBSURFACE FLOW
AND TRANSPORT
Mariush W. Kemblowki and Gilberto E. Urroz
Utah Water Research Laboratory
Utah State University
Logan, Utah

4.1 INTRODUCTION
This chapter begins with the mathematical description of the constitutive relationships for
flow and transport in porous media. Following this, simple analytical solutions are pre-
sented for a variety of subsurface flow and transport problems. The principles of flow and
transport are outlined, and solutions are provided for practical problems of flow and trans-
port in both the saturated and the unsaturated zones. The latter includes problems of
transport in the vapor phase. The major focus is on the processes that are relevant to
subsurface mitigation.

4.2 CONSTITUTIVE RELATIONSHIPS


This section presents the basic concepts and laws used to describe flow and transport in the
subsurface. In particular, the constitutive relationships defining the fluid flow in fully and
partially saturated media are given as well as the relationships that describe diffusive and dis-
persive mass fluxes in porous media. Finally, we show the relations used to describe parti-
tioning of chemicals in the subsurface environment.

4.2.1 Darcy’s law

Consider the flow of a fluid through a pipe filled with a granular material, as shown in Fig.
4.1. In the figure, z1 and z2 represent the elevations of the pipe centerline above a reference
level at Sections 1 and 2, respectively, whereas p1/γ and p2/γ represent the water
pressure head at Sections 1 and 2, respectively. We define the piezometric head at any
location in the porous media as

h  z  p/γ (4.1)

where γ  specific weight (weight per unit volume) of water, typically, γ  9810 N/m3 or
62.4 lb/ft3. Let q be the average water velocity in the cross section of the pipe: i.e.,

q  Q/A (4.2)

4.1
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

4.2 Chapter Four

∆h
p1

γw

1 p2

γw
h1
2 h2
L
z1
z2

Arbitrary datum

FIGURE 4.1 Porous media flow.

where Q  volumetric discharge (volume per unit time) and A  total cross-sectional area
of the pipe (including the soil matrix). French hydrologist Henry Darcy discovered that
the average flow velocity could be estimated from
q  K (h1  h2)/L (4.3)
where L is the distance, measured along the pipe, between cross sections 1 and 2, and K
is a parameter that depends on the nature of the porous media as well as on the properties
of the transported fluid. For water, K is known as the hydraulic conductivity or the coef-
ficient of permeability. Typical values of K are given in many references (e.g., Bureau of
Reclamation, 1985) see tables 4.1 ard 4.2. Eq. (4.3) is known as Darcy's law and is com-
monly used to model the flow of fluids in porous media. Notice that the velocity V is not
the fluid velocity in the soil pores, it is an average velocity calculated over the entire area
of the flow cross section. The average pore velocity is calculated as
q
v   (4.4)
θW
where θW is the volumetric water moisture content. Note that for saturated flow, θW  n,
where n is porosity.
Darcy’s law (i.e., Eq. 4.3) also can be written more concisely as
q  KI (4.5)
where I is the hydraulic gradient defined as
(h1  h2)
I  4.6)
L
Hydraulic conductivity K is a function of aquifer and fluid properties –specifically of
the intrinsic soil permeability k, fluid viscosity µ, and fluid density ρ–and is given by
K  k ρg/µ (4.7)
For saturated flow of a constant density fluid in isotopic porous media, the Darcy law
can be written as

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

Subsurface Flow and Transport 4.3

 ∂h 
qi  K 
∂x  (4.8)
 j 
For anisotropic media, the Darcy law is written as
 
qi   Kij ∂h (4.9)
 ∂xj 

where Kij is the conductivity tensor. For saturated flow of a fluid of variable density in
anisotropic porous media, we have
k il,ij  ∂p ∂z 
qi   so   ρfluid g (4.10)
µfluid  ∂xi ∂xi 
where ksoil,ij is the intrinsic permeability tenser. Finally, for unsaturated flow of variable-
density fluid in anisotropic porous media, we have
kr(θ)ksoil,ij  ∂p ∂z 
qi       ρfluid g (4.11)
µfluid  ∂xj ∂xj 
where kr(θ)  relative permeability of the porous media. Relative permeability is a func-
tion of soil saturation, which in turn is a function of the capillary pressure. These rela-
tionships for partially saturated flow are discussed in the next section.

4.2.2 Unsaturated Flow–Constitutive Relationship

In unsaturated flow, the concern is water movement in the zone above the water table. In
this case, the water saturation Sw is a function of the difference between air and water
pressures because the water is resulting from held by capillary forces resulting from sur-
face tension. This difference is known as the capillary pressure and is defined as
haw  ha  hw (4.12)
Typically, in unsaturated flow theory we assume negligible resistance to the gas-phase
flow in porous media; as a result, we also can assume that the gas-phase pressure is uni-
form and equal to the atmospheric pressure. Hence haw   hw. The aqueous pressure in
the unsaturated zone is lower than the atmospheric pressure; thus, the capillary pressure
is positive. The negative pressure head hw also is known as the soil matrix suction Ψ. Thus,
the total head in the aqueous phase is h  Ψ + z.
To remove water from the pore space i.e., to reduce the water content we have to apply
more negative pressure to the aqueous phase, i.e., increase the capillary pressure haw. This
relationship is typically called the soil-water retention curve, and can be expressed by the
following commonly used parametric models: the Brooks-Corey (BC) model and the van
Genuchten (VG) model. The BC model is

θθ  λ
θe  r  Ψ (4.13)
n  θr  hb 
for Ψ  hb and is otherwise (capillary fringe zone),

θe  1 (4.14)

where n  porosity, θ  volumetric moisture content (equal to n Sw), θe  effective volu-


metric moisture content, θr  residual (irreducible) moisture content, λ  Brooks-Corey
parameter, and hb  capillary fringe height. The van Genuchten model is

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

4.4 Chapter Four


θ
θe  θ  r  θr  [1  (αΨ)n]m (4.15)
n
where: n  curve fitting parameter that depends on the type of soil, α  soil property
index, α ≈ 1/hb, and m  1  1/n.
Note that parameter n depends on pore-size distribution. For a well-graded soil (wide
pore-size distribution, which results in a flatter moisture content curve θ(Ψ)), n is small,
whereas for poorly graded soils (narrow pore-size distribution, which results in a steeper
moisture content curve θ(Ψ), n is large: typically values of n higher than 2.5.
Also note that although the BC and VG models are the most commonly used models
in analysis, there is no restriction on using different mathematical representations to
describe the characteristics of soil-water retention. For example, a simple exponential
model, such as θe  exp( β Ψ), is in some cases, sufficient to describe the physics of the
retention.
As the moisture content in partially saturated media decreases, so does the volume of
pores available to fluid flow. Thus, hydraulic conductivity for the partially saturated media
depends on the water content and, in turn, on the metric suction. To describe this rela-
tionship, we modify the value of intrinsic permeability k by the factor of kr(θ) or kr(Ψ),
called relative permeability. Several models for kr are shown below:
BC model of relative permeability:
2  3λ
kr  θe λ (4.16)

VG model of relative permeability:

kr  θe0.5(1  (1  θe1/m)m)2 (4.17)

Mualem model of relative permeability:

kr  exp [ α Ψ] (4.18)

4.2.3 Diffussive and Dispersive fluxes

4.2.3.1 Molecular diffusion. Molecular diffusion describes the process by which a con-
taminant species dissolved in an environmental fluid moves from regions of higher
concentration to regions of lower concentration. When the only mechanism affecting the
diffusion of the contaminant species is the random motion of its molecules, the process is
referred to as molecular diffusion. The mass flux of a solute along a single direction, in a
liquid or gaseous body, is described by Fick's law:

冢 冣
dC
q  D  (4.19)
dx
In this equation, q  mass flux of solute per unit area per unit time [ML2 T1],
D  diffusion coefficient (L2 T 1), C  solute concentration  mass of solute/volume of
solution (M/L3), dC/dx  concentration gradient along the x direction. The minus sign in
Eq. (4.19) indicates that the solute flux will go from regions of larger concentration to
those of lower concentration. Values of the diffusion coefficient, D, depend on the type of
solute and the type of environmental fluid. For major cations and amnions dissolved in
water, values of D range from 1  109 to 2  109 m2/s (Fetter, 1994).

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

Subsurface Flow and Transport 4.5

4.2.3.2 Molecular diffusion in porous media Molecular diffusion in a porous medium


is affected by the nature of the medium. For example, the paths that diffusing molecules
follow in a porous medium are, in general, more complicated than if they were diffusing
in water. Although Eq. (4.19) can still be used to describe the diffusion in a porous medi-
um, the diffusion coefficient must be modified, and Eq. (4.19) is rewritten in terms of an
effective diffusion coefficient. One widely accepted expression for the effective diffusion
coefficient in porous media is the Millington-Quirk equation
θ3.33
D  Do  
solution
(4.20)
n2
where Do is the molecular diffusion coefficient of the compound in pure solution fluid,
θsolution is the solution fluid-filled porosity of the soil, and n is the total porosity of the soil.

4.2.3.3 Mechanical dispersion and macro-dispersion. Mechanical dispersion refers to


the component of dispersion caused by differences in velocity at the pore level that are a
consequence of the pore geometry. Water will move at different rates as a result of differ-
ences in pore sizes and tortuosity. A contaminant dissolved in the water flowing through
a porous medium will be dispersed in both the longitudinal and transverse directions
because of the fluctuations in the water velocity field. A way to incorporate the influence
of pore geometry in the dispersion process is to define longitudinal and transverse disper-
sivities αL and αT. Longitudinal and transverse mechanical dispersion coefficients can thus
be defined in terms of the dispersivities and the average pore velocity. For example, the
longitudinal mechanical dispersion coefficient (DL)mech will be given by (DL)mech  αL v.
This dispersion coefficient is not treated separately from the effective diffusion coefficient
defined in (8); instead, they are both combined in a coefficient of hydrodynamic disper-
sion. Thus, Fick’s Law, which describes molecular diffusion in a fluid, can be used to
describe longitudinal and transverse dispersion in a porous medium if the diffusion coef-
ficient D in Eq. (4.19) is replaced by a coefficient of longitudinal (or transverse) hydro-
dynamic dispersion, DL or DT, given by

DL  αL v  D* (4.21)

or

DT  αT v  D* (4.22)

where αL, αT  longitudinal and transverse dispersivities, respectively, and D*  effec-


tive porous-media diffusion coefficient.

In addition to the pore-scale dispersion, we also have formation-scale dispersion or,


more accurately, spreading, which is a result of the variability in transport velocity caused
by the heterogeneity of the hydraulic conductivity field. In terms of magnitude, this
microdispersive flux is significantly larger than the one related to mechanical dispersion.
Mathematically, macrodispersive flux is the flux equal to the expected value of the prod-
uct of Darcian velocity (q’) and of the contaminant concentration (C’) fluctuations:

q macrodispersive   q’ C’  (4.23)

In otherwords, this flux can be described using an expression similar to the Fickian diffu-
sion equation (Eq. 4.19) with a macrodispersivity coefficient. In the most general (three-
dimensional) case, the equation defining the macrodispersive flux in the flowing fluid is

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

4.6 Chapter Four

∂苶苶
C
qi  ( Aij苶
v ) (4.24)
∂xj
where Aij represents the macrodispersivity tensor and the bar indicates averaged quanti-
ties. The fluctuations q’ and c’ result from the heterogeneous nature of the aquifer. This is
expressed in the way the macrodispersivity tensor is estimated. For example, the longitu-
dinal macrodispersivity is estimated using
σ 2 λ1
A11  f  (4.25)
γ2
where σf2  variance of log-conductivity (f  Ln[K]), λ1  correlation scale in the direc-
tion of flow, and γ is given by

q  σf2 
γ    exp   (4.26)
KgJ1  6 

In summary, the longitudinal and transverse components of the total diffusive and dis-
persive mass flux (per unit bulk area) in heterogeneous geologic formations is estimated
as follows:
∂C
qL  (θFLUID AL v FLUID  D)   (4.27)
∂x

∂C
qT,HOR  (θFLUIDAT,HORvFLUID  D)   (4.28)
∂y

∂C
qT,vert  (θFLUIDAT,vertvFLUID  D)   (4.29)
∂y

where D  Millington-Quirk effective dispersion coefficient.

4.2.4 Partitioning

Equilibrium partitioning and sorption are the most common chemical processes that
affect reactive transport. These processes are dealt with by equating the total concentra-
tion to the sum of the concentrations in each phase multiplied by their respective volumes.
Furthermore, by equating the concentration in each phase to the concentration in a com-
mon phase —say, the concentration in water— the total concentration can be expressed in
terms of the common phases concentration and a retardation coefficient R

CT  θWATERCWATER(1  冘
I  WATER
K θ1
1
θWATER
)  θWATERCWATERR (4.30)

where: KI  partitioning coefficient between the ith phase and the common phase KI  CI /
CWATER, θI  the volumetric content of the ith phase, and θWATER  the volumetric content
of the common phase. This type of equilibrium partitioning is frequently used to describe
the relationships between concentrations in the following scenarios: (1) vapor and aque-
ous phases, (2) soil and vapor phases, (3) soil and aqueous phases, (4) partitioning of a
tracer between aqueous and NAPL or DNAPL phases, and (5) partitioning of a tracer
between vapor and NAPL or DNAPL phases.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

Subsurface Flow and Transport 4.7

For partitioning of compounds present in a NAPL or DNAPL mixture and aqueous


phase, we use

CI,WATER  xISI (4.31)

where SI is the solubility of compound I in water and xI is its mole fraction in the mixture.
Finally, for partitioning of compounds present in a NAPL or DNAPL mixture and vapor
phase, use
Pvi[atm]* Mw,i[g/mole]
Vi[mg/L]  103[mg/g]xi  (4.32)
R  0.0821[L  atm/mole  °K]T[°K]

where VI is the vapor concentration of compound I, Pv,I is the vapor pressure of compound
I, MW,I is its molecular weight, R  gas constant, and T  temperature in °K.

4.2.5 Degradation

In addition to partitioning, degradation of compounds also may affect the fate and
transport of reactive compounds significantly. Typically, degradation is modeled using
either power-order decay models or growth-process-based models. In environmental
subsurface hydrology, three basic power-order models are used: (1) zero-order decay,
(2) first-order decay, and (3) a combination of the first two models. According to these
models, the total decrease of mass in unit bulk volume caused by degradation is
expressed by

∂CT
  
∂t 冘
I  PHASES
θIK0,I  冘
I  PHASES
θIK1,ICI (4.33)

where K0,I  zero-order degradation rate of the compound in phase I, K1,I  first-order
degradation rate of the compound in phase I, and CI = mass/volume concentration of the
compound in phase I.
In addition to the zero-order and first-order degradation processes, the Monad kinetics
is frequently used to describe oxygen limited aerobic degradation of organic compounds
in the aqueous phase. According to the Monod model, the degradation rate in terms of the
total concentration is expressed by the following system of equations:

∂CT  CWATER   OWATER 


  θWATERMtτ  (4.34)
∂t  KC  CWATER   KO  OWATER 

and

∂OT  CWATER   OWATER 


  θWATERχMtτ  (4.35)
∂t  KC  CWATER   KO  OWATER 

where CWATER is the aqueous concentration of the contaminant, OWATER is the dissolved oxy-
gen concentration, Mt is the total concentration of the active microbial biomass, τ is the
maximum rate of organic solute utilization, KC is the concentration of the organic solute
at which the utilization rate is half the maximum, KO is the electron acceptor (oxygen)
concentration at which the utilization rate is half the maximum, and χ is the substrate uti-
lization ratio.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

4.8 Chapter Four

4.3 FLOW AND TRANSPORT IN SATURATED ZONES


In this section, we present some solutions to saturated flow and transport problems that
are encountered in the practice of subsurface hydrology. We begin with well hydraulics,
an understanding of which is important to the design of pump-and-treat systems, and dis-
cuss several models of transport of soluble plumes.

4.3.1 Flow to a Single Well

Darcy's law describes the average water flux through a porous medium when the local
hydraulic gradient is known. To determine discharges we use the law of conservation of
mass for the water, also known as the continuity equation. For the analysis of wells, it is
assumed that the flow toward the well caused by the well pumping is radially symmetric.
In this situation, it is convenient to use the equation of continuity in radial coordinates
(r, θ). For steady-state state flow toward a single well in a confined aquifer without
recharge, the equation reads

冢 冣
dh
Q  2πKB  (4.36)
dr
where K and B  aquifer conductivity and thickness, respectively. The equation of conti-
nuity for well flow in an unconfined aquifer is similar to Eq. (4.19): namely,

冢 冣
dh
Q  2πrKh  (4.37)
dr
with B replaced by the variable flow depth h. Implicit in Eq. (4.36) and (4.37) is the
Dupuit-Forchheimer assumption, the implication of which is that the flow in the aquifer
can be assumed to be practically horizontal.
Equations (4.36) and (4.37) are used to obtain solutions for the steady-state discharge
to a well in confined and unconfined aquifers, respectively, if the piezometric heads (con-
fined aquifers) or water-table elevations (unconfined aquifers) h1 and h2 are known at two
radial distances r1 and r2, respectively:

Q  2πKB(h2  h1)/ln(r2/r1)  2πT(h2  h1)/ln(r2/r1) (4.38)

where T  KB  aquifer transmissivity, and

Q  πKD(h22  h12)/ln(r2/r1) (4.39)

The solutions given in Eqs. (4.38) and (4.39) assume that the well penetrates to the
impermeable bottom of the aquifer and that there is no recharge into the aquifer. They also
assume an infinitely large aquifer with no interaction with surface streams or imperme-
able boundaries. Well solutions are often given in terms of the drawdown s as a function
of the radial distance r. The drawdown is defined as

sHh (4.40)

where H is the elevation of the original piezometric surface before pumping at the well
starts. The distance R for which h  H and s  0 is called the radius of influence of the
well. Using the concepts of drawdown and radius of influence, the steady-state flow equa-
tion in a confined aquifer can be rewritten as
Q  
s   LnR (4.41)
2πT  r 

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

Subsurface Flow and Transport 4.9

By combining the solutions for the steady state flow in confined and unconfined
aquifers, one can derive a relationship between the drawdown calculated from confined
conditions (assuming constant in space and time aquifer thickness equal to H), and that
estimated for unconfined conditions (when the change in aquifer thickness caused by
pumping is taken into account):

sUNC  H  兹H
苶2苶
苶苶s
2苶CON
苶苶FH (4.42)

Using this formula, known as Jacob’s correction, one can initially assume constant
aquifer thickness B  H in calculations and use “confined” aquifer solutions to calculate
drawdown, then correct the drawdown using Jacob’s correction. This approach is particu-
larly useful when dealing with transient flow. For transient flow conditions in an aquifer
with constant flow thickness, the transient drawdown is given by

Q ∞ exp[x] Q
s (r,t)   兰  dx   W [u(r,t)] (4.43)
4πT u x 4πT
where
r2S
u(r,t)    (4.44)
4Tt
and S  storativity (for confined aquifers) or porosity (for unconfined aquifers) and W(u)
is known in subsurface hydrology as well function and in mathematics as exponential inte-
gral. This function is tabulated in almost every groundwater hydrology textbook. It also is
available in many engineering mathematics software packages, such as Mathematica©, as
a library function.

4.3.2 Superposition and Convolution

For a time-variable pumping rate, the principle of convolution can be used to estimate the
transient drawdown. This approach is strictly valid for linear systems: i.e., systems in
which the response (drawdown) is a linear function of the excitement (pumping rate). The
linearity assumption is strictly valid for confined aquifers only; however, as long as the
drawdowns do not exceed 20% of the initial aquifer thickness, it also may be used for
unconfined aquifers. Using the convolution approach, the transient drawdown for a pump-
ing rate changing in a step-wise fashion is given by


n
s(r,t)  1 (QK  QK  1)W(u(r,∆tK  1)) (4.45)
4πTk  1

where the drawdown is estimated at time t, tn  t  tn  1, QK is the pumping rate for


tK  1  t  t K,tO  0, Q0  0.0, and ∆tK  1  t – tK  1. When several wells are present,
the superposition approach is used to estimate the cumulative drawdown by adding the
drawdown contributions from all the wells:


m
s(t)  1 QLW(u(rL,t)) (4.46)
4πT L  1

where rL is the distance between the point of interest (where the drawdown is estimated) and
well L. When several wells are pumping at variable rates, the superposition and convolution
approaches are used simultaneously. The superposition principle also can be used to super-
impose the drawdown on the natural (ambient) flow conditions. Using this principle leads to

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

4.10 Chapter Four

h(x, y, t)  H(x, y, t)  s(x, y, t) (4.47)

where h  transient potentiometric surface that combines ambient conditions and well
impact, H  potentiometric surface under natural (ambient) conditions, and s  transient
drawdown.

4.3.3 Interception Wells

With respect to contaminant transport in the subsurface, interception wells are used to trap
the contaminant plume within the well flow field. It is assumed in this analysis that there
is an ambient steady-state uniform flow through the aquifer. The combination of well-
related flow and ambient uniform flow satisfies the conditions of two-dimensional poten-
tial flow in a horizontal plane, where the discharge described by stream function y is relat-
ed to potential φ, or the piezometric head h. The extent of the aquifer through which water
travels to the well and is captured by it is called the capture zone. The derivation of the
analytical solution for steady-state flow capture-zone uses the following assumptions:
(1) ahomogeneous, isotropic, infinitely large aquifer, (2) uniform flow, (3) no leakage, (4)
aquifer storativity or specific yield neglected, (not relevant for steady-state analysis),
(5) hydrodynamic dispersion neglected, (6) the Dupuit assumption applies, and (7) the
well is fully penetrating and pumping at a constant rate. Three important parameters are
used in delineating the capture zone: namely, the stagnation point, the upgradient maxi-
mum width of the capture zone, and the equation for the capture zone boundary.
For a confined aquifer, the distance from the well to the stagnation point (measured in
the direction of the uniform flow) is
Qw
xSTAG   (4.48)
2πTI
where Qw  well discharge, T  aquifer transmissivity  KB (K  aquifer permeability,
B  aquifer depth), and I  natural hydraulic gradient: i.e., the gradient responsible for
the ambient steady-state uniform flow in the aquifer. The upgradient divide, defined by
the maximum width of the capture zone far upgradient of the well, for the confined aquifer
is given by
Q
wDIV  w (4.49)
TI
and the equation of the dividing streamline is
y
x   (4.50)

冢 冣
2πTIy
tan π  
Qw
The procedure for delineating the capture zone consists of the following steps: (1) esti-
mate the location of the stagnation point (xSTAG, 0), (2) estimate the maximum width of the
capture zone wDIV, and (3) vary y between zero and wDIV/2 and use the capture zone bound-
ary to estimate the boundary location (x,y).

4.3.4 Partially Penetrating Wells

Performance of wells that penetrate only partially through the bearing strata is discussed
in this section. The simplest case consists of a well that is barely penetrating into an semi-

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

Subsurface Flow and Transport 4.11

infinite porous medium so that the aquifer flow is three-dimensional and spherically sym-
metric. In this case, the following relationship applies between the flow into the partially
penetrating well Qp and the flow to a fully penetrating one Q:
Qp rw R
   ln  (4.51)
Q B rw
where B is the aquifer thickness, rw is the well radius, and R is the radius of influence of
the partially penetrating well. Because, in general, rw  B, then the equation above indi-
cates that the spherical flow to a partially penetrating well is highly inefficient compared
with simple radial flow: i.e., for the same drawdown in the well, it results in a significantly
smaller pumping rate.
In the general case of partial penetration, one may consider the total drawdown sT,
which consists of the drawdown equivalent to that of a fully penetrating well s and addi-
tional head loss because of the partial penetration of the well ∆s:

sT  s  ∆s (4.51)

Additional head loss for a well penetrating from the top (or the bottom) of the aquifer
is estimated as follows:
Q(1  p)  (1  p)h 
∆s   ln s  (4.53)
2πTp  rw 

where p  penetration factor; p  hs/B; hs  penetration depth; and B  aquifer


thickness (Fig. 4.2). For the well centrally positioned in the aquifer, the following formu-
la is used:
Q(1  p)  (1  p)h 
∆s   ln s  (4.54)
2πTp  2rw 

Thus, when the pumping rate is defined for a well, we calculate the drawdown correc-
tion ∆s and add it to the full penetration drawdown s. However, when the drawdown is
given for a well, we have to recalculate the pumping rate. In this case, the true pumping

hs B/2

B hs
hs/2

2rw

FIGURE 4.2 Partially-penetrating well.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

4.12 Chapter Four

rate is given by
S  ∆s
Qs  Qs  Q T
 (4.55)
s  ∆s S
T

where s  drawdown defined at the well, Q  pumping rate estimated using s, and
Qp  actual pumping rate

4.3.5 Well Duplets

Well duplets, each of which consists of one pumping and one recharge well, are frequent-
ly used as a means of injecting and removing aquifer mitigation solutes, such as co-sol-
vents, surfactants, or both. Typically, the recharge well is positioned directly upgradient
from the discharge well, and the magnitudes of pumping and injection rates are the same.
In this case, the two wells form a flow circulation cell: i.e., all the injected water is
pumped out by the discharge well. In the case of co-solvent flushing, it is important to
understand what region of the aquifer is subject to the mitigation: i.e., what the boundary
is of the circulation cell. This boundary is defined by the upgradient and downgradient
stagnation points (xSTAG, 0) and (xSTAG, 0) and the cell boundary equation. For the x-axis
parallel to the direction of ambient flow and the origin of the coordinate system located at
the mid-point between the two wells, the two stagnation points, (xSTAG, 0) and (xSTAG, 0)
are given as the roots of the quadratic equation

qoBd 1  x/d x/d 


   2  2  0 (4.56)
Qw 2π  (x/d  1) (x/d  1) 

where qo  ambient flux, 2d  distance between the wells, and Qw  pumping/injection


rate. The boundary of the circulation cell is defined by
 

qoBy 1 y/d   y/d  1
  tan1  tan1    (4.57)
Qw 2π   (x/d  1)   (x/d  1)  2

The circulation cell is symmetric with respect to the y axis. The cell delineation pro-
cedure consists of estimating the locations of the stagnation points and varying x between
zero and xSTAG and using the cell-boundary equation to solve for y. This implicit equation
can be solved by any calculation software, such as Mathematica© or MS Excel.

4.3.6 Transport Equations

The following general form of mass transport equation in the saturated zone is derived assum-
ing one-dimensional advective and three-dimensional diffusive-dispersive transport in the
aqueous phase, linear partitioning of a compound between the three phases (water-soil-
(D)NAPL), and first-order degradation in the aqueous phase. For this conditions, we have

冢冢α v  θ 冣∂x冣 v∂x λC


∂CW ∂ D ∂CW ∂CW
R   (4.58)
∂xi i
W i 1
W

where CW  aqueous phase concentration, v  pore-water velocity, and θW  volumetric


moisture content, and αi  longitudinal, transverse horizontal, and transverse vertical

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

Subsurface Flow and Transport 4.13

macrodispersivities, and D  Millington-Quirk dispersion coefficient, and λ  first-order


degradation rate in the aqueous phase, and R  retardation factor,

kSOILγβ kNAPLθNAPL
R1    (4.59)
θW θW

where kSOIL  water-soil partitioning coefficient, kSOIL  S/CW, S  weight/weight con-


centration of absorbed compound in soil, kNAPL  NAPL – water partitioning coefficient,
kNAPL  CNAPL/CW, CNAPL  concentration of compound in the NAPL phase, γB  bulk den-
sity of dry soil, and θNAPL  volumetric NAPL content.
Assuming that the diffusive fluxes are negligible compared the macrodispersive flux-
es, the transport equation can be simplified to yield

∂CW ∂  ∂CW  ∂CW


   αivR   vR   λRCW (4.60)
∂t ∂xi  ∂xi  ∂xi

where vR  v/R and λR  λ/R.

4.3.7 Selected Analytical Solutions

Closed-form solutions are available for a variety of flow, boundary, and initial conditions.
Van Genuchten and Alves (1982) presented a good summary of these solutions. Some of
the most useful solutions are presented below.

4.3.7.1 One-dimensional transport with step change in concentration–no degrada-


tion. This simple case has the initial condition C(x,0)  0 for x 0, and it is subject to
the following boundary conditions: C(0,t)  Co, t 0 and C(∞,t)  0, t 0. The solution
of the transport equation for these conditions is given by

冢 冣
CO  x  vRt   x   x  vRt 
C(x,t)   Erfc    expErfc  1/2  (4.61)
 2(αxvRt)   αx   2(αxvRt) 
1/2
2

4.3.7.2 One-dimensional transport with step change in concentration and first-order


degradation. The initial and boundary conditions are the same as in Sec. 4.7.1.
The solution is given by

CO
 x   4λRαx 1/2
C(x,t)   exp1  1    
2  2αx   vR  

冢 冣
4λ α 1/2

冤 冥
x  vR 1  Rx t
vR
Erfc  1/2 (4.62)
2(αxvRt)

where Erfc  complementary error function.

4.3.7.3 Continuous point injection, 2-D dispersive transport, no retardation, and no


degradation. A tracer is continuously injected at a rate Q (per unit depth of the aquifer)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

4.14 Chapter Four

with a concentration Co into a uniform flow field from a point (x  0, y  0). Let the
uniform velocity be vx. The asymtotic solution, i.e., for t →∞, is given by

 
冪莦
 莦莦
4D D莦莦莦 D莦莦
2
 Co   vx  v x  2
y 2
C(x,y)    exp x Ko    
x
  (4.63)
苶苶
 2π兹D LDr   2DLDT   L  L T

where Ko  modified Bessel function of the second kind and of 0th order (Bear, 1972).
The time-dependent solution is

 CoQ  vx
C(x,y)    expx Ko[W(0,β)]  W(t,β) (4.64)
苶苶
 2π兹D D苶
L T   2DL 

where

冪莦4v莦D莦莦冢莦Dx莦莦莦莦
 莦
D 冣
y
2 2 2
β x
(4.65)
L L T

W (t, b)  leaky well function (see, for example, Hunt, 1983, p. 100).

4.3.7.4 Point slug injection into a uniform flow field—3-D transport and retardation.
In this case, a slug of contaminant of the mass M  CoV is injected at point (0,0,0). The
transient distribution of concentration is described by
VoCo
C(x, y, z, t)   
8(πvRt)3/2(αxαyαz)1/2

 (x  vRt)2 y2 z2 
exp       (4.66)
 4αxvRt 4αyvRt 4αzvRt 

4.3.7.5 Continuous injection from a finite-sized source with retardation and degrada-
tion. In this case, consider transport from a rectangular source that is perpendicular
to the direction of flow. The source width is Y, and its depth below the water table is Z.
The transient concentration distribution in the presence of retardation and degradation is
given by

C  x   1  4λRαx 1/2
C(x, y, z, t)  o exp1    
8  2αx   vR  
αx 1/2
x  vRt(1  4λRv)
 
Erfc  2(αxvRt)1/2 
 

  y  Y/2   y  Y/2 
Erfc   Erfc  
 2(αyx)1/2   2(αyx) 
1/2

  zZ   z  Z 
Erfc  1/2   Erfc   1/2  (4.67)
  2(αzx)   2(αzx) 

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

Subsurface Flow and Transport 4.15

4.4 FLOW AND TRANSPORT IN UNSATURATED


ZONE — AQUEOUS PHASE
In this section, we briefly discuss the flow continuity equations and present some simple
solutions to selected flow problems. That discussion is followed by a presentation of mass
transport in the water phase of unsaturated zone.

4.4.1 Flow in an Unsaturated Zone

The continuity equation for an unsaturated flow system can be written as


∂θ
  ∇.q (4.68)
∂t
or
∂θ  ∂qx ∂qy ∂qz 
       (4.69)
∂t  ∂x ∂y ∂z 

Combining Darcy's law with the mass continuity equation, we can write the final flow
equations is
∂θ
  ∇(K(Ψ)∇h) (4.70)
∂t
The flow equations can be simplified for horizontal and vertical flow conditions.
a. One-dimensional horizontal flow:
∂θ ∂  ∂Ψ 
   K(Ψ) (4.71)
∂t ∂x  ∂x 
where θ  volumetric water content. In this, the contribution of the elevation head, z,
vanishes, since ∂z/∂x  0.
b. One-dimensional vertical flow:

∂θ ∂   ∂Ψ 
   K(Ψ)  1 (4.72)
∂t ∂z   ∂z 

Note that the flow equations are characterized by the presence of two dependent, albeit
related, variables: namely, θ and Ψ. To simplify this situation, we describe the relation-
ship between θ and Ψ by a term called soil diffusivity D(θ) as
K(θ)
D(θ)   (4.73)
C(θ)
where C(θ) is called specific moisture capacity and is defined as

C(θ)   ∂θ
 (4.74)
∂Ψ
Using these definitions, the flow equation can be written as follows:
c. One-dimensional horizontal flow:

∂θ ∂ ∂θ 
  Dx(θ) (4.75)
∂t ∂x  ∂x 

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

4.16 Chapter Four

d. One-dimensional 1-D vertical flow:


∂θ ∂ ∂θ
   (Dz(θ)   Kz(θ)) (4.76)
∂z ∂z ∂z
As can be seen, we now have only one dependent variable: namely, θ. The only limitation of
this formulation is that specific moisture capacity C(θ) becomes zero in the capillary fringe
zone, thus making the solution impossible. Therefore, this formulation is valid only in the par-
tially saturated zone (water content less than saturated value), not in the capillary fringe.
Another way to solve this problem, which does not have the limitation discussed
above, is to formulate the flow equations in terms of soil suction ψ. For the vertical flow,
we obtain

冢 冣
∂Ψ
C(θ)   ∂ Kz(Ψ)(
∂Ψ  1)
 (4.74)
∂z ∂z ∂z
Exercise. Consider steady-state vertical infiltration from the soil surface to the water table
at depth L. The relative hydraulic conductivity of the soil is described by the following
exponential law:

kr(Ψ)  exp[ α Ψ] (4.78)

Derive an expression for the vertical distribution of h. After the first integration of the
flow equation, we obtain
 ∂Ψ 
Kz(Ψ)  1  q (4.79)
 ∂z 
where q  infiltration rate and K is a function of capillary pressure y.

Kz(Ψ)  Ksatexp[αΨ] (4.80)

where h  z  ψ. Substitution of these relationships into the flow equation leads to


 
expα(h  z)∂h  
q
(4.81)
  ∂z KSAT
or
q
exp[αh]dh  exp[αx]dx (4.82)
KSAT
After the second integration, we obtain
1 q
exp[αh]   exp[αz]  C1 (4.83)
α αKSAT
Substituting the boundary condition h(0)  0 and solving for C1 yields the following
expression for the total head h:
 q 
h(z)  1 ln (exp[αz]  1)  1 (4.84)
α  KSAT 

4.4.2 Transport in an Unsaturated Zone

The mass continuity equation for an unsaturated flow system with advection and diffu-
sion/dispersion in the aquoeus phase, diffusion in the vapor phase, partitioning between
four phases (soil, water, vapor, and (D)NAPL), and first-order degradation in the aqueous
phase can be written as

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

Subsurface Flow and Transport 4.17


∂CW ∂  D kAIR  ∂CW
v
∂CW
R   αiv  D      v   λCW (4.85)
∂t ∂xi  θW θW  ∂xi  ∂x1

where CW  aqueous phase concentration; v  pore-water velocity; θW  volumetric


moisture content; αi  longitudinal, transverse horizontal, and transverse vertical
macrodispersivities; D  Millington-Quirk dispersion coefficient; Dv  Millington-Quirk
dispersion coefficient in the vapor phase; λ  first-order degradation rate in the aquoeus
phase; and R  retardation factor;

kSOILγB kNAPLθNAPL kAIRθAIR


R1      (4.86)
θW θW θW

where kSOIL  water-soil partitioning coefficient, kSOIL  S/CW, S  weight/weight con-


centration of absorbed compound in soil, kNAPL  NAPL-water partitioning coefficient,
kNAPL  CNAPL/CW, CNAPL  concentration of compound in the NAPL phase, kAIR  vapor-
water partitioning coefficient, kAIR  V/CW, V  concentration of compound in the vapor
phase, γB  bulk density of dry soil, θAIR  volumetric vapor content, and θNAPL  volu-
metric NAPL content.
Assuming again that the diffusive fluxes are negligible compared with the macrodis-
persive and advective fluxes, the transport equation can be simplified to yield

∂CW ∂  ∂CW ∂CW


   αivR  vR λRCW (4.87)
∂t ∂xi  ∂xi  ∂xi

where vR  v/R and λR  λ/R. Note that the form of this equation is the same as the form
of the one for transport in the saturated zone; therefore, the analytical solutions presented
for the saturated transport are valid for the unsaturated conditions. This is particularly true
for the one-dimensional (vertical) transport equations, which are of primary interest in the
case of unsaturated fate and transport of compounds.

4.5 FLOW AND TRANSPORT IN VAPOR PHASE


This section, presents the soil vapor flow equations. This is followed by selected solutions
to vapor flow problems. Finally, we discuss diffusive transport in the vapor phase.

4.5.1 Soil Vapor Flow

The flow of gases in a porous medium can be described by combining the following
equations:
a. The equation of continuity is
∂ρ
∇(ρv)  n (4.88)
∂t
with ρ  vapor density, v  Darcy’s flux vector, θAIR  air-filled porosity, and t  time.
b. The perfect gas law is
pM
ρ  m   (4.89)
v RT

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

4.18 Chapter Four

with m  gas mass, V  gas volume, p  absolute pressure, M  molar mass, T 


absolute temperature ºK, and R  universal gas constant.
c. Darcy’s law for vapor flow is
v  k ∇p (4.90)
µ
where k  soil intrinsic permeability and µ  gas viscosity. The resulting governing
equation is 1953
2nµ ∂p
∇2p2    (4.91)
k ∂t
(Bruce et al., 1953).
d. The molar flux [moles/unit area-time] is given by
k
q   p∇p (4.92)
µRT
For one-dimensional flow, the governing equation reduces to
∂2p2 2θAIR µ ∂p
    (4.93)
∂x2 k ∂t
and for radial flow, the governing equation is
∂2p2 1 ∂p2 2θAIR µ ∂p
      . (4.94)
∂r2 r ∂r k ∂t
For steady state, exact analytical solutions for gas flow are obtained by using the fol-
lowing transformation (Cho, 1991):
K ( p2  pr2 )
m   (4.95)
2
where m is referred to as the discharge potential. The governing equation now becomes
Laplace's equation:
∇2m  0 (4.96)

For example, the exact solution for the point source in three-dimensional space is given by
Q
m   (4.97)
4πr
where Q  source strength and r  distance from source point.
To estimate the time required to achieve steady-state vapor flow, Johnson et al. (1990)
presented a method based on the solution of radial flow of vapor to a well. Their results
are summarized in Fig. 2 of Johnson et al. (1990 b) for values of k corresponding to sandy
soil. They also presented a method to estimate vapor flow rates, pressure distributions, and
vapor velocities in unsaturated soils based on the steady-state solution to the governing
equation of vapor radial flow: namely,
2nµ ∂p
∇2p2    (4.98)
k ∂t
The pressure p can be expressed in terms of the ambient pressure pAtm and a deviation
p’ from this pressure: p’ is equivalent to the vacuum that would be measured in the soil. If
this substitution is used in the flow equation and if we neglect the product p'2 relative to
the product pAtm p’ (linearization), then the resulting equation for radial flow is

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

Subsurface Flow and Transport 4.19


θAIRµ ∂p’ 1 ∂  ∂p’ 
     r (4.99)
kpATM ∂t r ∂r  ∂r 

The solution to this equation for the following boundary conditions is

p’  0, for r → ∞

冢 冣
∂p’ Q
limr→0 r   (4.100)
∂r 2πBk/µ
as given by
Q ∞ Q
p’   exp[ x]dx   W(u)
4πB(k/µ) u  兰
r2θAIRµ
4kp ATM
4πB(k/µ)
(4.101)

where Q is the volumetric flow rate to the vapor well. The well function W(u) is tabulat-
ed in almost all groundwater textbooks. The behavior of the integral is such that for
(r2θAIRµ/4kpAtmt)  0.001, its value is close to the asymptotic steady-state limit.
Exercise. Given the following parameters—hydraulic conductivity K  10–2 cm/sec,
air viscosity µAIR  0.018 cp, volumetric air content (equal to porosity) θ AIR  0.3, and
local pressure gradient dp/dx  0.01 atm/cm—estimate the pore-vapor velocity. From the
conversion table (Domenico and Schwartz, 1990), we have

k[darcy]  K[m/s]*1.04*105 (4.102)

Thus, k  10.4 darcy. The vapor flux is given by

kr(θ)kfluid  ∂p ∂z 
qi      ρg (4.103)
µfluid  ∂xi ∂xi 

Assuming that kr  1.0, we obtain

kr*k[darcy] ∂p  atm  10.4* 0.01


q[cm/sec]         5.78[cm/sec] (4.104)
µ[centipoise] ∂x  cm  0.018

Finally, we estimate the pore-vapor velocity v  q/θAIR  19.26 cm/s.

4.5.2 Transport in Vapor Phase

It is usual to assume that whenever advective vapor flow is present, it dominates the trans-
port process and the diffusive/dispersive processes can be neglected. In this case, the fate
and transport equation for a compound that partitions between the four phases (soil, water,
vapor, and (DNAPL) and is subject to first-order degradation in the aqueous phase is
∂V  v ∂V λ θW k V
R  AIR   AW (4.105)
∂t ∂x1 θAIR
where: V  vapor phase concentration, vAIR  pore-air velocity, θW  volumetric mois-
ture content, θAIR  volumetric vapor content, λ  first-order degradation rate in the
aqueous phase, kA  W  air-water partitioning coefficient, kAW  CW/V, and R  retar-
dation factor,

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

4.20 Chapter Four


kA  WkSOILγB kA  WkNAPLθNAPL kA  WθW
R1      (4.106)
θAIR θAIR θAIR

where: kSOIL  water-soil partitioning coefficient, kSOIL  S/CW, S  weight/weight con-


centration of absorbed compound in soil, kNAPL  NAPL-water partitioning coefficient,
kNAPL  CNAPL/CW, CNAPL  concentration of compound in the NAPL phase, kAIR  vapor-
water partitioning coefficient, kAIR  V/CW, V  concentration of compound in the vapor
phase, γB  bulk density of dry soil, θAIR  volumetric vapor content, and θNAPL  volu-
metric NAPL content.
Division of the transport equation by the retardation factor yields

∂V ∂V
  vAIR,R   λ*RV (4.107)
∂t ∂x1

where vAir,R  vAIR/R and λR  kA  WλθW/(θAIR R). Note that the form of this equation is the
same as the form of the one for transport in the saturated zone, except for the absence of
the dispersive term. Therefore, the analytical solutions presented for the saturated trans-
port are valid for the vapor transport conditions.
When there is no advective transport in the vapor phase, the transport equation must
include diffusive fluxes in vapor and aqueous phases to yield
∂V  ∂ Dv ∂V  Dk ∂V  θW
R    A  W   λ  kA  WV (4.108)
∂t ∂xi  ∂xi ∂xi  θAIR
where D  Millington-Quirk dispersion coefficient for the aqueous phase, and Dv 
Millington-Quirk dispersion coefficient in the vapor phase. Assuming that the dispersion
coefficients do not vary in space leads to the following form of the transport equation:
∂V ∂2V
  D̂2  λ̂V (4.109)
∂t ∂x i
where
Dv  DkA  W
D̂   (4.110)
R
and
θW
λ̂  λ (4.111)
θAIRR
Again, we note the similarity of this fate and transport equation to the one presented
for saturated transport and conclude that all the analytical solutions presented in Sec. 4.2
can, in principle, be used to analyze vapor phase transport.

Exercise. Given that water saturation Sw  0.20, porosity n  0.4, compound concentra-
tion in soil vapor at depth L  2 m Co  100 mg/L, compound concentration at the soil
surface Cs  0.01, and molecular diffusion coefficient of the compound Do 
0.087cm2/sec, estimate the compound mass flux in the vapor phase at the soil surface. The
effective vapor-phase diffusion coefficient is given by

θ 3.33
D  Do A  0.0026 cm3/s (4.20)
n2
and the mass flux of compound A is estimated as follows:

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

Subsurface Flow and Transport 4.21


∂C Co  Cs  0.0026[cm2/s] 100*10 [mg/cm ]  1.3[mg/(cm2s)] (4.112)
3 3
q  D  D  
∂x L 200[cm]

Exercise. Consider advective transport of a compound in vapor phase. The compound


partitions between vapor and aqueous phases according to the relationship

VKC (4.113)

where V  concentration in vapor phase, C  concentration in aqueous phase, and K 


partitioning coefficient. Given vapor flux q, porosity n, and water saturation Sw, estimate
the apparent (retarded) velocity of the compound. From the advective transport equation,
we have

∂CT ∂V
   q (4.114)
∂t ∂x
where
 SW 
CT  C SW n  V(1  SW)n  Vn (1  SW)  1 (4.115)
 K(1  SW 

Thus, for nonretarded tracers, we have


∂V ∂V
  V (4.116)
∂x
where pore-vapor velocity is
q
v   (4.117)
n(1  Sw)
whereas for retarded compounds, we have

∂V ∂V
  vR (4.118)
∂t ∂x

where the retarded velocity is given by

v v
vR     SW 1 (4.119)
R

K(1  SW)
Exercise. Consider diffusive vertical transport of a compound in vapor phase. The com-
pound is subject to first-order degradation in the aqueous phase at rate λ and to partitions
between vapor and aqueous phases according to the following relationship:

VKC (4.113)

where V  concentration in vapor phase, C  concentration in aqueous phase, and K 


partitioning coefficient. At the depth of 100 cm below the ground surface, the vapor con-
centration of the compound was measured to be Vo, whereas at the ground surface the con-
centration was Vs. Given the compound’s diffusion coefficient in vapor phase Do, porosi-
ty n, and water saturation Sw, estimate the diffusive flux of the compound at the soil sur-
face. The relevant mass transport equation is given by

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

4.22 Chapter Four

∂ V λnS C  0
D
2
 (4.120)
∂x2 w

where
θ 3.33
D  Do A (4.19)
n2
Substituting C  V/K into the mass transport equation leads to
∂2V
2 λ*2V  0 (4.121)
∂x
where
λnS
λ*  w
2
(4.122)
KD
We solve the modified mass transport equation to obtain

V(x)  C1exp[ λ∗x]  C2exp[λ∗x] (4.123)

where constants C1 and C2 are obtained from the boundary conditions

V(0)  Vo, and V(100)  Vs (4.124)

The compound's mass flux at the soil surface is estimated from


∂V(x)
q  D @x  100 (4.125)
∂x

REFERENCES
Bear, J., and Y. Bachmat, Introduction to Modeling of Transport Phenomena in Porous Media,
Kluwer Academic, Dordrecht, The Netherlands, 1990.
Bear, J., Dynamics of Fluids in Porous Media, Dover Publications, New York,k 1972.
Bear, J., Hydraulics of Groundwater, McGraw Hill, New York, 1979.
Bouwer H., Groundwater Hydrology, McGraw Hill, New York, 1978.
Bruce, G. H., D. W. Peaceman, and H. H. Rachford, Jr., 1953. “Calculations of Unsteady-State Gas
Flow Through Porous Media,” Petroleum Transactions, AIME, 198: 1953.
Bureau of Reclamation, Ground Water Manual (Reprint), U.S. Department of the Interior
Washington, DC, 1995.
Cedergren H. R., Seepage, Drainage, and Flow Nets, 3rd ed., John Wiley & Sons, Inc., New York,
1989.
Charbeneau, R. J., “Kinematic Models for Soil Moisture and Solute Transport,” Water Resources
Research, 20: 699–706, June, 1984.
Chirlin, G. R., “A Critique of the Hvorslev Method for Slug Test Analysis: The Fully Penetrating
Well,” Ground Water Monitoring Review, 130–138, 1989.
Cho, J. S., 1991. Forced Air Ventilation for Remediation of Unsaturated Soils Contaminated by
VOC., Publication No. EPA/600/S2–91/016, U.S. Environmental Protection Agency, Washington,
D.C.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

Subsurface Flow and Transport 4.23

Dagam, G., “Solute transport in heterogeneous porous formations,” Journal of Fluid Mechanics,
145: 151–177, 1984.
De Josselin Jong, G. “Singularity Distribution for the Analysis of Multiple-Fluid Flow Through
Porous Media,” Journal of Geophysical Research, 65: 3739–3758, 1960.
De Marsily G., Quantitative Hydrogeology—Groundwater Hydrology for Engineers, Academic
Press, San Diego, CA,
De Smedt F., and P. J. Wirenga, “Solute Transport Through Soil With Nonuniform Water Content,”
Soil Science Society of America Journal, 42. (1): 1978.
Domenico, P. A., and F. W. Schwartz, Physical and Chemical Hydrogeology, John Wiley & Sons,
New York, 1990.
Dullien F. A. L., Porous Media: Fluid Transport and Pore Structure, 2nd ed., Academic Press, San
Diego, CA, 1992.
Edelman J. H., Groundwater Hydraulics of Extensive Aquifers, 2nd., International Institute for
Land Reclamation and Improvement, Bulletin No. 13, The Netherland, 1983.
Fetter, C. W., Contaminant Hydrogeology, Macmillan, New York, 1993.
Fetter, C. W., Applied Hydrogeology, Simon & Schuster, Company Englewood, NJ, 1994.
Freeze, R. A., and J. A. Cherry, Groundwater, Prentice-Hall, Englewood Cliffs, NJ, 1979.
Gelhar, L. W. and C. L. Axness. “Three-Dimensional Stochastic Analysis of Macrodispersion in
Aquifers”, Water Resources Research, 19 (1): 161–180, 1983.
Germann, P. F., M. S. Smith, and G. W. Thomas, “Kinematic Wave Approximation to the Transport
of Escherichia coli in the Vadose Zone,” Water Resources Research. 23 (7), 1281–1287, 1987.
Girinsky, N. K. Determination of the Coefficient of Permeability, Gosgeolizdat, 1950.
Grubb, S. “Analytical Model for Estimation of Steady-state Capture Zones of Pumping Wells in
Confined and Unconfined Aquifers,” Ground Water, 31(1) 27–32, 1993.
Hantush, M. S., Hydraulics of Wells, “in Advances in Hydroscience,” V. T. Chow, ed., Academic
Press, New York, 1964.
Hantush, M. S. “Growth and Decay of Groundwater-mounds in Response to Uniform Percolation.”
Water Resources Research, 3 (1): 227–234, 1967.
Harr, M. E., Groundwater and Seepage, Dover Publications, New York, 1990.
Haverkmp, R., M. Vauclin, J. Touma, P. J. Wierenga, and G. Vachaud, A “Comparison of
Numerical Simulation Models for One-Dimensional Infiltration,” Soil Science Society America
Journal, 41: 285–294, 1977.
Hinchee, R. E. ed., Air Sparging for Site Remediation, Lewis Publishers, Boca Raton, FL, 1994.
Huisman, L., Groundwater Recovery, Winchester Press, New York, 1972.
Hunt, B. “Seepage to Collection Gallery Near Seacoast,” Water Resources Research, 21: 311–316,
1985.
Hunt, B., Mathematical Analysis of Groundwater Resources, Butterworths, London, UK, 1983.
Jaffe, P. R., and R. A. Ferrara, “Desorption Kinetics in Modeling of Toxic Chemicals,” Journal of
Environmental Engineering, American Society of Civil Engineers,109: 859–867, 1983.
Javandel, I., and C. F. Tsang, “Capture-zone Type Curves: A Tool for Aquifer Cleanup.” Ground
Water, 24: 616–625, 1985.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

4.24 Chapter Four

I., C. Doughty, and C. F. Tsang, Groundwater Transport: Handbook of Mathematical Models,


American Geophysical Union, Washington, DC, 1987.
Johnson, P. C., M. W. Kemblowski, and J. D. Colthart, “Quantitative Analysis for the Cleanup of
Hydrocarbon-Contaminated Soils by in-situ Soil Venting,” Ground Water, 1990.
Johnson, P. C., C. C. Stanley, M. W. Kemblowski, D. L. Byers, and J. D. Colthart. A “Practical
Approach to the Design, Operation, and Monitoring of in situ Soil-Venting Systems.” Ground
Water Monitoring Review, Spring, 1990.
Jury, W. A., “Chemical Transport Modeling: Current Approaches and Unresolved Problems,”
Chemical Mobility and Reactivity in Soil Systems, 1983, pp. 49–64.
Jury, W. A., R. Grover, W. F. Spencer, and W. J. Farmer, “Modeling Vapor Losses of Soil
Incorporated Triallate,” Soil Science Society of America Journal, 44: 445–450, 1980.
Jury. W. A., W. F. Spencer, and W. J. Farmer, “Use of Models for Assessing Relative Volatility,
Mobility, and Persistence of Pesticides and other Trace Organics in Soil Systems.” Hazard
Assessment of Chemicals: Current Developments, Vol. 2, 1983.
Keely, J. F. and C. F. Tsang, “Velocity Plots and Capture Zones of Pumping Centers for Ground-
Water Investigations.” Ground Water, 21: 701–714, 1983.
Kishi, Y. and Y. Fukuo, “Studies on Salinization of Groundwater,” I. Journal of Hydrology, 35:
1–29, 1977.
Kool, J. B., J. C. Parker, and M. T. van Genchten, “Parameter Estimation for Unsaturated Flow and
Transport Modles—A Review.” Journal of Hydrology, 91: 255–293, 1987.
Kozeny, J., Thorie und Berchnung der Brunnen. Wasserkraft und Wasserwirtschaft, Nos. 8–10,
1933.
Marino, M. A., “Artificial Groundwater Recharge: I. Circular Recharging Area,” Journal of
Hydrology, 25: 201–208, 1975.
Marshall, T. J., J. W. Holmes, and C. W. Rose, Soil Physics, 3rd ed., Cambridge University Press,
Cambridge, UK, 1996.
McElwee, C., and M. Kemblowski, “Theory and Application of an Approximate Model of
Saltwater Upconing in Aquifers,” Journal of Hydrology, 115: pp 139–163, 1990.
McWhorter, D. B. Steady and Unsteady Flow of Fresh Water in Saline Aquifers, Water
Management Technical Report No. 20, Council of U.S. Universities for Soil and Water
Development in Arid and Sub-Humid Areas, 1972.
Musa, M. and M. W. Kemblowski. “Effective Capture Zone for a Single Well,” Submitted to
Ground Water July 1994.
Newsom, J. M., and J. L. Wilson, “Flow of Ground Water to a Well Near a Stream - Effect of
Ambient Ground-water Flow Direction.” Ground Water, 25: 703–711, 1988.
Oberlander, P. L. and R. W. Nelson, “An Idealized Ground-Water Flow and Chemical Transport
Model (S-PATHS),” Ground Water, 22: 441–449, 1984.
Ostendorf, D. W., R. R. Noss, and D. O. Lederer, “Landfill Leachate Migration through Shallow
Unconfined Aquifers,” Water Resources Research, 20: 291–296, 1984.
Palmer, C. M., Principles of Contaminant Hydrogeology, Lewis Publishing, Chelsea, MI, 1992.
Pankow, J. F., R. L. Johnson, and J. A. Cherry, “Air Sparging in Gate Wells in Cutoff Walls and
Trenches for Control of Plumes of Volatile Organic Compounds (VOCs),” Ground Water, 31:
654–663, 1993.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

Subsurface Flow and Transport 4.25

Parker, J. C., and M. T. van Genuchten, Determining Transport Parameters from Laboratory ad
Field Tracer Experiments, Virginia Agricultural Experiment Station Bulletin No. 84–3, Virginia
Polytechnic Institute and State University, Blacksburg, 1984.
Parker, J. C., and M. T. van Genuchten, “Flux-averaged and Volume-Averaged Concentrations in
Coninuum Approaches to Solute Transport,” Water Resources Research, 20: 886–872, 1984.
Parker, J. C., K. Unlu, and M. W. Kemblowski, “A Monte Carlo Model to Assess Effects of Land
Disposed E & P Waste on Groundwater,” SPE Annual Technical Conference & Exhibition, 1993.
Philip, J. R. “The theory of infiltration: 1. The infiltration equation and its solution.” Soil Science.
83: 345–357, 1957.
Raudkivi, A. J., and R. A. Callander, Analysis of Groundwater Flow, Edward Arnold, London, UK,
1976.
Rosenshein, J., and G. D. Bennet, eds., Groundwater Hydraulics, American Geophysical Union,
Washington, DC, 1984.
Rubin, H., and G. F. Pinder., “Approximate Analysis of Upconing.” Advances in Water Resources, 1
(2): 97–101, 1977.
Sallam, A., W. A. Jury, and J. Letey, “Measurement of Gas Diffusion Coefficient Under Relatively
Low Air-Filled Porosity,” Soil Science Society of America Journal 48:3–6, 1983.
Schiegg, H. O., “Considerations on Water,. Oil and Air in Porous Media,” Water Science
Technology, 17: 467–476. 1984.
Shafer, J. M., “Reverse Pathline Calculation of Time-Related Capture Zones in Nonuniform Flow,”
Ground Water 25: 283–289, 1987.
Sikkema, P. C. and J. C. Van Dam, “Analytical Formulae for the Shape of the Interface in a Semi-
Confined Aquifer,” Journal of Hydrology, 56: 201–220, 1982.
Sposito, G., and W. A. Jury, “Inspectional Analysis in the Theory of Water Flow Through
Unsaturated Soil,” Soil Science Society of America Journal, 42 (1): 1985.
Sposito, G., “Chemical Models of Inorganic Pollutants in Soils,” CRC Critical Reviews in
Environmental Control, 15 (1): 1–24, undated.
Strausberg, S. I. “Estimating Distances to Hydrologic Boundaries from Discharging Well Data,”
19th Annual Meeting of the Rocky Mountain Section of the Geological Society of America, Las
Vegas, NV, 1966.
Thornton, J. S., and W. L. Wootan, Jr., “Venting for the Removal of Hydrocarbon Vapors from
Gasoline Contaminated Soil,” Journal of Environmental Science and Health, A17 (1), 31–44, 1982.
Todd, D. K., Groundwater Hydrology, 2th John Wiley & Sons, New York, 1980.
Todd, D. K., “Salt-Water Intrusion and Its Control.” Journal of the American Water Works
Associations, 180–187, 1973.
U. S. Department of Agriculture Agricultural Research Service, Analytical Solutions of the One-
Dimensional Convective-Dispersive Solute Transport Equation, Technical Bulletin No. 1661,
Unlu, K., M. W. Kemblowski, J. C. Parker, D. Stevens, P. K. Chong, and I. Kamil, “A Screening
Model for Effects of Land-Disposed Wastes on Groundwater Quality” Journal of Contaminant
Hydrology, 11: 27–49, 1992. Washington, DC, 1982
Van Genuchten M. T. and W. J. Alves, “Analytic Solution of the One–Dimensional Convective
Solute Transport Equation, Technical Bulletin. 1661, U.S. Departament of Agriculture,
Washington D.C., 1982.
Ward, C. H., M. B. Tomson, P. B. Bedient, and M. D. Lee, “Transport and Fate Processes in the
Subsurface,” Water Resources Symposium, Vol 13, WARSAG, 1987.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
SUBSURFACE FLOW AND TRANSPORT

4.26 Chapter Four

Warrick, A. W. , J. W. Biggar, and D. R. Nielsen, “Simultaneous Solute and Water Transfer for an
Unsaturated Soil,” Water Resources Research. 7: 1216–1225, 1971.
Watson, K. K., and M. J. Jones, “Algebraic Equations for Solute Movement During Absorption,”
Water Resources Research, 20: 1131–1136, 1984.
Wilson, J. L., and L. W. Gelhar, “Analysis of Longitudinal Dispersion in Unsaturated Flow 1: The
Analytical Method,” Water Resources Research, 17 (1): 122-130, 1984.
Wilson, J. L., and P. J. Miller, “Two-Dimensional Plume in Uniform Ground-Water Flow -
Discussion,” Journal of the Hydraulics Division, American Society of Civil Engineers, 103
(HY12): 1567–1570, 1979.
Wilson, J. L., and P. J. Miller., “Two-Dimensional Plume in Uniform Ground-Water Flow,” Journal
of the Hydraulics Division, American Society of Civil Engineers, 104 (HY4): 503–514, 1978.
Wirojanagud, P., and R. Charbeneau., “Saltwater Upconing in Unconfined Aquifers,” Journal of
Hydraulic Engineering, American Society of Civil Engineers, 111: 417–434, 1985.
Yeh, G. T., Analytical Transient One-, Two-, and Three-Dimensional Simulation of Waste Transport
in the Aquifer System, Environmental Sciences Division Publication No. 1439, Oak Ridge
National Laboratory, Oak Ridge, TN, 1981.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

You might also like