ArticlePDF Available

Analytical Models for the Design of Iron-Based Permeable Reactive Barriers

Authors:

Abstract and Figures

The preliminary design of iron-based permeable reactive barriers is often accomplished using analytic expressions for one-dimensional groundwater flow and contaminant transport. Typically, one or more of the governing processes is simplified or neglected to facilitate development of a tractable solution. This paper presents a set of improved design equations that include the effects of dispersion, finite domain boundary, sequential decay, and production processes, and increased flow through high conductivity barriers. When applied to realistic example problems, application of the expanded design equations typically results in the specification of a larger permeable reactive barrier thickness than obtained using conventional approaches.
Content may be subject to copyright.
Analytical Models for the Design of Iron-Based
Permeable Reactive Barriers
Alan J. Rabideau, P.E., M.ASCE
1
; Raghavendra Suribhatla
2
; and James R. Craig
3
Abstract: The preliminary design of iron-based permeable reactive barriers is often accomplished using analytic expressions for one-
dimensional groundwater flow and contaminant transport. Typically, one or more of the governing processes is simplified or neglected to
facilitate development of a tractable solution. This paper presents a set of improved design equations that include the effects of dispersion,
finite domain boundary, sequential decay, and production processes, and increased flow through high conductivity barriers. When applied
to realistic example problems, application of the expanded design equations typically results in the specification of a larger permeable
reactive barrier thickness than obtained using conventional approaches.
DOI: 10.1061/ASCE0733-93722005131:111589
CE Database subject headings: Barriers; Design; Analytical techniques; Ground-water pollution
.
Introduction
A popular technology for remediating contaminated groundwater
is to emplace a permeable reactive barrier PRB in the path of
the plume e.g., Starr and Cherry 1994; USEPA 1998. According
to Environmental Technologies Inc. ETI, over 125 PRB instal-
lations have been reported worldwide ETI 2005a. The large ma-
jority of these applications utilize zero-valent iron to promote the
degradation of chlorinated organic compounds.
For chlorinated ethenes such as tertrachloroethylene PCE,
the zero-valent iron induces rapid transformations within the PRB
that are typically described as a network of coupled first-order
decay reactions e.g., Roberts et al. 1996; Scherer et al. 2000.
One of the key design tasks is to determine the PRB thickness
in the direction of groundwater flow needed to provide the
residence time to reduce the concentrations of target compounds
to the desired effluent concentration. The most straightforward
approach to PRB design is to conduct bench or pilot testing to
directly determine the required residence times e.g., Gavaskar et
al. 1997, 1998, 2000. Alternatively, mathematical modeling can
be used to extrapolate the PRB performance using reaction rate
constants measured from small scale laboratory experiments
and/or obtained from the literature.
For PRB design, a common simplification is to apply a single-
solute transport model to the parent compound, with the produc-
tion of reaction products represented by approximate correction
factors and/or safety factors e.g., Tratnyek et al. 1997. In some
cases, complex geochemical models have been developed to con-
sider multiple degradation pathways and other chemical consid-
erations such as the impact of mineral precipitation on barrier
performance e.g., Yabusaki et al. 2001. However, for prelimi-
nary design, it is desirable to maintain an analytical framework
that can be implemented simply and easily, while maintaining a
suitably accurate treatment of the chemical degradation processes.
This paper presents several improvements to current models that
increase the degree of accuracy and conservatism while retaining
a simple analytic framework that can be implemented quickly and
easily.
Transport Models
Current Design Practice
For PRBs based on engineered transformation reactions, transport
through the barrier is typically modeled using a one-dimensional
form of the advective–dispersive–reactive equation ADRE. For
the transport of a single decaying solute in a one-dimensional
homogeneous porous medium, the governing ADRE is
c
t
=−
c
x
+ D
2
c
x
2
k
1
where caqueous phase contaminant concentration M/L
3
;
ttime T; xdistance from the entrance of the PRB L;
␯⫽interstitial fluid velocity L/T; Ddispersion coefficient
L
2
/T; and k
first-order decay constant l/T.
Application of Eq. 1 to a PRB setting is commonly accom-
plished by neglecting the dispersion term and treating the PRB as
an ideal plug flow reactor, which leads to the following simple
design equation e.g., Gavaskar et al. 1998; USEPA 1998:
cx = L
c
0
= exp
k
L
2
where Lbarrier thickness L; and c
0
constant contaminant
concentration entering the barrier M/L
3
.
1
Associate Professor, Dept. of Civil, Structural, and Environmental
Engineering, Univ. at Buffalo, Buffalo, NY 14260 corresponding
author. E-mail: rabideau@eng.buffalo.edu
2
Research Assistant, Dept. of Civil, Structural, and Environmental
Engineering, Univ. at Buffalo, Buffalo, NY 14260.
3
Assistant Professor, Dept. of Civil Engineering, Univ. of Waterloo,
Waterloo, Ontario, Canada N2L 3G1.
Note. Discussion open until April 1, 2006. Separate discussions must
be submitted for individual papers. To extend the closing date by one
month, a written request must be filed with the ASCE Managing Editor.
The manuscript for this paper was submitted for review and possible
publication on August 2, 2004; approved on April 6, 2005. This paper is
part of the Journal of Environmental Engineering, Vol. 131, No. 11,
November 1, 2005. ©ASCE, ISSN 0733-9372/2005/11-1589–1597/
$25.00.
JOURNAL OF ENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2005 / 1589
In addition to neglecting dispersion, the form of Eq. 2
implies a constant concentration entrance boundary condition,
and a semi-infinite domain. The velocity term is frequently
assigned based on the local groundwater flow prior to barrier
construction, although it is sometimes reduced to reflect the
higher porosity associated with iron-based PRBs.
For chlorinated solvents, it is often necessary to consider the
production and transport of toxic reactions products. Retaining
the assumptions of Eq. 2, the resulting problem can be
expressed as a set of coupled linear ordinary differential equations
ODEs
c
i
=−k
i
c
i
+
j1
y
ij
k
j
c
j
3
i =1,2,¯ n
where nnumber of solutes in the decay chain; ␶⫽PRB residence
time L / T; ksolute decay constant l/T; and y
ij
fractional
yield of solute i produced by the decay of solute j.
In practice, the system represented by Eq. 3 is often solved
numerically using a standard stiff ODE solver e.g., ETI 2005b,
with the barrier thickness L adjusted to reduce the contaminant
concentrations to the cleanup targets. This approach is rigorous
with respect to the PRB chemistry, but simplifies the hydrody-
namics of the PRB by neglecting dispersion and exit boundary
effects. Furthermore, the placement of a high-permeability PRB
may change the regional groundwater flow patterns in a manner
that results in increased flow through the barrier. While the
effect of these factors may be small, the common simplifications
are nonconservative with respect to the predicted concentrations
exiting the barrier e.g., Eykholt and Sivavec 1995.Inthe
following sections, several modifications are presented to address
these issues while retaining the simple analytic framework.
Governing Equations for Chlorinated Ethenes
Roberts et al. 1996 proposed a conceptual model for the
transformation of chlorinated ethenes by zero-valent iron that has
been widely accepted. Although the general framework includes
both series and parallel reactions i.e., daughter products with
multiple parents, laboratory bench testing to characterize PCE/
trichloroethylence TCE transformation has indicated that the
reactions involving the production of toxic products may be
considered in terms of a single series of transformations
i.e., negligible “branching”兲共ETI 2005b. The resulting ADRE is
given by
c
i
t
=−
c
i
x
+ D
2
c
i
x
2
k
i
c
i
+ y
i
k
i−1
c
i−1
4
i =1,2,¯ ,n
where the yield coefficient y
i
contains a single subscript to indi-
cate that reaction product i is produced only by the single parent
compound i−1 i.e., y
i
is equivalent to y
i,i−1
in Eq. 3, with all
other y
i,j
terms= zero.
The restriction of Eq. 4 to a single decay chain does not
imply that other decay pathways are absent, but only that they do
not lead to significant production of the toxic reaction products of
interest. For example, other end products of PCE degradation
could include acetylene and ethylene e.g., Roberts et al. 1996,
which are typically not of regulatory concern. The form of Eq. 4
also implies that the dispersion coefficients are equal for all
solutes; i.e., the solute-specific contribution of molecular
diffusion is much less than the hydrodynamic component of
dispersion.
The solution to Eq. 4 requires two boundary conditions and
an initial condition. For field conditions, the contaminant source
can again be represented by a constant-concentration first type
entrance condition
c
i
x =0,t = c
0i
5
i =1,2,¯ ,n
where c
0i
concentration of solute i entering the PRB.
For laboratory columns, which are often operated by measur-
ing “steady-state” resident concentrations along the column
length, the mass-conserving third-type condition may be preferred
c
i
x =0,t = c
0i
+
D
c
i
x
x =0,t
6
i =1,2,¯ ,n
For the barrier exit, a popular choice is to treat the PRB as a
semi-infinite medium, which leads to the following boundary
condition:
c
i
x
x = ,t =0
7
i =1,2,¯ ,n
The application of Eq. 7 is appealing because it facilitates the
development of closed form solutions for both transient and
steady-state conditions. However, it incorrectly implies that the
reactive properties of the PRB extend into the adjoining aquifer. A
more plausible condition for advection-dominated transport is the
zero-gradient exit condition, which is given by
c
i
x
x = L,t =0
8
i =1,2,¯ ,n
where L=thickness of the reactive medium.
In addition to providing a more conceptually appealing
description of the transition occurring at the PRB exit, Eq. 8
leads to higher predicted effluent concentrations when the con-
stant concentration influent condition is used, and thus provides
more conservative predictions for the purpose of PRB design.
For PRB applications, it is common to assign the initial con-
dition based on the assumption that the contaminant is initially
absent within the barrier
c
i
x,t =0 =0
9
i =1,2,¯ ,n
Solutions
For design purposes, the consideration of steady-state conditions
is customary and leads to conservative predictions. Closed-form
steady-state solutions to Eq. 1 for several combinations of
boundary conditions are summarized in the Appendix.
Sun et al. 1999 developed a simple transformation procedure
that can be used to extend solutions based on single solute
1590 / JOURNAL OF ENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2005
transport to the case of multiple solutes subject to sequential first-
order decay. A transformed concentration variable a is defined
as
a
i
= c
i
+
j=1
i−1
l=j
i−1
y
l+1
k
l
k
l
k
i
c
j
10
i =2,3,¯ ,n
By applying this transformation to Eq. 4, the governing ADRE
for each transformed variable a
i
is reduced to the single-solute
form Eq. 1兲兴. The entrance boundary condition is similarly
transformed as
a
i0
= c
i0
+
j=1
i−1
l=j
i−1
y
l+1
k
l
k
l
k
i
c
j0
11
i =2,3,¯ ,n
Once the appropriate single-solute solution to Eq. 1 is identified,
a back transformation is performed to develop closed-form
solutions for each solute in the decay chain
c
i
= a
i
j=1
i−1
l=j
i−1
y
l+1
k
l
k
l
k
i
c
j
12
i =2,3,¯ ,n
The multisolute expansions for a four-solute chain are given in
the Appendix, and the resulting steady-state solutions to Eq. 1
can be readily implemented in a spreadsheet. For known influent
concentrations and specified constraints on the effluent concentra-
tions, the design equations can be solved iteratively to determine
the minimum required barrier thickness L.
Example
Consider the case of a PRB designed to treat a plume containing
PCE. Within the PRB, the pathway of interest is the sequential
decay of PCE to TCE, cis1,2,dichlorethylene DCE, and vinyl
chloride VC. The federal drinking water standards for PCE,
TCE, DCE, and VC are 5, 5, 70, and 2 g /L, respectively. For
this system, the third reaction product VC often controls the
system design and a four-solute model is needed. Fig. 1 shows the
simulated effluent concentrations for a typical PRB system based
on the parameters summarized in Table 1, for a variety of bound-
ary conditions.
Although small differences are apparent between the semi-
infinite and finite boundary conditions, the most significant
difference in predicted PRB performance occurs when dispersion
is included in the model. For advection-dominated groundwater
systems, the dispersion coefficient is typically assumed to be
linearly proportional to the velocity, i.e.,
D ␣␯ 13
where =D /␯⫽dispersivity of the medium L.
Field measurements of for zero-valent iron systems were not
located in the literature. In recognition of the scale dependence of
dispersion processes, dispersivity values are often estimated as a
fraction of the solute travel distance e.g., Gelhar et al. 1992. For
the example simulations shown in Fig. 1, was set at a value
equal to 2% of the domain length 0.02L, based loosely on the
Fig. 1. a Predicted permeable reactive barrier performance for
parameters listed in Table 1, with dispersion neglected and
semi-infinite exit boundary; b predicted permeable reactive barrier
performance for parameters listed in Table 1, with dispersion
included and finite zero-gradient exit boundary; and c predicted
permeable reactive barrier performance for parameters listed in
Table 1, with dispersion included and semi-infinite exit boundary
JOURNAL OF ENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2005 / 1591
results of column tests conducted over length scales similar to an
installed PRB Sivavec et al. 1996; Casey et al. 2000; Farrell et al.
2000.
Comparison of the various solutions indicates that the model
in which dispersion is neglected Eq. 2兲兴 predicts effluent con-
centrations that are significantly lower than the others. Applica-
tion of the finite exit boundary condition Eq. 7兲兴 leads to higher
predicted effluent concentrations, but the resulting difference is
small compared to the effect of neglecting dispersion. To further
explore the impact of dispersion, an optimization was performed
to identify the smallest PRB thickness that would produce effluent
concentrations at or below the MCL values for the transport
parameters given in Table 1. The optimization approach may be
expressed as follows:
Minimize
L = barrier thickness
subject to
c
i
L c
ei
i =1,2,¯ ,n 14
L L
min
where c
ei
=design effluent concentration for solute i e.g., the
drinking water standard and L
min
=minimum barrier thickness
provided to facilitate convergence.
For the reaction parameters given in Table 1, the “baseline” L
value was calculated at 1.05 m for the advection-only model.
For the recommended finite thickness design model based on
Eq. 7兲兴, the calculated L values were 1.10, 1.23, and 1.98 m for
the /L ratios of 0.005, 0.02, and 0.1, respectively other param-
eters retained at Table 1 values. Application of the semi-infinite
exit condition yielded slightly smaller L values of 1.09, 1.21, and
1.85 m. It is noteworthy that application of the commonly used
“10%” dispersivity relationship e.g., Aziz et al. 2000 resulted in
a calculated PRB thickness nearly twice that determined by the
model that neglected dispersion. Application of a more realistic
dispersivity ratio of 2% increased the design PRB thickness by
23% over the pure advection scenario.
Estimating the Velocity Term
Although the inclusion of dispersion significantly impacts the pre-
dicted PRB performance, the most sensitive variables, in general,
are the groundwater velocity and the reaction constants. The
velocity within the PRB is often estimated based on the precon-
struction groundwater flow patterns. In some cases, a correction
may be applied to account for the higher porosity within the PRB,
which may be 2–3 times the aquifer porosity. A more rigorous
analysis would utilize a numerical model of groundwater flow to
model the local velocity patterns e.g., Starr and Cherry 1994;
Gupta and Fox 1999; Das 2002. This approach is appealing be-
cause it accounts for the tendency of the high-permeability barrier
to preferentially channel groundwater flow, leading to increased
interior velocities. However, implementation of a numerical
model for this purpose can be cumbersome, and the accuracy of
the computed velocities may be subject to significant errors from
numerical discretization.
For aquifers that are characterized by horizontal regional flow,
an appealing alternative is the analytic element method AEM,
which can simulate both regional and local hydrologic features
without discretization artifacts Strack 1989. By configuring a
model to appropriately capture the regional flow features that
control local flow patterns, the degree of flow distortion caused
by the placement of a PRB can be easily quantified. Applications
of the AEM are usually limited to steady-state flow in a single
layer aquifer, which is appropriate for the design of systems
that contain fully penetrating barriers. An advantage of the AEM
is its applicability to confined, unconfined, or partially confined
conditions.
Emplacement of a PRB that has a different hydraulic conduc-
tivity from the regional aquifer can be represented using
“inhomogeneity” elements of various geometries. Fitts 1997
proposed the use of line elements to represent the influence of
emplaced barriers on regional flow field. However, to facilitate
the calculation of the velocity within the PRB, the simple solution
for an elliptical inhomogeneity in uniform flow developed inde-
pendently by Strack 1989 and Obdam and Veling 1987 is a
more appealing choice. The elliptical geometry is a good repre-
sentation of a typical fully penetrating PRB configuration and
avoids the singularities and/or numerical artifacts associated with
modeling barriers using line segments in an analytical or numeri-
cal model. While some distortion of the barrier geometry may
occur at the ends of the ellipse, these areas usually are outside the
reactive zone of interest. Also, when used in a regional model
where the uniform flow assumption is applicable, the elliptical
geometry captures the residence time distribution with good
accuracy, with shorter residence times calculated near the edges
of the PRB. The public domain AEM program Visual Bluebird
Craig and Matott 2004 supports high-order elliptical elements
as developed by Surlbhatla et al. 2004 and provides user-
specified velocity transects useful for characterizing the behavior
of barrier systems.
For preliminary design purposes, a simple equation can be
developed using the conceptual model of a single fully penetrat-
ing elliptical hydraulic conductivity inhomogeneity placed in a
uniform regional flow field. For the conceptual model shown in
Fig. 2, the discharge across the PRB is given by
Table 1. Parameters used in Permeable Reactive Barrier PRB
Simulation
Parameter Value
PRB thickness L 1.25 m
Groundwater velocity v 0.5 m/ day
Scaled dispersivity D /L
v
0.02
Decay rates
a
PCE k
1
3.61 day
−1
TCE k
2
5.73 day
−1
cDCE k
3
2.97 day
−1
VC k
4
3.6101 day
−1
Conversion ratio
a
PCE–TCE y
12
0.40
TCE–DC y
23
0.02
DC–VC y
34
0.01
Entrance concentration
a
PCE c
01
10,000 g/L
TCE c
02
1,000 g/L
cDCE c
03
100 g/L
VC c
04
10 g/L
a
Based on ETI 2005b.
1592 / JOURNAL OF ENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2005
Q = a + bQ
0
k
e
2
sin
2
bk + ak
e
2
+
k
e
2
cos
2
ak + bk
e
2
1/2
15
where Q =flow per unit width inside the PRB L
2
/T;
Q
0
=regional flow oriented at an angle from the longitudinal
axis of the PRB L
2
/T; a and b= long and short semiaxes of the
elliptical PRB, respectively; and k
e
and k= hydraulic conductivi-
ties of the elliptical PRB and the aquifer, respectively L/T.
The derivation of Eq. 15 is given in the Appendix. Although
the effect of aquifer recharge is not considered, the effects
of recharge are negligible if the regional flow is estimated from
field observations of the piezometric head, which is typically
accomplished using the following approximation:
Q
0
h
2
h
1
x
kh
ave
16
where h
2
and h
1
=observed piezometric heads at upgradient and
downgradient locations L, respectively; x =separation distance
in the direction of flow; k=mean aquifer hydraulic conductivity
L/T; and h
ave
=average saturated thickness.
Eq. 15 can be used to examine the influence of PRB configu-
ration on velocity. The most significant increases in the PRB dis-
charge relative to the regional flow occur when the PRB is not
oriented perpendicular to the regional flow and when the aspect
ratio a/ b is small. However, when the PRB is not perpendicular
to the regional flow, the direction of flow within the PRB and
the resulting distribution of residence times is complex. For the
simpler case of flow perpendicular to the PRB, the velocity
normal to the PRB can be determined from
Q = a + bQ
0
k
e
bk + ak
e
17
=
Q
h
PRB
18
where = porosity of the PRB; h
PRB
=average saturated thickness
within the PRB L; and = fluid velocity appropriate for use in
the ADRE.
Fig. 3 illustrates the effect of the relative hydraulic conductiv-
ity on the PRB discharge for various aspect ratios a /b. For low
aspect ratios, as the relative hydraulic conductivity of the PRB
increases, the PRB exerts a “wicking” effect on the groundwater
flow field and the resulting velocity is approximately 20%
higher than the pre-PRB condition. However, as the PRB aspect
ratio increases, the asymptotic velocity approaches the regional
condition.
Conclusions
This paper presents a set of simple closed-form equations suitable
for the preliminary design of iron-based permeable reactive bar-
riers for commonly encountered scenarios e.g., Eqs. 4144
combined with Eqs. 1618兲兴. Sensitivity calculations indicate
that some of the assumptions commonly used in PRB analysis—
negligible dispersion, semi-infinite domain, and negligible flow
channeling—can lead to nonconservative designs.
By including dispersion and a finite reactive domain, the
model reflects a more realistic and conservative representation
than the commonly applied “batch” type equations. At the present
stage of development, the approach is applicable only to a single
reaction pathway. The understanding of PRB reaction chemistry
continues to evolve, and the single pathway approximation may
not be applicable to more complex systems containing other
contaminants such as chlorinated ethanes or other zero-valent
metals such as zinc e.g., Arnold and Roberts 1998, 2000.If
parallel reactions are included i.e., multiple parents contributing
to a reaction product, the common approach of neglecting
dispersion converts the problem into a set of coupled ordinary
differential equations that can be solved numerically using a stan-
dard ODE solver or, in some cases, analytically e.g., Eykholt
1999. However, for the decay of chlorinated ethylene by zero-
Fig. 2. Schematic of conceptual model for fully penetrating elliptical
permeable reactive barrier placed in homogeneous aquifer with
uniform regional flow
Fig. 3. Ratio of barrier flow rate to regional flow rate for various
permeable reactive barrier geometries, for barrier perpendicular to
regional flow. Contrast in conductivity k
e
/k and barrier aspect ratio
a /b control degree of difference between flux within barrier and
regional flow conditions
JOURNAL OF ENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2005 / 1593
valent iron, the current understanding of reaction pathways
e.g., ETI 2005b is consistent with the single pathway approach.
Sample calculations indicate that the predicted PRB perfor-
mance is quite sensitive to the assumed dispersion coefficient.
The potential influence of dispersion is well known in reactor
engineering e.g., Levenspiel 1999, but has not been a focus of
PRB-related research, perhaps due to the large safety factors
typically used to account for various parameter uncertainties.
Until actual measurements of dispersion coefficients are avail-
able, one promising approach would be to develop models that
explicitly account for dispersion in a probabilistic framework
such as Monte Carlo analysis e.g., Vidumsky and Landis 2001.
However, as the PRB technology matures, a reliable procedure
for estimating the PRB dispersivity could reduce the contribution
of dispersion to prediction uncertainty, potentially leading to
smaller safety factors.
The other contribution of this work is in the use of elliptical
analytic elements to compute the velocity through an idealized
PRB, in a manner that accounts for the potential increase in
flow through the barrier due to its higher hydraulic conductivity.
Future work will examine analytical approaches for analyzing
PRB hydraulics in more complex flow regimes.
Acknowledgments
This research was partially supported by Grant No. R82-7961
from the United States Environmental Protection Agency’s EPA
Science to Achieve Results STAR program. This paper has
not been subjected to any EPA review and therefore does not
necessarily reflect the views of the Agency, and no official
endorsement should be inferred. James Craig’s graduate study
was supported by the National Science Foundation Integrated
Graduate Education and Research Training IGERT program in
Geographic Information Science.
Appendix
Steady-State Solutions to One-Dimensional Multisolute
Advective–Dispersive–Reactive Equation
The application of the transformation developed by Sun et al.
1999 to the multisolute form of the advective–dispersive–
reactive equation Eq. 4兲兴 leads to
a
i
t
=−
a
i
x
+ D
2
a
i
x
2
ka
i
19
i =1,2,¯ ,n
To develop the solute-specific solutions, the entrance boundary
conditions for the reaction products are first transformed using
Eq. 10 as follows: For i =2
a
20
= c
20
+
y
2
k
1
k
1
k
2
c
10
20
For i=3
a
30
= c
30
+
j=1
2
l=j
2
y
i+1
k
l
k
l
k
3
c
j0
21
a
30
= c
30
+
y
2
k
1
k
1
k
3
冊冉
y
3
k
2
k
2
k
3
c
10
+
y
3
k
2
k
2
k
3
c
20
22
For i=4
a
40
= c
40
+
j=1
3
l=j
3
y
l+1
k
l
k
l
k
4
c
j0
23
a
40
= c
40
+
y
2
k
1
k
1
k
4
冊冉
y
3
k
2
k
2
k
4
冊冉
y
4
k
3
k
3
k
4
c
10
+
y
3
k
2
k
2
k
4
冊冉
y
4
k
3
k
3
k
4
c
20
+
y
4
k
3
k
3
k
4
c
30
24
The solutions for the solute concentration are then developed
using the back transformation given by Eq. 11. For i =2
c
2
= a
2
j=1
1
l=j
1
y
l+1
k
l
k
l
k
2
c
j
25
c
2
= a
2
y
2
k
1
k
1
k
2
c
1
26
For i=3
c
3
= a
3
j=1
2
l=j
2
y
l+1
k
l
k
l
k
3
c
j
27
c
3
= a
3
l=1
2
y
l+1
k
l
k
l
k
3
c
1
l=2
2
y
l+1
k
l
k
l
k
3
c
2
28
c
3
= a
3
y
2
k
1
k
1
k
3
冊冉
y
3
k
2
k
2
k
3
c
1
y
3
k
2
k
2
k
3
c
2
29
For i=4
c
4
= a
4
j=1
3
l=j
3
y
l+1
k
l
k
l
k
4
c
j
30
c
4
= a
4
l=1
3
y
l+1
k
l
k
l
k
4
c
1
l=2
3
y
l+1
k
l
k
l
k
4
c
2
l=3
3
y
l+1
k
l
k
l
k
4
c
3
31
c
4
= a
4
y
2
k
1
k
1
k
4
冊冉
y
3
k
2
k
2
k
4
冊冉
y
4
k
3
k
3
k
4
c
1
y
3
k
2
k
2
k
4
冊冉
y
4
k
3
k
3
k
4
c
2
y
4
k
3
k
3
k
4
c
3
32
For PRB design, it is useful to consider the simplified
case where the reaction products are not initially present
i.e., c
20
=c
40
=c
30
=0. For this case, the transformed boundary
conditions are
a
40
=
y
2
k
1
k
1
k
4
冊冉
y
3
k
2
k
2
k
4
冊冉
y
4
k
3
k
3
k
4
c
10
33
a
30
=
y
2
k
1
k
1
k
3
冊冉
y
3
k
2
k
2
k
3
c
10
34
a
20
=
y
2
k
1
k
1
k
2
c
10
35
1594 / JOURNAL OF ENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2005
Steady-State Solutions
The solutions given above are applicable to transient and steady-
state transport, for a variety of boundary conditions. For the
design of barrier systems, several steady-state solutions are
potentially applicable e.g., van Genuchten 1981.
For the constant concentration entrance condition Eq. 5兲兴 and
zero gradient exit condition Eq. 8兲兴 the steady-state solution to
Eq. 19 is
a
i
x = a
0i
exp
u
i
x
2D
+
u
i
u
i
+
exp
+ u
i
x
2D
u
i
L
D
1+
u
i
u
i
+
exp
u
i
L
D
36
where
u
i
=
2
+4k
i
D 37
For the constant concentration entrance condition and semi-
infinite exit Eq. 7兲兴, the solution is
a
i
= a
i0
exp
x
2
+4k
i
D
2D
38
If dispersion is ignored, the solution for the constant concentra-
tion entrance condition and semi-infinite exit is
a
i
= a
i0
exp
k
i
x
39
For the third-type entrance boundary condition and the semi-
infinite exit, the solution is
a
i
= a
0i
exp
u
i
x
2D
+
u
i
u
i
+
exp
+ u
i
x
2D
u
i
L
D
u
i
+
2
u
i
2
2u
i
+
exp
u
i
L
D
40
Expanded Solution for Permeable Reactive Barrier Design
Solutions for particular applications can be developed by combin-
ing the single-solute solutions given by Eqs. 3640 with the
multisolute expansions given by Eqs. 2035. The full
expansion is provided here for the recommended PRB design
Eq. 36 and the special case where reaction products are initially
absent Eqs. 3335. Application to other scenarios is straight-
forward. The resulting equations can be applied to PRB design by
evaluating the concentration at the barrier exit x=L .
c
1
= c
10
exp
u
1
x
2D
+
u
1
u
1
+
exp
+ u
1
x
2D
u
1
L
D
1+
u
1
u
1
+
exp
u
1
L
D
41
c
2
=
y
2
k
1
k
1
k
2
c
10
exp
u
2
x
2D
+
u
2
u
2
+
exp
+ u
2
x
2D
u
2
L
D
1+
u
2
u
2
+
exp
u
2
L
D
y
2
k
1
k
1
k
2
c
1
42
c
3
=
y
2
k
1
k
1
k
3
冊冉
y
3
k
2
k
2
k
3
c
10
exp
u
3
x
2D
+
u
3
u
3
+
exp
+ u
3
x
2D
u
3
L
D
1+
u
3
u
3
+
exp
u
3
L
D
y
3
k
2
k
2
k
3
c
2
y
2
k
1
k
1
k
3
冊冉
y
3
k
2
k
2
k
3
c
1
43
c
4
=
y
2
k
1
k
1
k
4
冊冉
y
3
k
2
k
2
k
4
冊冉
y
4
k
3
k
3
k
4
c
10
exp
u
4
x
2D
+
u
4
u
4
+
exp
+ u
4
x
2D
u
4
L
D
1+
u
4
u
4
+
exp
u
4
L
D
y
2
k
1
k
1
k
4
冊冉
y
3
k
2
k
2
k
4
冊冉
y
4
k
3
k
3
k
4
c
1
y
3
k
2
k
2
k
4
冊冉
y
4
k
3
k
3
k
4
c
2
y
4
k
3
k
3
k
4
c
3
44
Solution for Groundwater Flow Influenced
by Elliptical Element
The solution for groundwater flow influenced by an elliptical
inhomogeneity may be found in Strack 1989 or Obdam and
Veling 1987, and is a simple case of the general solution devel-
oped by Suribhatla et al. 2004. The development is based on the
AEM Strack 1989 and is applicable to steady-state horizontal
flow in a confined or unconfined aquifer with a uniform isotropic
regional hydraulic conductivity k and a single elliptical zone
of a different hydraulic conductivity k
e
. Uniform flow Q
0
is
oriented at an angle to the x axis as shown in Fig. 2. The
long and short semiaxes a ,b of the elliptical inhomogeneity are
oriented along the x and y axes, respectively. There is no recharge
anywhere in the model domain.
The complex potential
u
due to the uniform flow in elliptical
coordinates may be expressed as
u
=−Q
0
d cosh · e
i
+
0
45
where d =focal distance of the elliptical inhomogeneity; and
0
=potential at the reference location. The transformed spatial
coordinate is defined in terms of the local elliptical coordinates
Suribhatla et al. 2004 as
= + i 46
where = constant along ellipses with the same foci and
=constant along hyperbolas orthogonal to those ellipses.
The complex potential due to the elliptical inhomogeneity
itself is
JOURNAL OF ENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2005 / 1595
e
=
n=0
n
e
n
+ e
n
inside ␩⬍␩
0
n=1
n
n
e
2n
0
· e
n
outside ␩艌␩
0
47
The complex coefficients
n
=
n
R
+i
n
I
can be determined using
the following integrals:
0
=
k
e
k
4k
e
0
,d 48
n
R
=
k
e
k
2k coshn
0
+ k
e
sinhn
0
兲兴
e
0
, · cos nd
49
n
I
=
k
e
k
2k sinn
0
+ k
e
coshn
0
兲兴
e
0
, · sinnd
50
The quantity
e
0
, in the above equations is the potential
due to all the elements except for the inhomogeneity at the bound-
ary of the elliptical inhomogeneity where =
0
. For the case of
interest,
e
0
, equals the potential due to the uniform regional
flow only. Taking the real part of Eq. 45 and solving for the
coefficients yields
0
=
k
e
k
2k
0
51
1
R
=
k k
e
Q
0
d cosh
0
· cos
2k cosh
0
+ k
e
sinh
0
52
1
I
=
k k
e
Q
0
d sinh
0
· sin
2k sinh
0
+ k
e
cosh
0
53
Substituting the coefficients into Eq. 47 and combining the
influence of uniform flow, the complex potential at any point
inside the elliptical inhomogeneity is
=−dQ
0
cosh e
i
+
0
+2
0
+
1
e
+ e
兲共54
It is useful to consider the flow inside the PRB in terms of the
complex discharge function W, which is defined as the derivative
of the complex potential with respect to the complex Cartesian
coordinate z=x +iy Strack 1989
W =−
d
dz
55
The discharge function inside the ellipse becomes
W = Q
0
e
i
+
1
a sinh
n=1
n
n
e
␩␶
e
␩␶
= Q
0
e
i
+
−2
d
1
R
+ i
1
I
= Q
0
k
e
kcosh
0
cos
k cosh
0
+ k
e
sinh
0
i
k
e
ksinh
0
cos
k sinh
0
+ k
e
cosh
0
+ e
i
56
This expression for complex discharge can be simplified using the
following geometric relationships:
a = d cosh
0
57a
b = d sinh
0
57b
leading to an expression for the discharge within the ellipse
W =
Q
0
k
e
a + bcos
ak + bk
e
i
Q
0
k
e
a + bsin
ak
e
+ bk
58
The magnitude of flow inside the elliptical inhomogeneity is
given as the absolute value of Eq. 58
Q = a + bQ
0
k
e
2
sin
2
bk + ak
e
2
+
k
e
2
cos
2
ak + bk
e
2
1/2
59
When the ratio of the conductivities of the ellipse to the back-
ground k
e
/k goes to infinity highly conductive barrier,
the total flow inside the ellipse reaches the value as given by
Q = a + bQ
0
sin
2
a
2
+
cos
2
b
2
1/2
60
For the case of flow perpendicular to the PRB =9,
the discharge becomes
Q = a + bQ
0
k
e
bk + ak
e
61
References
Arnold, W. A., and Roberts, A. L. 1998. “Pathways of chlorinated
ethylene and chlorinated acetylene reduction with Zn0.” Environ.
Sci. Technol.,3219, 3017–3025.
Arnold, W. A., and Roberts, A. L. 2000. “Pathways and kinetics of
chlorinated ethylene and chlorinated acetylene reaction with Fe0
particles.” Environ. Sci. Technol.,349, 1794–1805.
Aziz, C. E., Newell, C. J., Gonzales, J. R., Haas, P., Clement, T. P., and
Sun, Y. 2000. BIOCHLOR: Natural attenuation decision support
system, version 1.0, users manual, EPA/600/R-00/008, United States
Environmental Protection Agency, Ada, Okla.
Casey, F. X. M., Ong, S. K., and Horton, R. 2000. “Degradation
and transformation of trichloroethylene in misciple-displacement
experiments through zero-valent metals.” Environ. Sci. Technol., 34,
5023–5029.
Craig, J. R., and Matott, L. S. 2004. Visual Bluebird user’s manual:
Version 1.8, Dept. of Civil, Structural, and Environmental Engineer-
ing, Univ. at Buffalo, Buffalo, N.Y. http://groundwater.buffalo.edu
Das, D. B., 2002. “Hydrodynamic modelling for groundwater
flow through permeable reactive barriers.” Hydrolog. Process., 16,
3394–3418.
Environmental Technologies Inc. ETI. 2005a.http://www.eti.ca,
accessed March 28, 2005, ETI, Waterloo, Ont., Canada.
Environmental Technologies Inc. ETI. 2005b. “First-order kinetic
degradation models.” Technical Note 2.06, ETI, Waterloo, Ont.,
Canada.
Eykholt, G. R. 1999
. “Analytical solution for networks of irreversible
first-order reactions.” Water Res.,333, 814826.
Eykholt, G. R., and Sivavec, T. M. 1995. “Contaminant transport issues
for reactive-permeable barriers.” Proc., Geoenvironment 2000,
Characterization, Containment, Remediation, and Performance in
Environmental Geotechnics, Vol. 2, Y. B. Acar and D. E. Daniel, eds.,
ASCE, New York, 1608–1621.
Farrell, J., Kason, M., Meitas, N., and Li, T. 2000. “Investigation of the
long-term performance of zero-valent iron for reductive dechlorina-
tion of trichloroethylene.” Environ. Sci. Technol.,343, 514–521.
Fitts, C., 1997. “Analytic modeling of impermeable and resistant
1596 / JOURNAL OF ENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2005
barriers.” Ground Water,352, 312–317.
Gavaskar, A. R., Gupta, N., Sass, B., Fox, T., Janosy, R. J., Cantrell, K.,
and Olfenbuttel, R. 1997. “Design guidance for application of
Permeable barriers to remediate dissolved chlorinated solvents.”
Report Prepared for the United States Army Corps of Engineers,
Battelle, Columbus, Ohio.
Gavaskar, A. R., Gupta, N., Sass, B., Janosy, R. J., and Hicks, J. 2000.
“Design guidance for application of permeable reactive barriers for
groundwater remediation.” Report Prepared for the United States Air
Force, Battelle, Columbus, Ohio.
Gavaskar, A. R., Gupta, N., Sass, B. M., Janosy, R. J., and O’Sullivan, D.
1998. Permeable barriers for groundwater remediation, Battelle,
Columbus, Ohio.
Gelhar, L. W., Welty, C., and Rehfeldt, K. R. 1992. “A critical review of
data on field-scale dispersion in aquifers.” Water Resour. Res.,287,
1955–1974.
Gupta, N., and Fox, T. C. 1999. “Hydrogeologic modeling for
permeable reactive barriers.” J. Hazard. Mater., 68, 19–39.
Levenspiel, O., 1999. Chemical reaction engineering, 3rd Ed., Wiley,
New York.
Obdam, A. N. M., and Veling, E. J. M. 1987. “Elliptical inhomogenities
in groundwater flow-An analytical description.” J. Hydrol., 95,
87–96.
Roberts, L. A., Totten, L. A., Arnold, W. A., Burris, D. R.,
and Campbell, T. J. 1996. “Reductive elimination of chlorinated
ethylenes by zero-valent metals.” Environ. Sci. Technol.,308,
2654–2659.
Scherer, M. M., Richter, S., Valentine, R. L., and Alvarez, P. J. J. 2000.
“Chemistry and microbiology of permeable reactive barriers for in
situ groundwater clean up.” Crit. Rev. Environ. Sci. Technol.,303,
363–411.
Sivavec, T. M., Horney, D. P., Mackenzie, P. D., and Salvo, J. J. 1996
.
“Zero-valent iron treatability study for groundwater contaminated
with chlorinated organic solvents at the Paducah Gaseous Diffusion
Plant Site.” Environmental Laboratory Rep. No. 96CRD040,GE
Research and Development Center, Schenectady, N.Y.
Starr, R. C., and Cherry, J. A. 1994. “In situ remediation of contami-
nated ground water: The funnel-and-gate system.” Ground Water,
323, 465–476.
Strack, O. D. L., 1989. Groundwater mechanics, Prentice-Hall,
Englewood Cliffs, N.J.
Sun, Y., Peterson, J. N., Clement, T. P., and Skeen, R. S. 1999. “Devel-
opment of analytical solutions for multispecies transport with serial
and parallel reactions.” Water Resour. Res.,351, 185–190.
Suribhatla, R., Bakker, M., Bandilla, K., and Jankovic, I. 2004. “Steady
two-dimensional groundwater flow through many elliptical inhomo-
geneities.” Water Resour. Res.,404, WO4202.
Tratnyek, P. G., Johnson, T. L., Scherer, M. M., and Eykholt, G. R.
1997. “Remediating ground water with zero-valent metals:
Chemical considerations in barrier design.” Ground Water Monit.
Rem.,174, 108–114.
United States Environmental Protection Agency USEPA. 1998.
“Permeable reactive barrier technologies for contaminant remedia-
tion.” EPA/600/R-98/125, Office of Solid Waste and Emergency
Response, USEPA, Washington, D.C.
van Genuchten, M. Th., 1981. “Analytical solutions for chemical
transport with simultaneous adsorption, zero-order production and
first-order decay.” J. Hydrol., 49, 213–233.
Vidumsky, J. E., and Landis, R. C. 2001. “Probabilistic design of a
combined permeable barrier and natural biodegradation study.”
Presented at the Proc., 2001 Int. Containment and Remediation
Technology Conf., Orlando, Fla.
Yabusaki, S., Cantrell, K., Sass, B., and Steefel, C. 2001
.
“Multi-component reactive transport in an in situ zero-valent iron
cell.” Environ. Sci. Technol.,357, 1493–1503.
JOURNAL OF ENVIRONMENTAL ENGINEERING © ASCE / NOVEMBER 2005 / 1597
... The kinetics of the different reactive materials is well understood and documented. While a number of different reactive materials have been used, most of the PRBs installed worldwide utilize zerovalant iron (ZVI) as the reactive material (Rabideau et al., 2005). An overview of hydrogeological modeling for PRBs is given in Gupta and Fox (1999). ...
... Estimates of various parameters for designing the PRB are pro vided in the following sections. A simple mathematical model (Rabideau et al., 2005) governing the trans port process through the reactive barrier is represented by a one-dimensional advective-dispersive-reactive equation (ADRE). The governing equation for the single decay ADRE assuming a homogenous subsurface medium is as follows: ...
Chapter
Full-text available
Permeable reactive barrier (PRB) technology is an increasingly viable option for remediating chlorinated hydrocarbon, petroleum hydrocarbon, and dissolved heavy metals contamination (Chapters 2 and 3). The PRB is an in situ passive remediation technology and has certain advantages compared to other active remediation technologies such as the pump-and-treat and chemical oxidation. This technology also prevents the contamination from migrating to uncontaminated aquifer systems. About 200 PRBs have been installed worldwide (Das, 2002; ETI, 2005, see Chapter 3) for treating common contaminants like chlorinated hydrocarbons (Burris et al., 1995; Orth and Gillham, 1996; Roberts et al., 1996; McMahon et al., 1999; Vogan et al., 1999; Schlicker et al., 2000), petroleum hydrocarbons (Guerin et al., 2002) and heavy metals (Powell et al., 1995; Gu et al., 1998; Shokes and Möller, 1999). A schematic diagram demonstrating the PRB technology is shown in Figure 4.1. The most important components of the design and implementation of the PRB are a detailed understanding of the subsurface hydrogeology, the kinetics of the reactive material chosen for the barrier, and the long-term monitoring plan. The kinetics of the different reactive materials is well understood and documented. While a number of different reactive materials have been used, most of the PRBs installed worldwide utilize zerovalant iron (ZVI) as the reactive material (Rabideau et al., 2005). An overview of hydrogeological modeling for PRBs is given in Gupta and Fox (1999). The most challenging component of PRB design and implementation is the site hydraulics, and several case studies of PRBs demonstrate this aspect of the technology.
... The kinetics of the different reactive materials is well understood and documented. While a number of different reactive materials have been used, most of the PRBs installed worldwide utilize zerovalant iron (ZVI) as the reactive material (Rabideau et al., 2005). An overview of hydrogeological modeling for PRBs is given in Gupta and Fox (1999). ...
... Estimates of various parameters for designing the PRB are pro vided in the following sections. A simple mathematical model (Rabideau et al., 2005) governing the trans port process through the reactive barrier is represented by a one-dimensional advective-dispersive-reactive equation (ADRE). The governing equation for the single decay ADRE assuming a homogenous subsurface medium is as follows: ...
Book
Full-text available
Remediation of groundwater is complex and often challenging. But the cost of pump and treat technology, coupled with the dismal results achieved, has paved the way for newer, better technologies to be developed. Among these techniques is permeable reactive barrier (PRB) technology, which allows groundwater to pass through a buried porous barrier that either captures the contaminants or breaks them down. And although this approach is gaining popularity, there are few references available on the subject. Until now. Permeable Reactive Barrier: Sustainable Groundwater Remediation brings together the information required to plan, design/model, and apply a successful, cost-effective, and sustainable PRB technology. With contributions from pioneers in this area, the book covers state-of-the-art information on PRB technology. It details design criteria, predictive modeling, and application to contaminants beyond petroleum hydrocarbons, including inorganics and radionuclides. The text also examines implementation stages such as the initial feasibility assessment, laboratory treatability studies (including column studies), estimation of PRB design parameters, and development of a long-term monitoring network for the performance evaluation of the barrier. It also outlines the predictive tools required for life cycle analysis and cost/performance assessment. A review of current PRB technology and its applications, this book includes case studies that exemplify the concepts discussed. It helps you determine when to recommend PRB, what information is needed from the site investigation to design it, and what regulatory validation is required.
... The kinetics of the different reactive materials is well understood and documented. While a number of different reactive materials have been used, most of the PRBs installed worldwide utilize zerovalant iron (ZVI) as the reactive material (Rabideau et al., 2005). An overview of hydrogeological modeling for PRBs is given in Gupta and Fox (1999). ...
... Estimates of various parameters for designing the PRB are pro vided in the following sections. A simple mathematical model (Rabideau et al., 2005) governing the trans port process through the reactive barrier is represented by a one-dimensional advective-dispersive-reactive equation (ADRE). The governing equation for the single decay ADRE assuming a homogenous subsurface medium is as follows: ...
Chapter
Full-text available
Contaminated sites represent a major challenge for the long-term sustainability of the environment. In addition to their potential adverse impacts on human health, surface and groundwater quality, and ecological processes, they also represent a lost economic opportunity. Sources of contaminants include those arising from anthropogenic activities such as industrial and agricultural practices, mining activities, accidental spillages, and so on (Barzi et al., 1996; Naidu, 1996), and natural geogenic processes (Naidu et al., 2006), with the latter largely associated with metals and metalloids such as arsenic, lead, cadmium, and mercury. Irrespective of the source of contaminants, they will interact with soil colloidal particles and moisture in the near-surface leachable zone. They can, therefore, be dissolved/solubilized into water infiltrating through any unsaturated zone present in the soil profile. They can penetrate below the water table and subsequently migrate laterally in flowing groundwater and transported off-site, thus posing a serious risk to groundwater quality.
... In some cases, mathematical/ numerical modelling has been performed to predict the long-term performance of PRBs, aid the evaluation of the materials and the design of the PRB as well as elucidate the contaminant removal mechanisms. For example, Rabideau et al. (2005) employed the advective-dispersivereactive equation (ADRE) with a zero concentration gradient imposed on the exit boundary of the PRB to determine the PRB thickness. Obiri-Nyarko et al. (2015) performed geochemical modelling using PHREEQC to predict the long-term performance of a zeolite-PRB to treat leadcontaminated groundwater. ...
Article
Permeable adsorptive barriers (PABs) consisting of individual (compost, zeolite, and brown coal) and composite (brown coal-compost and zeolite-compost) adsorbents were evaluated for their hydraulic performance and effectiveness in removing aqueous benzene using batch and column experiments. Different adsorption isotherms and kinetic models and different formulations of the equilibrium advection-dispersion equation (ADE) were evaluated for their capabilities to describe the benzene sorption in the media. The batch experiments showed that the adsorption of benzene by the adsorbents was favourable and could be adequately described by the Freundlich and Langmuir isotherms and the pseudo-second-order kinetic model. Particle attrition and structural reorganization occurred in the columns, possibly introducing preferential flow paths and resulting in slight changes in the final hydraulic conductivity values (4.3 × 10⁻⁵ cm s⁻¹–1.7 × 10⁻³ cm s⁻¹) relative to the initial values (4.2 × 10⁻⁵ cm s⁻¹–2.14 × 10⁻³ cm s⁻¹). Despite the fact that preferential flow appeared to have an impact on the performance of the investigated adsorbents, the brown coal-compost mixture proved to be the most effective adsorbent. It significantly delayed benzene breakthrough (R = 29), indicating that it can be applied as a low-cost effective adsorbent in PABs for sustainable remediation of benzene-contaminated groundwater. The formulated transport models could fairly describe the behaviour of benzene in the investigated media under dynamic flow conditions; however, model refinement and additional experimental studies are needed before pilot/full-scale applications to improve the fits and verify the benzene removal processes. Our results generally demonstrate how such studies can be useful in evaluating potential reactive barrier materials.
Article
Permeable reactive barriers (PRBs) represent a recent in situ technology for groundwater remediation. The simplest configuration consists of a permeable diaphragm filled with a reactive material and placed mainly perpendicularly to the groundwater flow in order to immobilize or transform the contaminants through chemical, physical or biological mechanisms. The filtration mechanism through the barrier is mainly due to the hydraulic gradient therefore no external energy is required (passive technology). In this paper the main installation methods, configurations and techniques used for PRB construction, their applications, the physical, chemical and geotechnical characteristics of the reactive media commonly used for PRB realization, with specific reference to zero valent iron (Fe0), and the knowledge necessary for their correct design, in the light of the most recent researches available in the literature, are described. Furthermore, this paper describes the main results of an extensive experimental study aimed at evaluating the efficiency of Fe0 and pumice mixtures to be used in PRBs for the remediation of groundwater contaminated by heavy metals like nickel, copper and zinc. The results of such researches allowed to evaluate the removal efficiency and the hydraulic behavior of the reactive media, in the short and long period, by varying flow velocity and contaminants concentration. The results allowed also to evaluate the reliability of accelerated column tests for PRB design.
Article
The ‘sustainable remediation’ concept has been broadly embraced by industry and governments in recent years in both the US and Europe. However, there is a strong need for more research to enhance its ‘practicability’. In an attempt to fill this research gap, this study developed a generalised framework for selecting the most environmentally sustainable remedial technology under various site conditions. Four remediation technologies were evaluated: pump and treat (P&T), enhanced in situ bioremediation (EIB), permeable reactive barrier (PRB), and in situ chemical reduction (ISCR). Within the developed framework and examined site condition ranges, our results indicate that site characteristics have a profound effect on the life cycle impact of various remedial alternatives, thus providing insights and valuable information for determining what is considered the most desired remedy from an environmental sustainability perspective.
Article
We present a set of new, semi-analytical solutions to simulate three-dimensional contaminant transport subject to first-order chain-decay reactions. The aquifer is assumed to be areally semi-infinite, but finite in thickness. The analytical solution can treat the transformation of contaminants into daughter products, leading to decay chains consisting of multiple contaminant species and various reaction pathways. The solution in its current form is capable of accounting for up to seven species and four decay levels. The complex pathways are represented by means of first-order decay and production terms, while branching ratios account for decay stoichiometry. Besides advection, dispersion, bio-chemical or radioactive decay and daughter product formation, the model also accounts for sorption of contaminants on the aquifer solid phase with each species having a different retardation factor. First-type contaminant boundary conditions are utilized at the source (x=0 m) and can be either constant-in-time for each species, or the concentration can be allowed to undergo first-order decay. The solutions are obtained by exponential Fourier, Fourier cosine and Laplace transforms. Limiting forms of the solutions can be obtained in closed form, but we evaluate the general solutions by numerically inverting the analytical solutions in exponential Fourier and Laplace transform spaces. Various cases are generated and the solutions are verified against the HydroGeoSphere numerical model.
Article
Full-text available
1] Transient semianalytical and steady state analytical solutions to assess solute transport in a permeable reactive barrier (PRB)–aquifer system are developed on the basis of a multidomain approach of an up-gradient PRB and a down-gradient aquifer. Sensitivity analysis of input parameters in the solution is addressed through mathematical modeling. A designing equation from the developed steady state closed-form analytical solution is derived, and the accuracy of the equation is tested. The results confirm that the thickness and the first-order reaction capability of the PRB are two of the most important design considerations. Through sensitivity analysis, the reduction of the reactive capability of the PRB has a significant impact on the PRB performance. The aquifer reaction capability is found to be nonnegligible for the determination of the required thickness of the PRB, and its importance becomes greater when the PRB porosity becomes smaller because of mineral fouling. Citation: Park, E., and H. Zhan (2009), One-dimensional solute transport in a permeable reactive barrier – aquifer system, Water Resour. Res., 45, W07502, doi:10.1029/2008WR007155.
Article
The permeable reactive barrier (PRB) remediation technology has proven to be more cost-effective than conventional pump-and-treat systems, and has demonstrated the ability to rapidly reduce the concentrations of specific chemicals of concern (COCs) by up to several orders of magnitude in some scenarios. This study derives new steady-state analytical solutions to multispecies reactive transport in a PRB-aquifer (dual domain) system. The advantage of the dual domain model is that it can account for the potential existence of natural degradation in the aquifer, when designing the required PRB thickness. The study focuses primarily on the steady-state analytical solutions of the tetrachloroethene (PCE) serial degradation pathway and secondly on the analytical solutions of the parallel degradation pathway. The solutions in this study can also be applied to other types of dual domain systems with distinct flow and transport properties. The steady-state analytical solutions are shown to be accurate and the numerical program RT3D is selected for comparison. The results of this study are novel in that the solutions provide improved modeling flexibility including: 1) every species can have unique first-order reaction rates and unique retardation factors, and 2) daughter species can be modeled with their individual input concentrations or solely as byproducts of the parent species. The steady-state analytical solutions exhibit a limitation that occurs when interspecies reaction rate factors equal each other, which result in undefined solutions. Excel spreadsheet programs were created to facilitate prompt application of the steady-state analytical solutions, for both the serial and parallel degradation pathways.
Article
Full-text available
A new analytic element solution has been derived for steady two-dimensional groundwater flow through an aquifer that contains an arbitrary number of elliptical inhomogeneities. The hydraulic conductivity of each inhomogeneity is homogeneous and differs from the conductivity of the homogeneous background. In addition to elliptical inhomogeneities, other elements (such as wells and line sinks) may be present. The method is based on a separable form of the solution for Laplace's differential equation in elliptical coordinates. The piezometric head and the stream function, expressed as continuous spatial functions (as components of the complex potential), may be obtained up to machine accuracy regardless of the shape, size, orientation, and conductivity of the elliptical inhomogeneities. Components of the discharge vector are expressed in a similar manner, using a complex discharge function. Problems with 10,000 or more inhomogeneities can be solved using parallel computing on distributed memory supercomputer clusters. Two examples are included to demonstrate the precision and capabilities of the method. The second example is used to perform a preliminary study of contaminant transport in a highly heterogeneous formation of lognormal conductivity distribution. The results of the transport study are compared with recent theoretical and numerical results that are based on circular inhomogeneities. The new results with elliptical inhomogeneities confirm the findings based on circular inhomogeneities, including a long dispersion setting time and a zero value of the asymptotic transverse dispersivity.
Article
Several research institutes have become interested in the concept of permeable barriers for the containment and/or destruction of contaminated groundwater. The purpose of these trench-like barriers is to provide in situ capture and possibly destruction of the contaminant while preserving groundwater flow to uncontaminated zones. For instance, a trichloroethylene (TCE) plume may be contained by a permeable barrier in which reactive iron reduces TCE to ethylene and ethane, compounds which can be easily biodegraded (Gillham and O'Hannesin (1994)). The permeable barrier technology may also be implemented by coupling two or more of the following processes: bioremediation, sorption, reactive metal reductive dechlorination, oxidation, electrokinetics, or other processes. Several possible configurations of permeable barriers will be discussed. A few laboratory results will be presented, and applicable analytical solutions from van Genuchten (1981) will be used to suggest field-scale performance of reactive metal permeable barriers for the reductive dechlorination of TCE.
Article
Permeable reactive barriers (PRBs) are receiving a great deal of attention as an innovative, cost-effective technology for in situ clean up of groundwater contamination. A wide variety of materials are being proposed for use in PRBs, including zero-valent metals (e.g., iron metal), humic materials, oxides, surfactant-modified zeolites (SMZs), and oxygen- and nitrate- releasing compounds. PRB materials remove dissolved groundwater contaminants by immobilization within the barrier or transformation to less harmful products. The primary removal processes include: (1) sorption and precipitation, (2) chemical reaction, and (3) biologically mediated reactions. This article presents an overview of the mechanisms and factors controlling these individual processes and discusses the implications for the feasibility and long-term effectiveness of PRB technologies.
Article
The funnel-and-gate system for in situ treatment of contaminant plumes consists of low hydraulic conductivity cutoff walls with gaps that contain in situ reactors, such as reactive porous media-, that remove contaminants by abiotic or biological processes. Funnel-and-gate systems can be installed at the front of plumes to prevent further plume growth, or immediately downgradient of contaminant source zones to prevent contaminants from moving into plumes. Cutoff walls (the funnel) modify flow patterns so that ground water flows primarily through high conductivity gaps (the gates). This approach is largely passive in that after installation, in situ reactors are intended to function with little or no maintenance for long periods. This approach contrasts with the energy and maintenance-intensive character of pump-and-treat systems. This paper describes the funnel-and-gate concept, and uses two-dimensional computer simulations to illustrate the effects of cutoff wall and gate configuration on capture zone size and shape and on the residence time for reaction of contaminants in gates.
Article
To date it does not appear to have been demonstrated in the literature that halogenated ethylenes can undergo reductive {beta}-elimination to alkynes under environmental conditions. The purpose of this paper is to provide experimental evidence that such pathways may be involved in the reaction of chloroethylenes with zero-valent metals as well as to speculate on the significance of the products that may result. Calculations indicate that reductive {beta}-elimination reactions of chloroethylenes are in fact comparable energetically to hydrogenolysis at neutral pH. Experiments were therefore initiated to assess whether {beta}-elimination reactions of chlorinated ethylenes could occur in the presence of two zero-valent metals, Fe and Zn. 76 refs., 3 figs., 1 tab.
Article
Pathways and kinetics through which chlorinated ethylenes and their daughter products react with Fe(O) particles were investigated through batch experiments. Substantial intra- and interspecies inhibitory effects were observed, requiring the use of a modified Langmuir-Hinshelwood-Hougen-Watson (LHHW) kinetic model in which species compete for a limited number of reactive sites at the particle-water interface. Results indicate that reductive β-elimination accounts for 87% of tetrachloroethylene (PCE), 97% of trichloroethylene (TCE), 94% of cis-dichloroethylene (cis-DCE), and 99% of trans-dichloroethylene (Trans-DCE) reaction. Reaction of 1,1-DCE gives rise to ethylene, consistent with a reductive α-elimination pathway. For the highly reactive chloro- and dichloro-acetylene intermediates produced from the reductive elimination of TCE and PCE, 100% and 76% of the reaction, respectively, occur via hydrogenolysis to lessen chlorinated acetylenes. The branching ratios for reactions of PCE or TCE (and their daughter products) with iron particles are therefore such that production of vinyl chloride is largely circumvented. Reactivity of the chlorinated ethylenes decreases markedly with increasing halogenation, counter to the trend that might be anticipated if the rate-limiting step were to involve dissociative electron transfer. The authors propose that the reaction of vinyl halides proceeds via a di-Ï-bonded surface-bound intermediate. The reactivity trends and pathways observed in this work explain why lesser-chlorinated ethylenes have only been reported as minor products in prior laboratory and field studies of PCE and TCE reaction with Fe(O).
Article
A new analytical solution technique for networks of irreversible chemical reactions in batch reactors with first-order kinetics is presented. The solution technique involves the iterative calculation of a coefficient matrix that is then applied for direct calculation of species concentrations at any time. The derivation, which is based on Laplace transforms of an arbitrary network, is presented in the appendix. Several examples and extensions to the solution are discussed. Extensions made to the solution include the consideration of phase equilibria, first-order and zero-order sources and sinks, common reaction products outside the network, and combined networks of reactions and CSTRs. Rate-limited mass transfer effects and reversible reactions were not considered. The solution is general, compact, easy to use, and allows several extensions that will likely aid scientists and engineers in solving problems related to the dynamics of these networks.
Article
In order to be able to decide whether and how heterogeneities should be taken into account for modelling a groundwater flow problem, the magnitude of those disturbances must be estimated. Therefore, in this paper we solve the problem of a uniform flow in a homogeneous medium with an elliptical shaped inhomogeneity with a different value for the permeability and positioned under an arbitrary angle with respect to the undisturbed flow. We calculate the ratio of the flow rate through this body with respect to the undisturbed case. Besides its use in solving groundwater flow problems, this formula may be helpful for problems relevant to field reconnaisance.